quant-ph0505213/dqc.tex
1: % STF draft of sections 4,5. Feb 7, 2005.
2: 
3: % AD draft of sections 1,2,3. Revisions of
4: % sections 4,5. Feb 8, 2005.
5: 
6: % STF Abstract added, figures & captions added
7: % with accompanying changes in the text.
8: % Proofread sections 4,5.  Feb 11, 2005.
9: 
10: % AD edit of section 2.  Mar 7 2005
11: 
12: % STF edit of sections 4,5.
13: % Rewrote section 3.
14: % Reformatting, including rearrangement
15: % of sections and addition of appendix.
16: % Conclusion added.
17: % Mar 11 2005
18: 
19: % AD and STF Fuchsian analysis of the introduction.
20: % Mar 14 2005
21: 
22: % CMC edit of intro
23: % Apr 12 2005
24: 
25: % CMC rewrite of Secs. I-IV. May 18, 2005.
26: 
27: % AD added stuff on asymptopia.
28: % STF fixed all the figures.  May 19, 2005.
29: 
30: % CMC final draft. May 26, 2005.
31: 
32: % Absolute final.  May 26, 2005.
33: 
34: \documentclass[pra,eqsecnum,amsmath,amsfonts,12pt,letterpaper,tightenlines,nofootinbib]{revtex4}
35: 
36: \usepackage{hyperref,graphicx}
37: 
38: \input{Qcircuit}
39: 
40: \newcommand{\braket}[2]{\langle #1 \vert #2 \rangle}
41: \newcommand{\abs}[1]{\lvert {#1} \rvert}
42: \def\ket#1{\lvert #1 \rangle}
43: \def\bra#1{{\langle #1 \rvert  }}
44: \def\tr{{\rm{tr}}}
45: \def\spec{{\rm{Spec}}}
46: 
47: \def\ba{\mathbf{a}}
48: \def\bb{\mathbf{b}}
49: \def\bc{\mathbf{c}}
50: \def\bx{\mathbf{x}}
51: \def\mz{\mathbb{Z}}
52: 
53: \def\sN{\mathcal{N}}
54: \def\sM{\mathcal{M}}
55: \def\sC{\mathcal{C}}
56: \def\sT{\mathcal{T}}
57: \def\sR{\mathcal{R}}
58: 
59: \newcommand{\be}{\begin{equation}}
60: \newcommand{\ee}{\end{equation}}
61: \newcommand{\ben}{\begin{eqnarray}}
62: \newcommand{\een}{\end{eqnarray}}
63: \newcommand{\bes}{\begin{subequations}}
64: \newcommand{\ees}{\end{subequations}}
65: \newcommand{\bF}{\begin{figure}}
66: \newcommand{\eF}{\end{figure}}
67: \newcommand{\dg}{\dagger}
68: 
69: 
70: \def\ket#1{ | #1 \rangle}
71: \def\bra#1{{\langle #1 |  }}
72: 
73: \begin{document}
74: 
75: \title{Entanglement and the Power of One Qubit}
76: \author{Animesh Datta}
77: \email{animesh@unm.edu}
78: \author{Steven T. Flammia}
79: \email{sflammia@unm.edu}
80: \author{Carlton M. Caves}
81: \email{caves@info.phys.unm.edu}
82: \affiliation{Department of Physics and Astronomy, University of New Mexico,
83:     Albuquerque, NM 87131}
84: \date{\today}
85: 
86: \begin{abstract}
87: The ``Power of One Qubit'' refers to a computational model that has
88: access to only one pure bit of quantum information, along with $n$
89: qubits in the totally mixed state.  This model, though not as powerful
90: as a pure-state quantum computer, is capable of performing some
91: computational tasks exponentially faster than any known classical
92: algorithm.  One such task is to estimate with fixed accuracy the
93: normalized trace of a unitary operator that can be implemented
94: efficiently in a quantum circuit.  We show that circuits of this type
95: generally lead to entangled states, and we investigate the amount of
96: entanglement possible in such circuits, as measured by the
97: multiplicative negativity.  We show that the multiplicative negativity
98: is bounded by a constant, independent of $n$, for all bipartite
99: divisions of the $n+1$ qubits, and so becomes, when $n$ is large, a
100: vanishingly small fraction of the maximum possible multiplicative
101: negativity for roughly equal divisions.  This suggests that the global
102: nature of entanglement is a more important resource for quantum
103: computation than the magnitude of the entanglement.
104: \end{abstract}
105: 
106: \maketitle
107: 
108: \section{Introduction}
109: \label{S:intro}
110: 
111: Fully controllable and scalable quantum computers are likely many years
112: from realization.  This motivates study and development of somewhat
113: less ambitious quantum information processors, defined as devices that
114: fail to satisfy one or more of DiVincenzo's five criteria for a quantum
115: computer~\cite{DiVincenzo00}.   An example of such a quantum
116: information processor is a mixed-state quantum system, which fails to
117: pass DiVincenzo's second requirement, that the system be prepared in a
118: simple initial state.
119: 
120: The prime example of a mixed-state quantum information processor is
121: provided by liquid-state NMR experiments in quantum information
122: processing~\cite{Jones01}.  Current NMR experiments, which operate with
123: initial states that are highly mixed thermal states, use a technique
124: called pseudo-pure-state synthesis to process the initial thermal state
125: and thereby to simulate pure-state quantum information processing. This
126: technique suffers from an exponential loss of signal strength as the
127: number of qubits per molecule increases and thus is not scalable. There
128: is a different technique for processing the initial thermal state,
129: called algorithmic cooling~\cite{Schulman99}, which pumps entropy from
130: a subset of qubits into the remaining qubits, leaving the special
131: subset in a pure state and the remaining qubits maximally mixed.
132: Algorithmic cooling provides an in-principle method for making
133: liquid-state NMR---or any qubit system that begins in a thermal
134: state---scalable, in essence by providing an efficient algorithmic
135: method for cooling a subset of the initially thermal qubits to a pure
136: state, thereby satisfying DiVincenzo's second criterion.
137: 
138: Knill and Laflamme~\cite{kl98} proposed a related mixed-state
139: computational model, which they called DQC1, in which there is just one
140: initial pure qubit, along with $n$ qubits in the maximally mixed state.
141: Although provably less powerful than a pure-state quantum
142: computer~\cite{Ambainis00}, DQC1 can perform some computational tasks
143: efficiently for which there are no known polynomial time classical
144: algorithms.  In particular, a DQC1 quantum circuit can be used to
145: evaluate, with fixed accuracy independent of $n$, the normalized trace,
146: $\tr(U_n)/2^n$, of any $n$-qubit unitary operator $U_n$ that can be
147: implemented efficiently in terms of quantum
148: gates~\cite{kl98,Laflamme02}.  In Sec.~\ref{S:classical} we consider
149: briefly whether there might be efficient classical algorithms for
150: estimating the normalized trace, and we conclude that this is unlikely.
151: The efficient quantum algorithm for estimating the normalized trace
152: provides an exponential speedup over the best known classical algorithm
153: for simulations of some quantum processes~\cite{pklo04,elpc04}.  Knill
154: and Laflamme referred to the power of this mixed-state computational
155: model as the ``power of one qubit.''
156: 
157: Study of the power of one qubit is motivated partly by NMR experiments,
158: but our primary motivation in this paper is to investigate the role of
159: entanglement in quantum computation, using DQC1 as a theoretical test
160: bed for the investigation.  For pure-state quantum computers, Jozsa and
161: Linden \cite{Josza99} have shown that exponential speedup over a
162: classical computer requires that entanglement not be restricted to
163: blocks of qubits of fixed size as problem size increases. Entanglement
164: that increases with problem size is thus a necessary prerequisite for
165: the exponential speedup achieved by a pure-state quantum computer.  On
166: the other hand, the Gottesman-Knill theorem \cite{N&C}, demonstrates
167: that global entanglement is far from sufficient for exponential
168: speedup.  While this means that the role of entanglement is not
169: entirely understood for pure-state quantum computers, far less is known
170: about the role of entanglement in mixed-state quantum computers. When
171: applied to mixed-state computation, the Jozsa-Linden proof does not
172: show that entanglement is a requirement for exponential speedup.
173: Indeed, it has not previously been shown that there is any entanglement
174: in the DQC1 circuits that provide an exponential speedup over classical
175: algorithms.
176: 
177: The purpose of this paper is to investigate the existence of and amount
178: of entanglement in the DQC1 circuit that is used to estimate the
179: normalized trace.  The DQC1 model consists of a \emph{special qubit\/}
180: (qubit~0) in the initial state $|0\rangle\langle0|={1\over2}(I_1+Z)$,
181: where $Z$ is a Pauli operator, along with $n$ other qubits in the
182: completely mixed state, $I_n/2^n$, which we call the \emph{unpolarized
183: qubits}.  The circuit consists of a Hadamard gate on the special qubit
184: followed by a controlled unitary on the remaining
185: qubits~\cite{Laflamme02}: \vspace{-1em}
186: \begin{equation}
187: \label{E:circuit}
188: \Qcircuit @C=.5em @R=-.5em {
189:     & \lstick{\ket{0}\!\bra{0}} & \gate{H} & \ctrl{1} & \meter & \push{\rule{0em}{4em}} \\
190:     & & \qw & \multigate{4}{U_n} & \qw & \qw \\
191:     & & \qw & \ghost{U_n} & \qw & \qw \\
192:     \lstick{\mbox{$I_n/2^n$}} & & \qw & \ghost{U_n} & \qw & \qw \\
193:     & & \qw & \ghost{U_n} & \qw & \qw \\
194:     & & \qw & \ghost{U_n} & \qw & \qw \gategroup{2}{2}{6}{2}{.6em}{\{}
195: }
196: \end{equation}
197: After these operations, the state of the $n+1$ qubits becomes
198: \begin{equation}
199: \rho_{n+1}=
200: {1\over2N}
201: \Bigl(|0\rangle\langle0|\otimes I_n+|1\rangle\langle1|\otimes I_n
202: +|0\rangle\langle1|\otimes U_n^\dagger
203: +|1\rangle\langle0|\otimes U_n\Bigr)
204: = \frac{1}{2N}
205:     \begin{pmatrix}
206:         I_n & \, U_n^\dag \\
207:         \, U_n & I_n
208:     \end{pmatrix} \;,
209: \label{E:rhoout}
210: \end{equation}
211: where $N=2^n$.  The information about the normalized trace of $U_n$ is
212: encoded in the expectation values of the Pauli operators $X$ and $Y$ of
213: the special qubit, i.e., $\langle X\rangle={\rm Re}[\tr(U_n)]/2^n$ and
214: $\langle Y\rangle=-{\rm Im}[\tr(U_n)]/2^n$.
215: 
216: To read out the desired information, say, about the real part of the
217: normalized trace, one runs the circuit repeatedly, each time measuring
218: $X$ on the special qubit at the output.  The measurement results are
219: drawn from a distribution whose mean is the real part of the normalized
220: trace and whose variance is bounded above by 1.  After $L$ runs, one
221: can estimate the real part of the normalized trace with an accuracy
222: $\epsilon\sim1/\sqrt L$.  Thus, to achieve accuracy $\epsilon$ requires
223: that the circuit be run $L\sim1/\epsilon^2$ times. More precisely, what
224: we mean by estimating with fixed accuracy is the following: let $P_e$
225: be the probability that the estimate is farther from the true value
226: than $\epsilon\,$; then the required number of runs is
227: $L\sim\ln(1/P_e)/\epsilon^2$.  That the number of runs required to
228: achieve a fixed accuracy does not scale with number of qubits and
229: scales logarithmically with the error probability is what is meant by
230: saying that the DQC1 circuit provides an efficient method for
231: estimating the normalized trace.
