1: %\documentstyle[epsf, epsfig, psfig, 11pt]{article}
2: \documentclass[11pt]{article}
3: \usepackage{epsf,psfig,epsfig,amsmath}
4:
5: \newcommand{\hcenterpage}[1]{
6: \setlength{\textwidth}{#1}
7: \setlength{\evensidemargin}{3.25in}
8: \addtolength{\evensidemargin}{-0.5\textwidth}
9: \setlength{\oddsidemargin}{\evensidemargin}
10: \setlength{\marginparwidth}{1.74cm}
11: \addtolength{\marginparwidth}{-\marginparsep}
12: \addtolength{\marginparwidth}{\evensidemargin}
13: }
14:
15: \newcommand{\vcenterpage}[1]{
16: \setlength{\textheight}{#1}
17: \setlength{\topmargin}{4.3in} % was 3.8in
18: \addtolength{\topmargin}{-0.5\textheight}
19: \setlength{\topskip}{0cm}
20: % \setlength{\parskip}{0.0ex plus0ex minus0ex}
21: }
22:
23: \renewcommand{\baselinestretch}{1.1}
24:
25: %\input{defs}
26:
27: \hcenterpage{6.75in}
28: \vcenterpage{9.25in}
29: %\thispagestyle{empty}
30: \title{Simulations of the adiabatic quantum optimization for the Set Partition Problem.}
31:
32: \vskip-0.15in
33:
34:
35:
36: \author{V.N. Smelyanskiy$^*$, U. V. Toussaint, D.A. Timucin \\
37: Computational Science Division\\ NASA Ames Research Center\\
38: $^*\,$ vadim@quantum.arc.nasa.gov}
39:
40: \begin{document}
41: \maketitle
42: \date{}
43:
44: \begin{abstract}
45: We analyze the complexity of the quantum optimization
46: algorithm based on adiabatic evolution for the NP-complete set partition
47: problem. We introduce a cost function defined on a logarithmic
48: scale of the partition residues so that the total number of
49: values of the cost function is of the order of the problem size.
50: We simulate the behavior of the algorithm by numerical
51: solution of the time-dependent Schr{\" o}dinger equation as well
52: as the stationary equation for the adiabatic eigenvalues. The
53: numerical results for the time-dependent quantum evolution
54: indicate that the complexity of the algorithm scales exponentially
55: with the problem size. This result appears to contradict the recent
56: numerical results for complexity of quantum adiabatic algorithm applied to
57: a different NP-complete problem (Farhi et al, {\it Science} 292, p.472 (2001)).
58: \end{abstract}
59:
60: \renewcommand{\baselinestretch}{1.5}
61: \normalsize
62:
63: \thispagestyle{empty}
64: \section{Introduction}
65: \label{sec:Intro}
66:
67:
68: Most common computationally intensive tasks encountered in
69: practice may be formulated as combinatorial optimization problems
70: (COPs), many of which are found to belong to the algorithmic class
71: \emph{nondeterministic-polynomial complete} (NP-complete)
72: \cite{Garey}. The NP-complete problems are computationally hard -
73: they are characterized (in the worst cases) by exponential scaling
74: of the running time or memory requirements with the problem size.
75: A special property of the class is that any NP-complete problem
76: can be converted into any other NP-complete problem in polynomial time on
77: a classical computer; therefore, it is sufficient to find a
78: deterministic algorithm that can be guaranteed to solve all instances
79: of just one of the NP-complete problems within a polynomial time
80: bound.
81:
82: An instance of a COP of size $n$ may be encoded using bit
83: strings $\textsf{z} = z_0 \, z_1 \, \cdots \, z_{n-1}$, $\,\,z_j
84: = 0, 1$, with a corresponding value of the cost function (or
85: ``energy'') $E = E_{\textsf{z}}$ for each string. The objective
86: is to find the bit string(s) with the minimum cost (and the
87: corresponding cost value). In quantum computation, bits $z_j$ are
88: replaced by spin-$\frac{1}{2}$ \emph{qubits}; the qubit states
89: $|0\rangle$ and $|1\rangle$ are eigenstates of the $\pm z$
90: component of the $j$-th spin, respectively. The Hilbert space of
91: a quantum \emph{register} with $n$ qubits is spanned by $N = 2^n$
92: basis vectors $|\textsf{z}\rangle = |z_0\rangle \otimes \cdots
93: \otimes |z_{n-1}\rangle$.