232: 
233: Throughout much of our analysis, we use a generalization of the DQC1
234: circuit, in which the initial pure state of the special qubit is
235: replaced by the mixed state ${1\over2}(I_1+\alpha Z)$,
236: which has polarization $\alpha$,
237: \begin{equation}
238: \label{E:circuitalpha}
239: \Qcircuit @C=.5em @R=-.5em {
240:     & \lstick{{1\over2}(I_1+\alpha Z)}
241:         & \gate{H} & \ctrl{1} & \meter & \push{\rule{0em}{4em}} \\
242:     & & \qw & \multigate{4}{U_n} & \qw & \qw \\
243:     & & \qw & \ghost{U_n} & \qw & \qw \\
244:     \lstick{\mbox{$I_n/2^n$}} & & \qw & \ghost{U_n} & \qw & \qw \\
245:     & & \qw & \ghost{U_n} & \qw & \qw \\
246:     & & \qw & \ghost{U_n} & \qw & \qw \gategroup{2}{2}{6}{2}{.6em}{\{}
247: }
248: \end{equation}
249: giving an overall initial state
250: \begin{equation}
251: \rho_i={1\over2N}(I_1+\alpha Z)\otimes I_n
252: ={1\over2N}\!\left[I_{n+1}+
253: \alpha
254: \begin{pmatrix}
255: I_n&0\\
256: 0&-I_n
257: \end{pmatrix}
258: \right]\;.
259: \end{equation}
260: We generally assume that $\alpha\ge0$, except where we explicitly note
261: otherwise.  After the circuit is run, the system state becomes
262: \begin{equation}
263: \rho_{n+1}(\alpha)=
264: {1\over2N}\!\left[I_{n+1}+
265: \alpha
266: \begin{pmatrix}
267: 0&U_n^\dag\\
268: U_n&0
269: \end{pmatrix}
270: \right]=
271: \frac{1}{2N}
272:     \begin{pmatrix}
273:         I_n &  \alpha U_n^\dag \\
274:         \alpha U_n & I_n
275:     \end{pmatrix} \;.
276: \label{E:rhooutalpha}
277: \end{equation}
278: The effect of subunity polarization is to reduce the expectation
279: values of $\langle X\rangle$ and $\langle Y\rangle$ by a factor of
280: $\alpha$, thereby making it more difficult to estimate the normalized
281: trace.  Specifically, the number of runs required to estimate the
282: normalized trace becomes $L\sim\ln(1/P_e)/\alpha^2\epsilon^2$.  Reduced
283: polarization introduces an additional overhead, but as long as the
284: special qubit has nonzero polarization, the model still provides an
285: efficient estimation of the normalized trace.  What we are dealing with
286: is really the ``power of even the tiniest fraction of a qubit.''
287: 
288: For $n+1$ qubits, \emph{all\/} states contained in a ball of radius
289: $r_{n+1}$ centered at the completely mixed state are
290: separable~\cite{b99,gb03} (distance is measured by the Hilbert-Schmidt
291: norm).  Unitary evolution leaves the distance from the completely mixed
292: state fixed, so at all times during the circuit~(\ref{E:circuitalpha}),
293: the system state is a fixed distance
294: $\sqrt{\tr(\rho_i-I_{n+1}/2N)^2}=\alpha 2^{-(n+1)/2}$ from the
295: completely mixed state.  This suggests that with $\alpha$ small enough,
296: there might be an exponential speedup with demonstrably separable
297: states.  This suggestion doesn't pan out, however, because the radius
298: of the separable ball decreases exponentially faster than
299: $2^{-(n+1)/2}$.  The best known lower bound on $r_n$ is $2\times
300: 6^{-n/2}$~\cite{gb04}; for the system state to be contained in a ball
301: given by this lower bound, we need $\alpha\le 2\times 3^{-(n+1)/2}$.
302: The exponential decrease of $\alpha$ means that an exponentially
303: increasing number of runs is required to estimate the normalized trace
304: with fixed accuracy.  More to the point, the possibility that the
305: actual radius of the separable ball might decrease slowly enough to
306: avoid an exponential number of runs is ruled out by the existence of a
307: family of $n$-qubit entangled states found by D\"ur~\emph{et
308: al.}~\cite{dur99}, which establishes an upper bound on $r_n$ that goes
309: as $2\times 2^{-n}$ for large $n$, implying that $\alpha\le 2\times
310: 2^{-(n+1)/2}$ if the system state is to be in the ball given by the
311: upper bound.  These considerations do not demonstrate the impossibility
312: of an exponential speedup using separable states, but they do rule out
313: the possibility of finding such a speedup within the maximal separable
314: ball about the completely mixed state.
315: 
316: We are thus motivated to look for entanglement in states of the
317: form~(\ref{E:rhooutalpha}), for at least some unitary operators $U_n$.
318: Initial efforts in this direction are not encouraging.  It is clear
319: from the start that the marginal state of the $n$ unpolarized qubits
320: remains completely mixed, so these qubits are not entangled among
321: themselves. Moreover, in the state~(\ref{E:rhooutalpha}), as was shown
322: in Ref.~\cite{pklo04}, the special qubit is unentangled with the $n$
323: unpolarized qubits, no matter what $U_n$ is used.  To see this, one
324: plugs the eigendecomposition of the unitary, $U_n=\sum_j
325: e^{i\phi_j}|e_j\rangle\langle e_j|$, into the expression for
326: $\rho_{n+1}(\alpha)$.  This gives a separable decomposition
327: \begin{equation}
328: \rho_{n+1}(\alpha)=
329: {1\over2N}\sum_j
330: (|a_j\rangle\langle a_j|+|b_j\rangle\langle b_j|)
331: \otimes|e_j\rangle\langle e_j|\;,
332: \end{equation}
333: where $|a_j\rangle=\cos\theta|0\rangle+e^{i\phi_j}\sin\theta|1\rangle$
334: and $|b_j\rangle=\sin\theta|0\rangle+e^{i\phi_j}\cos\theta|1\rangle$,
335: with $\sin2\theta=\alpha$.
336: 
337: No entanglement of the special qubit with the rest and no entanglement
338: among the rest---where then are we to find any entanglement?  We look
339: for entanglement relative to other divisions of the qubits into two
340: parts.  In such bipartite divisions the special qubit is grouped with a
341: subset of the unpolarized qubits.  To detect entanglement between the
342: two parts, we use the Peres-Horodecki partial transpose
343: criterion~\cite{p96,hhh96}, and we quantify whatever entanglement we
344: find using a closely related entanglement monotone which we call the
345: \emph{multiplicative negativity\/}~\cite{Vidal02}.  The Peres-Horodecki
346: criterion and the multiplicative negativity do not reveal all
347: entanglement---they can miss what is called bound entanglement---but we
348: are nonetheless able to demonstrate the existence of entanglement in
349: states of the form~(\ref{E:rhoout}) and~(\ref{E:rhooutalpha}). For
350: convenience, we generally refer to the multiplicative negativity simply
351: as the negativity.  The reader should note, as we discuss in
352: Sec.~\ref{S:negativity}, that the term ``negativity'' was originally
353: applied to an entanglement measure that is closely related to, but
354: different from the multiplicative negativity.
355: 
356: The amount of entanglement depends, of course, on the unitary
357: operator~$U_n$ and on the bipartite division.  We present three results
358: in this regard.  First, in Sec.~\ref{S:examples}, we construct a family
359: of unitaries $U_n$ such that for $\alpha>1/2$, $\rho_{n+1}(\alpha)$ is
360: entangled for all bipartite divisions that put the first and last
361: unpolarized qubits in different parts, and we show that for all such
362: divisions, the negativity is $(2\alpha+3)/4$ for $\alpha\ge1/2$ ($5/4$
363: for $\alpha=1$), independent of~$n$.  Second, in Sec.~\ref{S:random},
364: we present numerical evidence that the state $\rho_{n+1}$ of
365: Eq.~(\ref{E:rhoout}) is entangled for typical unitaries, i.e., those
366: created by random quantum circuits.  For $n+1=5,\ldots,10$, we find
367: average negativities between 1.155 and just above 1.16 for the
368: splitting that puts $\left\lfloor n/2\right\rfloor$ of the unpolarized
369: qubits with qubit~0.  Third, in Sec.~\ref{S:bounds}, we show that for
370: all unitaries and all bipartite divisions of the $n+1$ qubits, the
371: negativity of $\rho_{n+1}(\alpha)$ is bounded above by the constant
372: $\sqrt{1+\alpha^2}$ ($\sqrt2\simeq1.414$ for $\alpha=1$), independent
373: of $n$.  Thus, when $n$ is large, the negativity achievable by the DQC1
374: circuit~(\ref{E:circuit}) becomes a vanishingly small fraction of the
375: maximum negativity, $\sim2^{n/2}$, for roughly equal bipartite
376: divisions.
377: 
378: The layout of the paper is as follows.  In Sec.~\ref{S:classical} we
379: examine the classical problem of estimating the normalized trace of a
380: unitary.  In Sec.~\ref{S:negativity} we review pertinent properties of
381: the negativity before applying it to obtain our three key results in
382: Secs.~\ref{S:examples}-\ref{S:bounds}.  We conclude in
383: Sec.~\ref{S:conclusion} and prove a brief Lemma in an Appendix.
384: Throughout we use $\breve A$ to stand for the partial transpose of an
385: operator $A$ relative to a particular bipartite tensor-product
386: structure, and we rely on context to make clear which bipartite
387: division we are using at any particular point in the paper.
388: 
389: \section{Classical Evaluation of the Trace}\label{S:classical}
390: 
391: In this section we outline briefly a classical method for evaluating the
392: trace of a unitary operator that can be implemented efficiently in terms
393: of quantum gates, and we indicate why this appears to be a problem that
394: is exponentially hard in the number of qubits.
395: 
396: The trace of a unitary matrix $U_n\equiv U$ is the sum over the
397: diagonal matrix elements of~$U$:
398: \begin{equation}
399: \label{E:sum1}
400: \tr(U)=\sum_{\ba} \bra{\ba}U\ket{\ba}\;.
401: \end{equation}
402: Here $\ba$ is a bit string that specifies a computational-basis state
403: of the $n$ qubits.  By factoring $U$ into a product of elementary gates
404: from a universal set and inserting a resolution of the identity between
405: all the gates, we can write $\tr(U)$ as a sum over the amplitudes of
406: Feynman paths.  A difficulty with this approach is that the sum must be
407: restricted to paths that begin and end in the same state.  We can
408: circumvent this difficulty by preceding and succeeding $U$ with a
409: Hadamard gate on all the qubits.  This does not change the trace, but
410: does allow us to write it as
411: \begin{equation}
412: \label{E:sum2}
413: \tr(U)=\sum_{\ba,\bb,\bc}
414: \bra{\ba}H^{\otimes n}\ket{\bb}\bra{\bb}U\ket{\bc}\bra{\bc}H^{\otimes n}\ket{\ba}
415: ={1\over2^n}\sum_{\ba,\bb,\bc}
416: (-1)^{\ba\cdot(\bb+\bc)}\bra{\bb}U\ket{\bc}\;.
417: \end{equation}
418: Now if we insert a resolution of the identity between the elementary
419: gates, we get $\tr(U)$ written as an unrestricted sum over Feynman-path
420: amplitudes, with an extra phase that depends on the initial and final
421: states.
422: 
423: Following Dawson~\emph{et al.}~\cite{dhhmno05}, we consider two
424: universal gate sets: (i)~the Hadamard gate $H$, the $\pi/4$ gate $T$,
425: and the controlled-NOT gate and (ii)~$H$ and the Toffoli gate.  With
426: either of these gate sets, most of the Feynman paths have zero
427: amplitude.  Dawson~\emph{et al.}~\cite{dhhmno05} introduced a
428: convenient method, which we describe briefly now, for including only
429: those paths with nonzero amplitude.  One associates with each wire in
430: the quantum circuit a classical bit value corresponding to a
431: computational basis state.  The effect of an elementary gate is to
432: change, deterministically or stochastically, the bit values at its
433: input and to introduce a multiplicative amplitude.  The two-qubit
434: controlled-NOT gate changes the input control bit $x$ and target bit
435: $y$ deterministically to output values $x$ and $y\oplus x$, while
436: introducing only unit amplitudes.  Similarly, the three-qubit Toffoli
437: gates changes the input control bits $x$ and $y$ and target bit $z$
438: deterministically to $x$, $y$, and $z\oplus xy$, while introducing only
439: unit amplitudes.  The $T$ gate leaves the input bit value $x$ unchanged
440: and introduces a phase $e^{ix\pi/4}$.  The Hadamard gate changes the
441: input bit value $x$ stochastically to an output value $y$ and
442: introduces an amplitude $(-1)^{xy}/\sqrt2$.
443: 
444: The classical bit values trace out the allowed Feynman paths, and the
445: product of the amplitudes introduced at the gates gives the overall
446: amplitude of the path.  In our application of evaluating the
447: trace~(\ref{E:sum2}), a path is specified by $n$ input bit values
448: (which are identical to the output bit values), $n$ random bit values
449: introduced by the initial Hadamard gates, and $h$ random bit values
450: introduced at the $h$ Hadamard gates required for the implementation of
451: $U$.  This gives a total of $2n+h$ bits to specify a path and thus
452: $2^{2n+h}$ allowed paths.  We let $\bx$ denote collectively the $2n+h$
453: path bits.
454: 
455: If we apply the gate rules to a Hadamard-Toffoli circuit, the only gate
456: amplitudes we have to worry about are the $\pm1/\sqrt2$ amplitudes
457: introduced at the Hadamard gates.  There being no complex amplitudes,
458: the trace cannot be complex.  Indeed, for this reason, achieving
459: universality with the $H$-Toffoli gate set requires the use of a simple
460: encoding, and we assume for the purposes of our discussion that this
461: encoding has already been taken into account.  With all this in mind,
462: we can write the trace~(\ref{E:sum2}) as a sum over the allowed paths,
463: \begin{equation}
464: \tr(U)=\frac{1}{2^{n+h/2}}\sum_{\bx}(-1)^{\psi(\bx)}\;.
465: \end{equation}
466: Here $\psi(\bx)$ is a polynomial over $\mz_2$, specifically, the mod-2
467: sum of the products of input and output bit values at each of the
468: Hadamard gates.  The downside is that a string of Toffoli gates
469: followed by a Hadamard can lead to a polynomial that is high order in
470: the bit values.  As pointed out by Dawson~\emph{et
471: al.}~\cite{dhhmno05}, we can deal with this problem partially by
472: putting a pair of Hadamards on the target qubit after each Toffoli
473: gate, thus replacing the quadratic term in the output target bit with
474: two new random variables and preventing the quadratic term from
475: iterating to higher order terms in subsequent Toffoli gates.  In doing
476: so, we are left with a cubic term in $\psi(\bx)$ from the amplitude of
477: the first Hadamard.  The upshot is that we can always make $\psi(\bx)$
478: a cubic polynomial.
479: 
480: Notice now that we can rewrite the trace as
481: \begin{equation}
482: \tr(U)
483: =\frac{1}{2^{n+h/2}}\left[
484: \begin{pmatrix}
485: \mbox{number of $\bx$ such}\\ \mbox{that $\psi(\bx)=0$}
486: \end{pmatrix}-
487: \begin{pmatrix}
488: \mbox{number of $\bx$ such}\\ \mbox{that $\psi(\bx)=1$}
489: \end{pmatrix}\right]\;,
490: \end{equation}
491: thus reducing the problem of evaluating the trace exactly to counting
492: the number of zeroes of the cubic polynomial $\psi(\bx)$.  This is a
493: standard problem from computational algebraic geometry, and it is known
494: that counting the number of zeroes of a general cubic polynomial over
495: any finite field is \#{\bf P} complete \cite{ek90}.  It is possible
496: that the polynomials that arise from quantum circuits have some special
497: structure that can be exploited to give an efficient algorithm for
498: counting the number of zeroes, but in the absence of such structure,
499: there is no efficient classical algorithm for computing the trace
500: exactly unless the classical complexity hierarchy collapses and
501: \emph{all\/} problems in \#{\bf P} are efficiently solvable on a
502: classical computer.
503: 
504: Of course, it is not our goal to compute the trace exactly, since the
505: quantum circuit only provides an efficient method for estimating the
506: normalized trace to fixed accuracy.  This suggests that we should
507: estimate the normalized trace by sampling the amplitudes of the allowed
508: Feynman paths.  The normalized trace,
509: \begin{equation}
510: {\tr(U)\over2^n}=\frac{1}{2^{2n+h}}\sum_{\bx}2^{h/2}(-1)^{\psi(\bx)}\;,
511: \end{equation}
512: which lies between $-1$ and $+1$, can be regarded as the average of
513: $2^{2n+h}$ quantities whose magnitude, $2^{h/2}$, is exponentially
514: large in the number of Hadamard gates.  To estimate the average with
515: fixed accuracy requires a number of samples that goes as $2^h$,
516: implying that this is not an efficient method for estimating the
517: normalized trace.  The reason the method is not efficient is pure
518: quantum mechanics, i.e., that the trace is a sum of amplitudes, not
519: probabilities.
520: 
521: If we apply the gate rules to a Hadamard-$T$-controlled-NOT circuit,
522: the bit value on each wire in the circuit is a mod-2 sum of appropriate
523: bit values in $\bx$, but now we have to worry about the amplitudes
524: introduced by the Hadamard and $T$ gates.  The trace~(\ref{E:sum2})
525: can be written as
526:  \begin{equation}
527:   \tr(U)={1\over2^{n+h/2}}
528:   \sum_{\bx}e^{i(\pi/4)\chi(\bx)}(-1)^{\phi(\bx)}\;.
529:   \label{E:sum3}
530:  \end{equation}
531: Here $\phi(\bx)$ is a polynomial over $\mz_2$, obtained as the mod-2
532: sum of the products of input and output bit values at each of the
533: Hadamard gates.  Since the output value is a fresh binary variable and
534: the input value is a mod-2 sum of bit values in $\bx$, $\phi(\bx)$ is a
535: purely quadratic polynomial over $\mz_2$.  The function $\chi(\bx)$ is
536: a mod-8 sum of the input bit values to all of the $T$ gates.  Since
537: these input bit values are mod-2 sums of bit values in $\bx$,
538: $\chi(\bx)$ is linear in bit values, but with an unfortunate mixture of
539: mod-2 and mod-8 addition.  We can get rid of this mixture by preceding
540: each $T$ gate with a pair of Hadamards, thus making the input to the
541: every $T$ gate a fresh binary variable.  With this choice, $\chi(\bx)$
542: becomes a mod-8 sum of appropriate bit values from $\bx$.
543: 
544: We can rewrite the sum~(\ref{E:sum3}) in the following way:
545: \begin{equation}
546:  \tr(U)={1\over2^{n+h/2}}
547:  \sum_{j=0}^7 e^{i(\pi/4)j} \left[
548: \begin{pmatrix}
549: \mbox{number of $\bx$ such that}\\
550: \mbox{$\chi(\bx)=j$ and $\phi(\bx)=0$}
551: \end{pmatrix}
552: -
553: \begin{pmatrix}
554: \mbox{number of $\bx$ such that}\\
555: \mbox{$\chi(\bx)=j$ and $\phi(\bx)=1$}
556: \end{pmatrix}
557: \right].
558: \end{equation}
559: Thus the problem now reduces to finding simultaneous (binary) solutions
560: to the purely quadratic $\mz_2$ polynomial $\phi(\bx)$ and the purely
561: linear $\mz_8$ polynomial $\chi(\bx)$.  One has to be careful here to
562: note that we are only interested in binary solutions, so we are not
563: solving $\chi(\bx)=j$ over all values in $\mz_8$.  The number of
564: solutions of a purely quadratic polynomial over $\mz_2$ can be obtained
565: trivially~\cite{ek90}, but the constraint over $\mz_8$ means that one
566: must count the number of solutions over a mixture of a field and a
567: ring.  The complexity class for this problem is not known, but given
568: the equivalence to counting the number of solutions of a cubic
569: polynomial over $\mz_2$, it seems unlikely that there is an efficient
570: classical algorithm.  Moreover, an attempt to estimate the normalized
571: trace by sampling allowed paths obviously suffers from the problem
572: already identified above.
573: 
574: \section{Properties of negativity}
575: \label{S:negativity}
576: 
577: In this section we briefly review properties of negativity as an
578: entanglement measure, focusing on those properties that we need in the
579: subsequent analysis (for a thorough discussion of negativity, see
580: Ref.~\cite{Vidal02}).
581: 
582: Let $A$ be an operator in the joint Hilbert space of two systems,
583: system~1 of dimension $d_1$ and system~2 of dimension $d_2$.  The
584: partial transpose of $A$ with respect to an orthonormal basis of
585: system~2 is defined by taking the transpose of the matrix elements of
586: $A$ with respect to the system-2 indices.  A partial transpose can also
587: be defined with respect to any basis of system~1.  Partial
588: transposition preserves the trace, and it commutes with taking the
589: adjoint.
590: 
591: The operator that results from partial transposition depends on which
592: basis is used to define the transpose, but these different partial
593: transposes are related by unitary transformations on the transposed
594: system and thus have the same eigenvalues and singular values.
595: Moreover, partial transposition on one of the systems is related to
596: partial transposition on the other by an overall transposition, which
597: also preserves eigenvalues and singular values.  Despite the
598: nonuniqueness of the partial transpose, we can talk meaningfully about
599: its invariant properties, such as its eigenvalues and singular values.
600: Similar considerations show that the eigenvalues and singular values
601: are invariant under local unitary transformations.
602: 
603: The singular values of an operator $O$ are the eigenvalues of
604: $\sqrt{O^\dag O}\equiv|O|$ (or, equivalently, of $\sqrt{O
605: O^\dag}$).  Any operator has a polar decomposition $O=T|O|$, where
606: $T$ is a unitary operator.  Writing $|O|=W^\dag SW$, where $W$ is
607: the unitary that diagonalizes $|O|$ and $S$ is the diagonal matrix of
608: singular values, we see that any operator can be written as $O=VSW$,
609: where $V=TW^\dag$ and $W$ are unitary operators.
610: 
611: We denote a partial transpose of $A$ generically by $\breve A$.  We
612: write the eigenvalues of $\breve A$ as $\lambda_j(\breve A)$ and denote
613: the singular values by $s_j(\breve A)$.  If $A$ is Hermitian, so is
614: $\breve A$, and the singular values of $\breve A$, i.e., the
615: eigenvalues of $|\breve A|$, are the magnitudes of the eigenvalues,
616: i.e, $s_j(\breve A)=|\lambda_j(\breve A)|$.
617: 
618: If a joint density operator~$\rho$ of systems 1 and 2 is separable, its
619: partial transpose $\breve\rho$ is a positive operator.  This gives the
620: Peres-Horodecki entanglement criterion~\cite{p96,hhh96}: if
621: $\breve\rho$ has a negative eigenvalue, then $\rho$ is entangled (the
622: converse is not generally true).  The magnitude of the sum of the
623: negative eigenvalues of the partial transpose, denoted by
624: \begin{equation}
625: \sN(\rho)\equiv-\sum_{\lambda_j(\breve\rho)<0}\lambda_j(\breve\rho)\;,
626: \end{equation}
627: is a measure of the amount of entanglement.  Partial transposition
628: preserves the trace, so $\tr(\breve\rho)=1$, from which we get
629: \begin{equation}
630: 1+2\sN(\rho)=\sum_j|\lambda_j(\breve\rho)|=
631: \tr|\breve\rho|\equiv\sM(\rho)\;,
632: \end{equation}
633: where $\sM(\rho)$ is a closely related entanglement measure.  The
634: quantity $\sN(\rho)$ was originally called the
635: \emph{negativity\/}~\cite{Vidal02}; we can distinguish the two measures
636: by referring to $\sM(\rho)$ as the \emph{multiplicative negativity}, a
637: name that emphasizes one of its key properties and advantages over
638: $\sN(\rho)$.  In this paper, however, we use the multiplicative
639: negativity exclusively and so refer to it simply as ``the negativity''.