94:
95: \section{Optimization by Adiabatic Quantum Evolution}
96: \label{sec:opt}
97:
98: Following \cite{Farhi,Farhiapp,Cerf}, we consider a quantum evolution of duration $T$ based on the time-dependent
99: Hamiltonian
100: \begin{equation}
101: H(t) = \alpha(s)\, V + \beta(s)\, H_{P}, \quad s\equiv s(t). \label{farhi}
102: \end{equation}
103: \noindent
104: Here, $H_P = \sum_{\textsf{z}} E_{\textsf{z}}
105: |\textsf{z}\rangle \langle \textsf{z}|$ is the ``problem''
106: Hamiltonian that embodies the problem structure in its energy
107: spectrum and eigenstates, the summation being performed over all $N$ $n$-bit
108: strings, and $V$ is a ``driver'' Hamiltonian that is constructed
109: in such a way as to cause transitions between those states -
110: essentially an Ising-type spin Hamiltonian corresponding to 1- and
111: 2-gate operations:
112: \begin{equation}
113: V = -\sum_{i=0}^{n-1} B_i \sigma_{x}^{i} - \frac{1}{2} \sum_{i, \,
114: j=0}^{n-1} J_{ij} \sigma_{x}^{i} \sigma_{x}^{j}. \label{driver}
115: \end{equation}
116:
117: Coefficients $\alpha\left(s\left(t\right)\right)$ and $\beta(s(t))$ vary in time in such
118: a way that at the initial instant of time $H(0)=V$ and at the
119: final instant $H(T)=H_P$. A particular choice of the coefficients
120: is \cite{Farhi}
121: \begin{equation}
122: \alpha(s)=1-s, \quad \beta(s)=s, \quad s(t)=t/T\label{alpha}
123: \end{equation}
124: \noindent The total Hamiltonian (\ref{farhi}) produces a
125: nontrivial quantum evolution from some initial (superposition)
126: state $\psi(0)$ to a final (solution) state $\psi(T)$. If no
127: knowledge about the solution is available \emph{a priori}, then
128: the initial state may be chosen as the symmetric state (cf.
129: \cite{Grover,Farhi})
130: \begin{equation}
131: \psi(0) =
132: 2^{-n/2} \sum_{\textsf{z}=0}^{2^{n}-1} |\textsf{z}\rangle. \label{ini}
133: \end{equation}\noindent
134: This choice is appropriate provided (\ref{ini}) is a ground state
135: of $V$ (e.g., $B_i, J_{ij} \geq 0$).
136: Now, if $T$ is sufficiently large, then functions $\alpha(t/T)$ and $\beta(t/T)$
137: vary in time slowly and the system will remain in the
138: instantaneous (adiabatic) ground state of $H(t)$ during its
139: entire evolution $0 < t < T$ (cf. \cite{Farhi}). Accordingly,
140: $\psi(T)$ will be a superposition of states $|z\rangle$ corresponding to
141: the ground state of $H_P$. It is clear that in this case a measurement
142: performed on the quantum register at $t = T$ will find with
143: certainty one of the solutions of COP. In this case the complexity
144: of the quantum algorithm is determined by its duration $T$. If we
145: expand the wavefunction of the system $\psi(t)$ in the basis of
146: the adiabatic eigenfunctions $\Psi_k(t)$ of the Hamiltonian $H(t)$
147: \begin{eqnarray}
148: &&H(t) \Psi_k(t) = g_k(t) \Psi_k(t), \quad k=0,1,\ldots,2^{n}-1.\label{adiab} \\
149: &&\psi(t)=\sum_{k=0}^{2^n-1}\, C_k(t) \Psi_k(t) \exp\left(-i/\hbar \int^{t}_{0} dt' g_k(t')\right),
150: \label{corr} \