640: 
641: The negativity $\sM(\rho)$ equals one for separable states, and it is
642: an entanglement monotone~\cite{Vidal02}, meaning that (i)~it is a
643: convex function of density operators and (ii)~it does not increase
644: under local operations and classical communication.  The negativity has
645: the property of being multiplicative in the sense that the $\sM$ value
646: for a state that is a product of states for many pairs of systems is
647: the product of the $\sM$ values for each of the pairs.  By the same
648: token, $\log\sM(\rho)$, called the \emph{log-negativity\/}, is
649: additive, but the logarithm destroys convexity so the log-negativity is
650: not an entanglement monotone~\cite{Vidal02}.  For another point of view
651: on the monotonicity of the log-negativity, see Ref.~\cite{Plenio05}.
652: 
653: The minimum value of the negativity is one, but we need to know the
654: maximum value to calibrate our results.  Convexity guarantees that the
655: maximum value is attained on pure states.  We can find the
656: maximum~\cite{SLee03} by considering the Schmidt decomposition of a
657: joint pure state of systems~1 and 2,
658: \begin{equation}
659: |\psi\rangle=\sum_{j=1}^d\sqrt{\mu_j}|j,j\rangle\;,
660: \end{equation}
661: where $d=\min(d_1,d_2)$.  Taking the partial transpose of $\rho$
662: relative to the Schmidt basis of system~2 gives
663: \begin{equation}
664: \breve\rho
665: =\sum_{j,k=1}^d\sqrt{\mu_j\mu_k}|j,k\rangle\langle k,j|\;,
666: \end{equation}
667: with eigenvectors and eigenvalues
668: \begin{eqnarray}
669: |j,j\rangle\;,&&\mbox{eigenvalue $\mu_j$,}\nonumber\\
670: {1\over\sqrt2}(|j,k\rangle\pm|k,j\rangle)\;,
671: &&\mbox{eigenvalue $\pm\sqrt{\mu_j\mu_k}$,}\quad j<k.
672: \end{eqnarray}
673: This gives a negativity
674: \begin{equation}
675: \sM(\psi)=
676: 1+2\sum_{j<k}\sqrt{\mu_j\mu_k}
677: =\sum_{j,k=1}^d\sqrt{\mu_j\mu_k}
678: =\biggl(\sum_{j=1}^d\sqrt{\mu_j}\biggr)^2\;.
679: \end{equation}
680: The concavity of the square root implies $\sum_j\sqrt{\mu_j}\le\sqrt
681: d$, with equality if and only if $\mu_j=1/d$ for all $j$, i.e.,
682: $|\psi\rangle$ is maximally entangled. We end up with
683: \begin{equation}
684: 1\le\sM(\rho)\le d\;.
685: \end{equation}
686: 
687: The negativity is the sum of the singular values of
688: $\breve\rho$.  For states of the form we are interested in, given by
689: Eq.~(\ref{E:rhooutalpha}), the negativity is determined by the
690: singular values of the partial transpose of the unitary operator~$U_n$.
691: To see this, consider any bipartite division of the qubits. Performing
692: the partial transpose on the part that does not include the special
693: qubit, we have
694: \begin{equation}
695: \breve\rho_{n+1}(\alpha)=
696: \frac{1}{2N}
697:     \begin{pmatrix}
698:         I_n &  \alpha\breve U_n^\dag \\
699:         \alpha\breve U_n & I_n
700:     \end{pmatrix} \;,
701: \label{E:breverhooutalpha}
702: \end{equation}
703: where $\breve U_n$ is the partial transpose of $U_n$ relative to the
704: chosen bipartite division.  Notice that if we make our division between
705: the special qubit and all the rest, then $\breve U_n=U_n^T$ is a
706: unitary operator, and $\breve\rho_{n+1}(\alpha)$ is the quantum state
707: corresponding to using $U_n^T$ in the circuit~(\ref{E:circuitalpha});
708: this shows that for this division, the negativity is 1, consistent with
709: our earlier conclusion that the special qubit is not entangled with the
710: other qubits.  For a general division, we know there are unitaries $V$
711: and $W$ such that $\breve U_n=VSW$, where $S$ is the diagonal matrix of
712: singular values $s_j(\breve U_n)$.  This allows us to write
713: \begin{equation}
714: \breve\rho_{n+1}(\alpha)=
715:     \begin{pmatrix}
716:         W^\dagger & 0 \\
717:         0 & V
718:     \end{pmatrix}
719:     \frac{1}{2N}
720:     \begin{pmatrix}
721:         I_n &  \alpha S \\
722:         \alpha S & I_n
723:     \end{pmatrix}
724:     \begin{pmatrix}
725:         W & 0 \\
726:         0 & V^\dagger
727:     \end{pmatrix}\;,
728: \end{equation}
729: showing that $\breve\rho_{n+1}(\alpha)$ is a unitary transformation
730: away from the matrix in the middle and thus has the same eigenvalues.
731: The block structure of the middle matrix makes it easy to find these
732: eigenvalues, which are given by $[1\pm\alpha s_j(\breve U_n)]/2N$.
733: This allows us to put the negativity in the form
734: \begin{equation}
735: \sM\bigl(\rho_{n+1}(\alpha)\bigr)={1\over2N}
736: \sum_{j=1}^N|1+\alpha s_j(\breve U_n)|+|1-\alpha s_j(\breve U_n)| =
737: {1 \over N} \sum_{j=1}^N \max\bigl(\abs{\alpha} s_j(\breve U_n), 1\bigr)\;,
738: \label{E:Msingular}
739: \end{equation}
740: which is valid for both positive and negative values of $\alpha$.  An
741: immediate consequence of Eq.~(\ref{E:Msingular}) is that
742: $\sM\bigl(\rho_{n+1}(\alpha)\bigr)=\sM\bigl(\rho_{n+1}(-\alpha)\bigr)$,
743: as one would expect.  Since $\rho_{n+1}(\alpha)$ is a mixture of
744: $\rho_{n+1}(+1)=\rho_{n+1}$ and $\rho_{n+1}(-1)$, convexity tells us
745: immediately that
746: $\sM\bigl(\rho_{n+1}(\alpha)\bigr)\le\sM\bigl(\rho_{n+1}\bigr)$, i.e.,
747: that a mixed input for the special qubit cannot increase the negativity
748: over that for a pure input.  More generally, we have that the
749: negativity cannot decrease at any point as $\alpha$ increases from 0
750: to~1.
751: 
752: \section{Entanglement in the DQC1 Circuit} \label{S:examples}
753: 
754: In this section, we construct a family of unitaries $U_n$ that produce
755: global entanglement in the DQC1 circuit~(\ref{E:circuit}).  For
756: $\alpha=1$, the negativity produced by this family is equal to $5/4$,
757: independent of $n$, for all bipartite divisions that put the first and
758: last unpolarized qubits in different parts.  We conjecture that this is
759: the maximum negativity that can be achieved in a circuit of the
760: form~(\ref{E:circuit}).
761: 
762: Before the measurement, the output state of the
763: circuit~(\ref{E:circuitalpha}) is given by Eq.~(\ref{E:rhooutalpha}).
764: To construct the unitaries $U_n$, we first introduce a two-qubit
765: unitary matrix
766: \begin{equation}
767: \label{E:U2}
768:     U_{2} \equiv \begin{pmatrix}
769:         A_1 & C_1 \\
770:         D_1 & B_1 \\
771:     \end{pmatrix} \;,
772: \end{equation}
773: where $A_1$, $B_1$, $C_1$, and $D_1$ are single-qubit ($2\times2$)
774: matrices that must satisfy $A_1^\dagger A_1+D_1^\dagger D_1=
775: B_1^\dagger B_1+C_1^\dagger C_1=I_1$ and $A_1^\dagger C_1+D_1^\dagger
776: B_1=0$ to ensure that $U_2$ is unitary.  The $n$-qubit unitary $U_n$
777: is then defined by
778: \begin{eqnarray}
779:     U_n &\equiv&
780:     \begin{pmatrix}
781:         I_{n-2} \otimes A_1 & X_{n-2} \otimes C_1 \\
782:         X_{n-2} \otimes D_1 & I_{n-2} \otimes B_1
783:     \end{pmatrix} \nonumber \\
784:     &=&|0\rangle\langle0|\otimes I_{n-2}\otimes A_1
785:     +|1\rangle\langle1|\otimes I_{n-2}\otimes B_1 \nonumber \\
786:     &&\phantom{|}
787:     +|0\rangle\langle1|\otimes X_{n-2}\otimes C_1
788:     +|1\rangle\langle0|\otimes X_{n-2}\otimes D_1
789:     \;.
790:     \label{E:Un}
791: \end{eqnarray}
792: Here we use $X_1$, $Y_1$, and $Z_1$ to denote single-qubit Pauli
793: operators.  A subscript $k$ on the identity operator or a Pauli operator
794: denotes a tensor product in which that operator acts on each of $k$
795: qubits.  If we adopt the convention that $X_0 = I_0 = 1$, then $U_n$
796: reduces to $U_2$ when $n=2$.  It is easy to design a quantum circuit
797: that realizes $U_n$.  The structure of the circuit is illustrated by
798: the case of $U_4$:
799: \begin{equation}
800: \label{E:Ucircuit}
801:     \Qcircuit @C=.5em @R=0.5em {
802:     & \lstick{\mbox{1st qubit}} & \ctrl{2} & \ctrl{3} & \multigate{1}{U_2} & \ctrl{3} & \ctrl{2} & \qw \\
803:     & \lstick{\mbox{4th qubit}} & \qw      & \qw      & \ghost{U_n}        & \qw    & \qw & \qw \\
804:     & \lstick{\mbox{2nd qubit}} & \targ    & \qw      & \qw                & \qw      & \targ & \qw \\
805:     & \lstick{\mbox{3rd qubit}} & \qw      & \targ    & \qw                & \targ    & \qw & \qw                    }
806: \end{equation}
807: In general, the two-qubit unitary $U_2$, acting on the first and last
808: qubits, is bracketed by controlled-NOT gates from the first qubit,
809: acting as control, to each of the other qubits, except the last, as
810: targets.
811: 
812: Because $I_1$ and $X_1$ are invariant under transposition, it is clear
813: from the form of $U_n$ that in the state~(\ref{E:rhooutalpha}), all
814: qubits, except~0, 1, and~$n$, are invariant under transposition.  We
815: can use this fact to find the negativity for all bipartite divisions.
816: First consider any bipartite division that puts qubits 1 and $n$ in the
817: same part.  There are two possibilities.  If the special qubit is in
818: the same part as qubits~1 and $n$, then partial transposition on the
819: other part leaves $\rho_{n+1}(\alpha)$ unchanged, so the negativity
820: is~1.  If the special qubit is not in the same part as~1 and $n$, then
821: partial transposition on the part that includes~1 and $n$ is the same
822: as partial transposition of all the unpolarized qubits, a case we
823: already know to have negativity equal to 1.  We conclude that any
824: bipartite division that puts~1 and $n$ in the same part has negativity
825: equal to 1.
826: 
827: Turn now to bipartite divisions that put qubits~1 and $n$ in different
828: parts.  There are two cases to consider: (i)~the special qubit is in
829: the same part as qubit~1, and (ii)~the special qubit is in the same part
830: as qubit~$n$.  In case~(i), partial transposition of the part that
831: contains qubit~$n$ gives
832: \begin{equation}
833: \breve\rho_{n+1}(\alpha)={1\over2N}
834:     \begin{pmatrix}
835:         I_n                 & \alpha \breve U_n^\dagger \\
836:         \alpha \breve U_n   & I_n
837:     \end{pmatrix}
838: \qquad\mbox{with}\qquad
839: \breve U_n=
840: \begin{pmatrix}
841:         I_{n-2} \otimes A_1^T & X_{n-2} \otimes C_1^T \\
842:         X_{n-2} \otimes D_1^T & I_{n-2} \otimes B_1^T
843:     \end{pmatrix}\;.
844: \end{equation}
845: In case~(ii), partial transposition of the part that contains qubit~1
846: gives
847: \begin{equation}
848: \breve\rho_{n+1}(\alpha)={1\over2N}
849:     \begin{pmatrix}
850:         I_n                 & \alpha \breve U_n^\dagger \\
851:         \alpha \breve U_n   & I_n
852:     \end{pmatrix}
853: \qquad\mbox{with}\qquad
854: \breve U_n=
855: \begin{pmatrix}
856:         I_{n-2} \otimes A_1 & X_{n-2} \otimes D_1 \\
857:         X_{n-2} \otimes C_1 & I_{n-2} \otimes B_1
858:     \end{pmatrix}\;.
859: \end{equation}
860: 
861: The basic structure of $\breve\rho_{n+1}(\alpha)$ is the same in
862: both cases.  Without changing the spectrum, we can reorder the rows and
863: columns to block diagonalize $\breve\rho_{n+1}(\alpha)$ so that
864: there are $N/4$ blocks, each of which has the form
865: \begin{equation}
866: {1\over2N}
867: \begin{pmatrix}
868:     I_2                 & \alpha \breve U_2^\dagger \\
869:     \alpha \breve U_2   & I_2
870: \end{pmatrix}
871: ={4\over N}\breve\rho_3(\alpha)\;,
872: \label{E:breverho2}
873: \end{equation}
874: where $\breve\rho_3(\alpha)$ is the appropriate partial transpose of
875: the three-qubit output state.  Thus the spectrum of
876: $\breve\rho_{n+1}(\alpha)$ is the same as the spectrum of
877: $\breve\rho_3(\alpha)$, except that each eigenvalue is reduced by a
878: factor of $4/N$.  In calculating the negativity, since each eigenvalue
879: is ($N/4$)-fold degenerate, the reduction factor of $4/N$ is cancelled
880: by a degeneracy factor of $N/4$, leaving us with the fundamental result
881: of our construction,
882: \begin{equation}
883: \sM\bigl(\rho_{n+1}(\alpha)\bigr)=\sM\bigl(\rho_3(\alpha)\bigr)\;.
884: \end{equation}
885: This applies to both cases of bipartite splittings that we are
886: considering, showing that all divisions have the same negativity as the
887: corresponding $n=2$ construction.
888: 
889: We now specialize to a particular choice of $U_2$ given by
890: \begin{equation}
891: A_1=
892:     \begin{pmatrix}
893:         0 & 0 \\ 0 & 1
894:     \end{pmatrix}\;,
895: \quad
896: B_1=
897:     \begin{pmatrix}
898:         1 & 0 \\ 0 & 0
899:     \end{pmatrix}\;,
900: \quad
901: C_1=
902:     \begin{pmatrix}
903:         0 & 1 \\ 0 & 0
904:     \end{pmatrix}\;,
905: \quad\mbox{and}\quad
906: D_1=
907:     \begin{pmatrix}
908:         0 & 0 \\ 1 & 0
909:     \end{pmatrix}\;.
910: \label{E:ABCD}
911: \end{equation}
912: For this choice, the two cases of bipartite division lead to the same
913: partial transpose.  The spectrum of
914: \begin{equation}
915: \breve\rho_3(\alpha)={1\over8}
916: \begin{pmatrix}
917: I_1 & 0 & \alpha A_1 & \alpha D_1 \\
918: 0 & I_1 & \alpha C_1 & \alpha B_1  \\
919: \alpha A_1 & \alpha D_1 & I_1 & 0\\
920: \alpha C_1 & \alpha B_1 & 0 & I_1
921: \end{pmatrix}
922: \end{equation}
923: is
924: \begin{equation}
925: \spec(\breve\rho_3\bigl(\alpha)\bigr)=
926: {1\over 8}(1+2\alpha,1,1,1,1,1,1-2\alpha)\;,
927: \end{equation}
928: giving a negativity equal to 1 for $\alpha\le1/2$ and a
929: negativity
930: \begin{equation}
931: \sM\bigl(\rho_{n+1}(\alpha)\bigr)=\sM\bigl(\rho_3(\alpha)\bigr)=
932: {1\over4}(2\alpha+3)
933: \quad\mbox{for}\quad
934: \alpha\ge1/2\;.
935: \end{equation}
936: This result shows definitively that the circuit~ can produce
937: entanglement, at least for $\alpha>1/2$.  We stress that the
938: negativity achieved by this family of unitaries is independent of
939: $n\ge2$.
940: 
941: For $\alpha=1$, the negativity achieved by this family reduces to
942: $5/4$.  For large $n$, this amount of negativity is a vanishingly small
943: fraction of the maximum possible negativity, $\sim2^{n/2}$, for roughly
944: equal divisions of the qubits.  This raises the question whether it is
945: possible for other unitaries to achieve larger negativities.  A first
946: idea might be to find two-qubit unitaries $U_2$ that yield a higher
947: negativity $\sM(\rho_3)=\sM(\rho_{n+1})$ when plugged into the
948: construction of this section, but the bounds we find in
949: Sec.~\ref{S:bounds} dispose of this notion, since they show that $5/4$
950: is the maximum negativity that can be achieved for $n=2$.  Another
951: approach would be to generalize the construction of this section in a
952: way that is obvious from the circuit~(\ref{E:Ucircuit}), i.e., by
953: starting with a $k$-qubit unitary in place of the two-qubit unitary of
954: Eq.~(\ref{E:Ucircuit}).  Numerical investigation of the case $k=3$ has
955: not turned up negativities larger than $5/4$.  We conjecture that $5/4$
956: is the maximum negativity that can be achieved by states of the
957: form~(\ref{E:rhoout}).  Though we have not been able to prove this
958: conjecture, we show in the next section that typical unitaries for
959: $n+1\le10$ achieve negativities less than $5/4$ and in the following
960: section that the negativity is rigorously bounded by $\sqrt2$.
961: 
962: We stress that we are not suggesting that the construction of this
963: section, with $U_2$ given by Eq.~(\ref{E:ABCD}), achieves the maximum
964: negativity for all values of $\alpha$, for that would mean that we
965: believed that the negativity cannot exceed 1 for $\alpha\le1/2$, which
966: we do not.  Although we have not found entanglement for $\alpha\le
967: 1/2$, we suspect there are states with negativity greater than 1 as
968: long as $\alpha$ is large enough that $\rho_{n+1}(\alpha)$ lies outside
969: the separable ball around the maximally mixed state, i.e.,
970: $\alpha\ge2^{(n+1)/2}r_{n+1}$.  The bound of Sec.~\ref{S:bounds} only
971: says that $\sM\bigl(\rho_{n+1}(\alpha)\bigr)\le\sqrt{1+\alpha^2}$, thus
972: allowing negativities greater than 1 for all values of $\alpha$ except
973: $\alpha=0$.  Moreover, since the negativity does not detect bound
974: entanglement, there could be entangled states that have a negativity
975: equal to~1.
976: 
977: \section{The Average Negativity of a Random Unitary} \label{S:random}
978: 
979: Having constructed a family of unitaries that yields a DQC1 state with
980: negativity $5/4$, a natural question to ask is, ``What is the
981: negativity of a typical state produced by the
982: circuit~(\ref{E:circuit})?''  To address this question, we choose the
983: unitary operator in the circuit~(\ref{E:circuit}) at random and
984: calculate the negativity.  Of course, one must first define what it
985: means for a unitary to be ``typical'' or ``chosen at random''.  The
986: natural measure for defining this is the Haar measure, which is the
987: unique left-invariant measure for the group ${\sf U}(N)$
988: \cite{Conway90}.  The resulting ensemble of unitaries is known as the
989: Circular Unitary Ensemble, or CUE, and it is parameterized by the
990: Hurwitz decomposition \cite{Hurwitz1897}.  Although this is an exact
991: parameterization, implementing it requires computational resources that
992: grow exponentially in the size of the unitary~\cite{Emerson03}.  To
993: circumvent this, a pseudo-random distribution that requires resources
994: growing polynomially in the size of the unitary was formulated and
995: investigated in Ref.~\cite{Emerson03}. This is the distribution from
996: which we draw our random unitaries, and we summarize the procedure for
997: completeness.
998: 
999: We first define a random ${\sf SU}(2)$ unitary as
1000: \begin{equation}
1001:     R(\theta,\phi,\chi) =
1002:     \begin{pmatrix}
1003:     e^{i \phi} \cos\theta & e^{i \chi} \sin\theta \\
1004:     -e^{-i \chi} \sin\theta & e^{-i \phi} \cos\theta \\
1005:     \end{pmatrix} \;,
1006: \end{equation}
1007: where $\theta$ is chosen uniformly between $0$ and $\pi/2$, and $\phi$
1008: and $\chi$ are chosen uniformly between $0$ and $2\pi$.  A random
1009: unitary applied to each of the $n$ qubits is then
1010: \begin{equation}
1011:     {\sf R} = \bigotimes_{i=1}^{n} R(\theta_i, \phi_i, \chi_i) \;,
1012: \end{equation}
1013: where a separate random number is generated for each variable at each
1014: value of $i$.  Now define a mixing operator ${\sf M}$ in terms of
1015: nearest-neighbor $Z\otimes Z$ couplings as
1016: \begin{equation}
1017:     {\sf M} =
1018:     \exp\left(i \frac{\pi}{4} \sum_{j=1}^{n-1} Z^{(j)} \otimes Z^{(j+1)} \right) \;.
1019: \end{equation}
1020: The pseudo-random unitary is then given by
1021: \begin{equation}
1022:     {\sf R}_j{\sf M}{\sf R}_{j-1}\cdots{\sf M}{\sf R}_2{\sf M}{\sf R}_1 \;,
1023: \end{equation}
1024: where $j$ is a positive integer that depends on $n$, and each ${\sf
1025: R}_k$ is chosen randomly as described above.  For a given $n$, the
1026: larger $j$ is, the more accurately the pseudo-random unitary
1027: distribution resembles the actual CUE.  From the results in
1028: Ref.~\cite{Emerson03}, $j=40$ gives excellent agreement with the CUE
1029: for unitary operators on at least up to 10 qubits, so this is what we
1030: use in our calculations.
1031: 
1032: Due to boundary effects, not all bipartite splittings that put $k$
1033: unpolarized qubits in one part are equivalent. Nevertheless, we
1034: consider only bipartite divisions that split the qubits along
1035: horizontal lines placed at various points in the circuit of
1036: Eq.~(\ref{E:circuit}).  We refer to the division that groups the last
1037: $k$ qubits together as the $(n+1-k,k)$ splitting.  For $\alpha=1$, we
1038: calculate the average negativity and standard deviation of a
1039: pseudo-random state $\rho_{n+1}$ for two different bipartite
1040: splittings, $(n,1)$ and $(\left\lfloor n/2 \right\rfloor +1,
1041: \left\lceil n/2 \right\rceil )$. These results are plotted in
1042: Fig.~\ref{F:random}.  For $n+1=5,\ldots,10$, the average negativity for
1043: the roughly equal splitting lies between 1.135 and just above 1.16. The
1044: standard deviation appears to converge exponentially to zero, as in
1045: Ref.~\cite{Scott03}, a behavior that is typical of asymptotically
1046: equivalent matrices.  In addition, for $9+1$ qubits, we calculate the
1047: average negativity and standard deviation for all nontrivial ($k\ne n$)
1048: bipartite splittings $(n+1-k,k)$, and the results are shown in
1049: Fig.~\ref{F:allsplits}.
1050: 
1051: \begin{figure}[t]
1052: \begin{center}
1053: \includegraphics[scale=.458]{meanneg.eps}
1054: \includegraphics[scale=.458]{devneg.eps}
1055: \caption{{\bf Left}: Average negativity of the state $\rho_{n+1}$ of
1056: Eq.~(\protect\ref{E:rhoout}) ($\alpha=1$) for a randomly chosen unitary
1057: $U_n$ for two different bipartite splittings, $(n,1)$ and
1058: $\left(\left\lfloor n/2 \right\rfloor +1, \left\lceil
1059: n/2 \right\rceil\right)$.  The $(n,1)$ splitting appears to
1060: reach an upper bound quickly, whereas the other splitting is still
1061: rising slowly at 10 qubits.  {\bf Right}: Semi-log plot of the standard
1062: deviation in the negativity of the randomly chosen state $\rho_{n+1}$.
1063: The fit curves show that the standard deviation is decaying
1064: exponentially, so that for large numbers of qubits, almost all
1065: unitaries give the same negativity.} \label{F:random}
1066: \end{center}
1067: \end{figure}
1068: 
1069: \begin{figure}[t]
1070: \begin{center}
1071: \includegraphics[scale=.6]{allsplits10.eps}
1072: \caption{Average negativity of the state $\rho_{10}$ of
1073: Eq.~(\protect\ref{E:rhoout}) ($\alpha=1$) for a randomly chosen unitary
1074: $U_9$ as a function of bipartite splitting number $k$ for bipartite
1075: splittings $(10-k,k)$.  The error bars give the standard deviations.
1076: The function attains a maximum when the bipartite split is made between
1077: half of the qubits on which the unitary acts.}
1078: \label{F:allsplits}
1079: \end{center}
1080: \end{figure}
1081: 
1082: \section{Bounds on the Negativity} \label{S:bounds}
1083: 
1084: In this section, we return to allowing the special qubit in the circuit
1085: (\ref{E:circuitalpha}) to have initial polarization~$\alpha$.  Since the
1086: value of $n$ is either clear from context or fixed, we reduce the
1087: notational clutter by denoting the state $\rho_{n+1}(\alpha)$ of
1088: Eq.~(\ref{E:rhooutalpha}) as~$\rho_\alpha$.
1089: 
1090: Given a particular bipartite division, the partial transpose of
1091: $\rho_\alpha$ with respect to the part that does not include the
1092: special qubit is
1093: \begin{equation}
1094:     \breve \rho_\alpha = \frac{I_n+\alpha \breve C}{2N} \;,
1095: \end{equation}
1096: where
1097: \begin{equation}
1098: \label{E:Cdef}
1099:     \breve C \equiv
1100:     \begin{pmatrix}
1101:         0 & \breve U_n^\dag \\
1102:         \breve U_n & 0
1103:     \end{pmatrix} \;.
1104: \end{equation}
1105: Using the binomial theorem, we can expand
1106: $\tr(\breve\rho^{\,s}_\alpha)$ in terms of $\tr(\breve C^k)$:
1107: \begin{equation}
1108: \label{E:binomialform}
1109:     \tr(\breve\rho^{\,s}_\alpha) =
1110:     \left({1\over2N}\right)^{\! \! s} \sum_{k=0}^s {s \choose k} \alpha^k \tr(\breve C^k) \;.
1111: \end{equation}
1112: When $k$ is odd, $\breve C^k$ is block off-diagonal, so its trace
1113: vanishes.  When $k$ is even, we have
1114: \begin{equation}
1115: \tr(\breve C^k) = 2\,\tr\Bigl((\breve U_n \breve U_n^\dag)^{k/2}\Bigr)\;.
1116: \end{equation}
1117: When $k=2$, this simplifies to $\tr(\breve C^2)=2\tr(\breve U_n\breve
1118: U_n^\dag)=2\tr(U_n U_n^\dag)=2N$.  The crucial step here follows
1119: immediately from the property $\tr(\breve A\breve B)=\tr(AB)$, which we
1120: prove as a Lemma in the Appendix.  Note that in general $\tr(\breve A_1
1121: \breve A_2 \ldots \breve A_l) \not= \tr(A_1 A_2 \ldots A_l)$ if $l>2$,
1122: so we cannot give a similar general calculation of
1123: $\tr(\breve\rho^{\,s}_\alpha)$ for even $s \ge 4$, since it involves terms
1124: of this form.
1125: 
1126: Using Eq.~(\ref{E:binomialform}), we can now obtain three independent
1127: constraint equations on the eigenvalues
1128: $\lambda_j=\lambda_j(\breve\rho_\alpha)$ of the partial transpose
1129: $\breve\rho_\alpha$:
1130: \begin{equation} \label{E:sumsofpowers}
1131:     \sum_{j=1}^{2N}\lambda_j^s=
1132:     \tr(\breve \rho^{\,s}_\alpha)
1133:     = \frac{1}{2^s N^{s-1}}[(1+\alpha)^s + (1-\alpha)^s]\;,\qquad s=1,2,3.
1134: \end{equation}
1135: Since the negativity is given by
1136: \begin{equation} \label{E:negdef}
1137:     \sM(\rho_\alpha) = \sum_j \abs{\lambda_i} \;,
1138: \end{equation}
1139: we can find an upper bound on the negativity by maximizing
1140: $\sum_j\abs{\lambda_j}$ subject to the
1141: constraints~(\ref{E:sumsofpowers}).  If we consider only the $s=1,2$
1142: constraints, we obtain a nontrivial upper bound on the negativity with
1143: little effort.  We find that adding the constraint $s=3$ adds nothing
1144: asymptotically for large $N$, but for small $N$ yields a tighter bound
1145: than we get from the $s=1,2$ constraints, although this comes at the
1146: cost of considerably more effort.  We emphasize that these bounds apply
1147: to all bipartite divisions and to all unitaries $U_n$.  Notice that we
1148: have no reason to expect these bounds to be saturated, since the traces
1149: of higher powers of $\breve\rho_\alpha$ impose additional constraints
1150: that we are ignoring.  The one exception is the case of three qubits,
1151: where the $s=1,2,3$ constraints are a complete set, and indeed, in this
1152: case, the $s=1,2,3$ bound is $5/4$, which is saturated by the unitary
1153: found in Sec.~\ref{S:examples}.
1154: 
1155: The remainder of this section is devoted to calculating the $s=1,2$ and
1156: $s=1,2,3$ upper bounds.  A graphical summary of our results is for
1157: $\alpha=1$ presented in Fig.~\ref{F:totalfig}.
1158: 
1159: \begin{figure}[t]
1160: \begin{center}
1161: \includegraphics[scale=.6]{totalfig.eps}
1162: \caption{Plot of the bounds on the negativity of states of the
1163: form~(\ref{E:rhoout}), i.e., for a pure-state input in the zeroth
1164: register ($\alpha=1$).  The uppermost plot is the simple analytic bound
1165: $\sM_{1,2}=\sqrt2$, obtained using the $s=1,2$~constraint equations;
1166: the next largest plot is the numerically constructed $s=1,2,3$ bound.
1167: One can see that the $s=1,2,3$ bound asymptotes to the $s=1,2$ bound.
1168: As noted in the text, these bounds are independent of the unitary $U_n$
1169: and the bipartite division.  The flat line shows the negativity $5/4$
1170: for the state constructed in Sec.~\ref{S:examples}, currently the state
1171: of the form~(\ref{E:rhoout}) with the largest demonstrated negativity;
1172: notice that for $n+1=3$, this state attains the $s=1,2,3$ bound.  The
1173: lowest two sets of data points display the expected negativities for a
1174: randomly chosen unitary using the bipartite splittings $(n,1)$ and
1175: $\left(\left\lfloor n/2 \right\rfloor +1, \left\lceil n/2
1176: \right\rceil\right)$, which were also plotted in Fig.~\ref{F:random}.}
1177: \label{F:totalfig}
1178: \end{center}
1179: \end{figure}
1180: 
1181: \subsection{The $s=1,2$ Bound}
1182: 
1183: We can use Lagrange multipliers to reduce the problem to maximizing a
1184: function of one variable, but first we must deal with the absolute
1185: value in Eq.~(\ref{E:negdef}).  To do so, we assume that $t$ of the
1186: eigenvalues are negative and the $2N-t$ others are nonnegative, where
1187: $t$ becomes a parameter that must now be included in the maximization.
1188: We want to maximize
1189: \begin{equation}
1190: \sM_{1,2} = -\sum_{i=1}^t \lambda_i + \sum_{j=t+1}^{2N} \lambda_j\;,
1191: \end{equation}
1192:  subject to the constraints
1193: \begin{equation} \label{E:cons}
1194:     \sum_{k=1}^{2N} \lambda_k = 1
1195:     \quad \mbox{and} \quad \sum_{k=1}^{2N}
1196:     \lambda_k^2 = \frac{1+\alpha^2}{2N} \;.
1197: \end{equation}
1198: The notation we adopt here for the indices is that $i$ labels negative
1199: eigenvalues and $j$ labels nonnegative eigenvalues, while $k$ can label
1200: either.  This serves to remind us of the sign of an eigenvalue just by
1201: looking at its index.
1202: 
1203: Introducing Lagrange multipliers $\mu$ and $\nu$, the function we want
1204: to maximize is
1205: \begin{equation}
1206:     f(\lambda_k,t) =
1207:     -\sum_{i=1}^t \lambda_i + \sum_{j=t+1}^{2N} \lambda_j +
1208:     \mu\!\left(\sum_{k=1}^{2N} \lambda_k -1\right)
1209:     + \nu\!\left(\sum_{k=1}^{2N} \lambda_k^2 -\frac{1+\alpha^2}{2N}\right)\;.
1210: \end{equation}
1211: Differentiating with respect to $\lambda_i$ and then $\lambda_j$, we find
1212: \begin{eqnarray}
1213:     -1+\mu+2\nu \lambda_i = 0 \;, \\
1214:     +1+\mu+2\nu \lambda_j = 0 \;.
1215: \end{eqnarray}
1216: We immediately see that in the maximal solution, all the negative
1217: eigenvalues are equal, and all the nonnegative eigenvalues are equal.
1218: We can now reformulate the problem in the following way.  If we call
1219: the two eigenvalues $\lambda_-$ and $\lambda_+$, our new problem is to
1220: maximize
1221: \begin{equation} \label{E:newneg}
1222:     \sM_{1,2} = \sum_k \abs{\lambda_k}= -t\lambda_-+(2N-t)\lambda_+\;,
1223: \end{equation}
1224: subject to the constraints
1225: \begin{eqnarray} \label{E:constraint1}
1226:     t \lambda_- + (2N-t) \lambda_+ &=& 1\;,\\
1227:     t \lambda_-^2 + (2N-t) \lambda_+^2 &=& \frac{1+\alpha^2}{2N}\;.
1228:     \label{E:constraint2}
1229: \end{eqnarray}
1230: We can now do the problem by solving the constraints for $\lambda_-$
1231: and $\lambda_+$ in terms of $t$, plugging these results into
1232: $\sM_{1,2}$, and then maximizing over $t$.
1233: 
1234: Before continuing, we note two things.  First, $t$ cannot be $2N$, for
1235: if it were, then all the eigenvalues would be negative, making it
1236: impossible to satisfy Eq.~(\ref{E:constraint1}).  Second, unless
1237: $\alpha=0$, $t$ cannot be 0, for if it were, then all the eigenvalues
1238: would be equal to $1/2N$ by Eq.~(\ref{E:constraint1}), a situation
1239: Eq.~(\ref{E:constraint2}) says can occur only if $\alpha=0$.  Since we
1240: are not really interested in the case $\alpha=0$, for which
1241: $\rho_\alpha$ is always the maximally mixed state, we assume $\alpha>0$
1242: and $0<t<2N$ in what follows.
1243: 
1244: Solving Eqs.~(\ref{E:constraint1}) and (\ref{E:constraint2}) and
1245: plugging the solutions into Eq.~(\ref{E:newneg}), we get the two
1246: solutions
1247: \begin{equation}
1248: \label{E:ellipse}
1249:     \sM_{1,2} = \frac{N-t \pm \alpha \sqrt{t(2N-t)}}{N} \;.
1250: \end{equation}
1251: We choose the positive branch, since it contains the maximum.
1252: Maximizing with respect to $t$ treated as a continuous variable, we
1253: obtain the upper bound,
1254: \begin{equation} \label{E:maxneg}
1255:     \sM_{1,2} = \sqrt{1+\alpha^2}\, \stackrel{\alpha \to 1}{=}
1256:     \sqrt 2 \simeq 1.414\;,
1257: \end{equation}
1258: which occurs when the degeneracy parameter is given by
1259: \begin{equation}
1260: \label{E:degeneracy}
1261:     t = N\!\left(1 -\frac{1}{\sqrt{1+\alpha^2}} \right)
1262:     \stackrel{\alpha \to 1}{\simeq} 0.292\,N \;.
1263: \end{equation}
1264: The numbers on the right are for the case $\alpha = 1$, corresponding
1265: to the special qubit starting in a pure state.  Notice that the upper
1266: bound~(\ref{E:maxneg}) allows a negativity greater than 1 for all
1267: $\alpha$ except $\alpha=0$.
1268: 
1269: Since we did not yet enforce the condition that $t$ be a positive
1270: integer, the bound~(\ref{E:maxneg}) can be made tighter for specific
1271: values of $N$ and $\alpha$ by calculating $t$ and checking which of the
1272: two nearest integers yields a larger $\sM_{1,2}$.  Asymptotically,
1273: however, the ratio $t/N$ can approach any real number, so this bound
1274: for continuous $t$ is the same as the bound for integer $t$ in the
1275: limit $N\to\infty$.
1276: 
1277: \subsection{The $s=1,2,3$ Bound}
1278: 
1279: To deal with this case, we again make the assumption that $t$ of the
1280: eigenvalues are negative and $2N-t$ are nonnegative and thus write
1281: \begin{equation}
1282: \sM_{1,2,3} = -\sum_{i=1}^t \lambda_i + \sum_{j=t+1}^{2N} \lambda_j\;,
1283: \end{equation}
1284: as before.  In addition to the constraints~(\ref{E:cons}), we now have a
1285: third constraint
1286: \begin{equation}
1287: \label{E:constraint3}
1288:     \sum_{k=1}^{2N} \lambda_k^3 = \frac{1+3 \alpha^2}{4N^2} \;.
1289: \end{equation}
1290: We specialize to the case $\alpha = 1$ for the remainder of this
1291: subsection, because it is our main interest, and the algebra for
1292: the general case becomes difficult.
1293: 
1294: Introducing three Lagrange multipliers, we can write the function we
1295: want to maximize as
1296: \begin{equation}
1297: \label{E:lagrange123}
1298:     f(\lambda_k,t) =
1299:     -\sum_{i=1}^t \lambda_i + \sum_{j=t+1}^{2N} \lambda_j +
1300:     \mu\!\left( \sum_{k=1}^{2N} \lambda_k - 1 \right)+
1301:     \nu\!\left( \sum_{k=1}^{2N} \lambda_k^2 - \frac{1}{N} \right) +
1302:     \xi\!\left( \sum_{k=1}^{2N} \lambda_k^3 - \frac{1}{N^2} \right)\;.
1303: \end{equation}
1304: Differentiating with respect to $\lambda_i$ and then $\lambda_j$ gives
1305: \begin{eqnarray}
1306:     -1+\mu+2\nu \lambda_i + 3 \xi \lambda_i^2 &=& 0 \;, \\
1307:     +1+\mu+2\nu \lambda_j + 3 \xi \lambda_j^2 &=& 0 \;.
1308: \end{eqnarray}
1309: These equations being quadratic, we see that there are at most two
1310: distinct negative eigenvalues and at most two distinct nonnegative
1311: eigenvalues.  Since the sum of the two solutions of either of these
1312: equations is $-2\nu/3\xi$, however, we can immediately conclude either
1313: that one of the potentially nonnegative solutions is negative or that
1314: one of the potentially negative solutions is positive.  Hence, we find
1315: that at least one of the four putative eigenvalues has the wrong sign,
1316: implying that there are at most three distinct eigenvalues, though we
1317: don't know whether one or two of them are negative.
1318: 
1319: Labelling the three eigenvalues by $A$, $B$, and $C$, we can reduce the
1320: problem to solving the three constraint equations,
1321: \begin{eqnarray} \label{E:123}
1322:     u A + v B + w C & = & 1\;,\nonumber \\
1323:     u A^2 + v B^2 + w C^2 & = & 1/N\;,\\
1324:     u A^3 + v B^3 + w C^3 & = & 1/N^2\;,\nonumber
1325: \end{eqnarray}
1326: for $A$, $B$, and $C$ and then maximizing $\sM_{1,2,3}$ over the
1327: degeneracy parameters $u$, $v$, and $w$, which are nonnegative positive
1328: integers satisfying the further constraint
1329: \begin{equation}
1330: u+v+w=2N\;.
1331: \end{equation}
1332: We do not associate any particular sign with $A$, $B$, and $C$; the
1333: signs are determined by the solution of the equations.
1334: 
1335: One might hope that the symmetry of Eqs.~(\ref{E:123}) would allow for
1336: a simple analytic solution, but this appears not to be the case. In
1337: solving the three equations, one is inevitably led to a sixth-order
1338: polynomial in one of the variables, with the coefficients given as
1339: functions of $u$, $v$, and $w$.  Rather than try to solve this
1340: equation, which appears intractable, we elected to do a brute force
1341: optimization for any given value of $2N$ by solving Eqs.~(\ref{E:123})
1342: for each possible value of $u$, $v$, and $w$.  Picking the solution
1343: that has the largest negativity then yields the global maximum.  We did
1344: this for each $N$ up to $2N=78$.  The values of
1345: $u$, $v$, and $w$ that maximize the negativity are always
1346: \begin{equation} \label{E:uvw}
1347:     u=\left[N\left(1-\frac{1}{\sqrt 2}\right)\right] \;,
1348:     \ v=1 \;,
1349:     \ w = 2N-1-u \;,
1350: \end{equation}
1351: where $[x]$ denotes the integer nearest to $x$.  The unique eigenvalue
1352: corresponding to $v=1$ is the largest positive eigenvalue, $w$ is the
1353: degeneracy of another positive eigenvalue, and $u$ is the degeneracy of
1354: the negative eigenvalue.  Notice that the degeneracy of the negative
1355: eigenvalue is exactly what was found in the $s=1,2$ case.  Using the
1356: results~(\ref{E:uvw}) as a guide, we did a further numerical
1357: calculation of the maximum for larger values of $N$, by considering
1358: only the area around the degeneracy values given by
1359: Eq.~(\protect\ref{E:uvw}).  While this is not a certifiable global
1360: maximum, the perturbation expansion described below matches so well
1361: that the two are indistinguishable if they are plotted together for
1362: $n+1>7$.  This gives us confidence that the numerically determined
1363: upper bound $\sM_{1,2,3}$, which we plot in Fig.~\ref{F:totalfig}, is
1364: indeed a global maximum for all $N$.
1365: 
1366: We have used the numerical work to help formulate a perturbation
1367: expansion that gives the first correction to the $N\to\infty$ behavior
1368: of the $s=1,2,3$ bound.  Defining $x=1/N$, we rewrite the constraint
1369: equations~(\ref{E:123}) as
1370: \begin{eqnarray} \label{E:con}
1371:     a A + b B + c\,C & = & x\;,\nonumber \\
1372:     a A^2 + b B^2 + c\,C^2 & = & x^2\;,\\
1373:     a A^3 + b B^3 + c\,C^3 & = & x^3\;,\nonumber
1374: \end{eqnarray}
1375: where $a=u/N$, $b=v/N$, and $c=w/N$.  We also have the constraint
1376: \begin{equation} \label{E:con2}
1377: a+b+c=2\;.
1378: \end{equation}
1379: As $x$ is the variable that is asymptotically small, we seek an
1380: expansion in terms of it.
1381: 
1382: Our numerical work tells us that there are two positive eigenvalues,
1383: one of which is larger and nondegenerate.  In formulating our
1384: perturbation expansion, we let $B$ and $C$ be the positive eigenvalues,
1385: with $B$ being the larger one, having degeneracy $v=b_1\ge1$.  We do
1386: not assume that $b_1$ is 1, as the numerics show, but rather let the
1387: equations force us to that conclusion.  With this assumption, the form
1388: of the constraints~(\ref{E:con}) shows that the variables have the
1389: following expansions to first order beyond the $N\to\infty$ form:
1390: \begin{eqnarray} \label{E:degs}
1391:     a & = & a_0 + a_1 x^{1/3}\;,\nonumber \\
1392:     b & = & b_1x\;,\\
1393:     c & = & c_0 + c_1x^{1/3}\;,\nonumber
1394: \end{eqnarray}
1395: and
1396: \begin{eqnarray} \label{E:evs}
1397:     A & = & A_0 x + A_1 x^{4/3}\;,\nonumber \\
1398:     B & = & B_1 x^{2/3}\;,\\
1399:     C & = & C_0 x + C_1 x^{4/3}\;.\nonumber
1400: \end{eqnarray}
1401: We see that we are actually expanding in the quantity $y=x^{1/3}$.  In
1402: terms of these variables, the negativity is given by
1403: \begin{eqnarray}
1404:  \sM_{1,2,3}&=&
1405:  \frac{-a A + b B + c C}{x}\nonumber\\
1406:  &=&-a_0A_0+c_0C_0+(-a_0A_1-a_1A_0+c_0C_1+c_1C_0)x^{1/3}+O(x^{2/3})\;,
1407: \end{eqnarray}
1408: which we now endeavor to maximize.
1409: 
1410: Substituting Eqs.~(\ref{E:degs}) and (\ref{E:evs}) into the
1411: constraints~(\ref{E:con}) and (\ref{E:con2}) and equating terms with
1412: equal exponents of $x$, we obtain, to zero order,
1413: \begin{eqnarray}
1414:     a_0 + c_0 & = & 2\;,\nonumber \\
1415:     a_0 A_0 + c_0 C_0 & = & 1\;,\\
1416:     a_0 A_0^2+ c_0 C_0^2 & = & 1 \;.\nonumber
1417: \end{eqnarray}
1418: Solving for $a_0$, $c_0$, and $C_0$ in terms of $A_0$ and substituting
1419: the results into the zero-order piece of $\sM_{1,2,3}$ gives
1420: \begin{equation}
1421:  \label{E:neg1}
1422:  \sM_{1,2,3}=\frac{1-4A_0+2A_0^2}{1-2A_0+2A_0^2}\;.
1423: \end{equation}
1424: Maximizing Eq.~(\ref{E:neg1}) gives $A_0^2=1/2$ and, hence,
1425: $A_0=-1/\sqrt{2}$, since $A$ is the negative eigenvalue.  This leads to
1426: $a_0=1-1/\sqrt{2}$, $c_0 = 1+1/\sqrt{2}$, and $C_0=1/\sqrt{2}$, and the
1427: resulting $N\to\infty$ upper bound is $\sM_{1,2,3}=\sqrt2$, as
1428: expected.
1429: 
1430: If we carry this process out to first order beyond the $N\to\infty$
1431: behavior, we obtain, after some algebraic manipulation, $\sM_{1,2,3} =
1432: \sqrt2-b_1^{1/3}x^{1/3}/2^{7/6}+O(x^{2/3})$.  Maximizing this simply
1433: means making $b_1$ as small as possible, i.e., choosing $b_1=1$, whence
1434: we obtain the following asymptotic expression for the $s=1,2,3$ upper
1435: bound:
1436: \begin{equation}
1437: \sM_{1,2,3} = \sqrt2 - \frac{1}{2^{7/6}N^{1/3}} +
1438: O\!\left(\frac{1}{N^{2/3}}\right)\;.
1439: \end{equation}
1440: This shows that the upper bound of $\sqrt 2$ is approached
1441: monotonically from below in the asymptotic regime.  In addition, the
1442: procedure verifies that in the maximum solution, the largest positive
1443: eigenvalue is nondegenerate.  For the case of qubits we have $N=2^n$,
1444: implying that the approach to the $N\to\infty$ bound is exponentially
1445: fast.
1446: 
1447: \section{Conclusion} \label{S:conclusion}
1448: 
1449: The mixed-state quantum circuit~(\ref{E:circuitalpha}) provides an
1450: efficient method for estimating the normalized trace of a unitary
1451: operator, a task that is thought to be exponentially hard on a
1452: classical computer.  If one believes that global entanglement is the
1453: essential resource for the exponential speedup achieved by quantum
1454: computation, then the question begging to be answered is whether there
1455: is any entanglement in the circuit's output
1456: state~(\ref{E:rhooutalpha}). The purpose of this paper was to
1457: investigate this question.
1458: 
1459: A notable feature of the circuit~(\ref{E:circuitalpha}) is that it
1460: provides an efficient method for estimating the normalized trace no
1461: matter how small the initial polarization $\alpha$ of the special qubit
1462: in the zeroth register, as long as that polarization is not zero.
1463: Since all the other qubits are initially completely unpolarized, we are
1464: led to characterize the computational power of this circuit as the
1465: ``power of even the tiniest fraction of a qubit.''  We provide
1466: preliminary results regarding the entanglement that can be achieved for
1467: $\alpha<1$. Our results are consistent with, but certainly do not
1468: demonstrate the conclusion that separable states cannot provide an
1469: exponential speedup and that entanglement is possible no matter how
1470: small $\alpha$ is.  The question of entanglement for subunity
1471: polarization of the special qubit deserves further investigation.
1472: 
1473: Our key conclusions concern the case where the special qubit is
1474: initially pure ($\alpha=1$).  We find that the
1475: circuit~(\ref{E:circuit}) typically does produce global entanglement,
1476: but the amount of this entanglement is quite small.  Using
1477: multiplicative negativity to measure the amount of entanglement, we
1478: show that as the number of qubits becomes large, the multiplicative
1479: negativity in the state~(\ref{E:rhoout}) is a vanishingly small
1480: fraction of the maximum possible multiplicative negativity for roughly
1481: equal splittings of the qubits.  This hints that the key to
1482: computational speedup might be the global character of the
1483: entanglement, rather than the amount of the entanglement.  In the
1484: spirit of the pioneering contribution of Wyler~\cite{Wyler74}, what
1485: happier motto can we find for this state of affairs than {\it Multum ex
1486: Parvo}, or A Lot out of A Little.
1487: 
1488: 
1489: \begin{acknowledgments}
1490: The authors thank H.~N. Barnum, A.~Denney, B.~Eastin, K.~Manne, N.~C.
1491: Menicucci, and A.~Silberfarb for useful
1492: discussions.  The Matlab code used to calculate the results of
1493: Sec.~\ref{S:random} made use of T.~S.~Cubitt's freely available
1494: algorithm for taking the partial transpose of a matrix; this and other
1495: useful algorithms written by Cubitt are available at the website
1496: \href{http://www.dr-qubit.org/matlab.html}{http://www.dr-qubit.org/matlab.html}.
1497: The quantum circuits in this paper were typeset using Qcircuit, which
1498: is freely available online at
1499: \href{http://info.phys.unm.edu/Qcircuit}{http://info.phys.unm.edu/Qcircuit}.
1500: This work was supported in part by US Army Research Office Contract
1501: No.~W911NF-04-1-0242.
1502: \end{acknowledgments}
1503: 
1504: \appendix
1505: \section{Proof of the Lemma}{\label{A:appendix}}
1506: 
1507: \textbf{Lemma:} $\tr(\breve A \breve B) = \tr(A B) $.\\
1508: 
1509: \textbf{Proof:} Define two operators $A$ and $B$ by
1510: \begin{equation}
1511: A=\sum_{i,j,k,l} a_{ij,kl} \ket{ij}\!\bra{kl}\;,
1512: \quad
1513: B=\sum_{m,n,p,q} b_{mn,pq} \ket{mn}\!\bra{pq} \;.
1514: \end{equation}
1515: Taking the partial transpose with respect to the second subsystem, we
1516: find
1517: \begin{equation}
1518: \breve A=\sum_{i,j,k,l} a_{ij,kl} \ket{il}\!\bra{kj}\;,
1519: \quad
1520: \breve B=\sum_{m,n,p,q} b_{mn,pq} \ket{mq}\!\bra{pn} \;.
1521: \end{equation}
1522: Calculating the quantities of interest, we find that they are indeed equal.
1523: \begin{eqnarray}
1524:     \tr(\breve A \breve B) &=&
1525:         \sum_{\scriptsize\begin{array}{l}i,j,k,l, \\ m,n,p,q \end{array}}
1526:         a_{ij,kl} b_{mn,pq} \braket{pn}{il} \braket{kj}{mq}
1527:     = \sum_{i,j,k,l} a_{ij,kl} b_{kl,ij} \;, \\
1528:     \tr\left(A  B\right) &=&
1529:         \sum_{\scriptsize \begin{array}{l}i,j,k,l, \\ m,n,p,q \end{array}}
1530:         a_{ij,kl} b_{mn,pq} \braket{pq}{ij} \braket{kl}{mn}
1531:     = \sum_{i,j,k,l} a_{ij,kl} b_{kl,ij}\;.
1532: \end{eqnarray}
1533: 
1534: \begin{references}
1535: 
1536: \bibitem{DiVincenzo00} D.~DiVincenzo, Fortschr.\ Phys. {\bf 48}, 9 (2000).
1537: 
1538: \bibitem{Jones01} J.~A. Jones, Prog. NMR Spectroscopy {\bf 38}, 325 (2001).
1539: 
1540: \bibitem{Schulman99} L.~J. Schulman and U.~V. Vazirani, in Proceedings
1541: of the 31st Annual ACM Symposium on Theory of Computing (ACM Press, New
1542: York, 1999), p.~322 (extended version available as {\tt arXiv.org
1543: e-print \href{http://arxiv.org/abs/quant-ph/9804060}{quant-ph/9804060}}).
1544: 
1545: \bibitem{kl98} E.~Knill and R.~Laflamme, Phys.\ Rev.\ Lett. {\bf 81}, 5672 (1998).
1546: 
1547: \bibitem{Ambainis00} A.~Ambainis, L.~J. Schulman, and U.~V. Vazirani,
1548: in Proceedings of the 32nd Annual ACM Symposium on Theory of Computing
1549: (ACM Press, New York, 2000), p.~697.
1550: 
1551: \bibitem{Laflamme02} R.~Laflamme, D.~G. Cory, C.~Negrevergne, and L.~Viola,
1552: Quant.\ Inf.\ Comp.\ {\bf 2}, 166 (2002).
1553: 
1554: \bibitem{pklo04} D.~Poulin, R.~Blume-Kohout, R.~Laflamme, and H.~Ollivier,
1555: Phys.\ Rev.\ Lett.\ {\bf 92}, 177906 (2004).
1556: 
1557: \bibitem{elpc04} J.~Emerson, S.~Lloyd, D.~Poulin, and D.~Cory, Phys.\ Rev.~A
1558: {\bf 69}, 050305(R) (2004).
1559: 
1560: \bibitem{Josza99} R.~Josza and N.~Linden, Proc.~Roy.~Soc.~Lond.~A {\bf 459},
1561: 2011 (2003).
1562: 
1563: \bibitem{N&C} M.~Nielsen and I.~Chuang, {\it Quantum Computation and Quantum
1564: Information\/} (Cambridge University Press, Cambridge, England, 2000).
1565: 
1566: \bibitem{b99} S.~L. Braunstein, C.~M. Caves, R.~Jozsa, N.~Linden, S.~Popescu,
1567: and R.~Schack, Phys.\ Rev.\ Lett.\ {\bf 83}, 1054 (1999).
1568: 
1569: \bibitem{gb03} L.~Gurvits and H.~Barnum, Phys.\ Rev.~A {\bf 68},
1570: 042312 (2003).
1571: 
1572: \bibitem{gb04} L.~Gurvits and H.~Barnum, {\tt arXiv.org e-print
1573: \href{http://arxiv.org/abs/quant-ph/0409095}{quant-ph/0409095}}.
1574: 
1575: \bibitem{dur99} W.~D{\"u}r, J.~I. Cirac, and R.~Tarrach, Phys.\ Rev.\ Lett.\
1576: {\bf 83}, 3562 (1999).
1577: 
1578: \bibitem{p96} A.~Peres,  Phys.\ Rev.\ Lett.\ {\bf 77}, 1413 (1996).
1579: 
1580: \bibitem{hhh96} M.~Horodecki, P.~Horodecki, R.~Horodecki, Phys.~Lett.~A
1581: {\bf 223}, 1 (1996).
1582: 
1583: \bibitem{Vidal02} G.~Vidal and R.~F. Werner, Phys.\ Rev.~A {\bf 65}, 032314
1584: (2002).
1585: 
1586: \bibitem{dhhmno05} C.~M.~Dawson, H.~L.~Haselgrove, A.~P.~Hines, D.~Mortimer,
1587: M.~A.~Nielsen, and T.~J.~Osborne, Quant.\ Inf.\ Comp.\ {\bf 5}, 102 (2005).
1588: 
1589: \bibitem{ek90} A.~Ehrenfeucht and  M.~Karpinski, Technical Report~tr-90-033,
1590: International Computer Science Institute, Berkeley, CA (1990),
1591: available online at {\tt\href{http://www.citeseer.com}{http://www.citeseer.com}}.
1592: 
1593: \bibitem{Plenio05} M.~B.~Plenio, {\tt arXiv.org e-print
1594: \href{http://arxiv.org/abs/quant-ph/0505071}{quant-ph/0505071}}.
1595: 
1596: \bibitem{SLee03}
1597: S.~Lee, D.~P.~Chi, S.~D.~Oh, and J.~Kim, Phys.\ Rev.~A {\bf 68}, 062304 (2003).
1598: 
1599: \bibitem{Conway90} J.~Conway, {\it A Course in Functional Analysis}
1600: (New York, Springer-Verlag, 1990).
1601: 
1602: \bibitem{Hurwitz1897} A.~Hurwitz, Nachr.\ Ges.\ Wiss.\ G\"ott.\ Math.-Phys.~{\bf Kl},
1603: 71 (1897).  The Hurwitz decomposition is discussed in the appendix of
1604: M.~Po\'zniak, K.~\.Zyczkowski, and M.~Ku\'s, J.~Phys.~A:
1605: Math.~Gen.~{\bf 31}, 1059 (1998).
1606: 
1607: \bibitem{Emerson03} J.~Emerson, Y.~S.~Weinstein, M.~Saraceno, S.~Lloyd,
1608: and D.~G.~Cory, Science {\bf 302}, 2098 (2003).
1609: 
1610: \bibitem{Scott03} A.~J.~Scott and C.~M.~Caves, J.~Phys.~A {\bf 36}, 9553
1611: (2003).
1612: 
1613: \bibitem{Wyler74} J.~A.~Wyler, Gen.~Rel.~Grav. {\bf 5}, 175 (1974).
1614: 
1615: 
1616: \end{references}
1617: 
1618: \end{document}
1619: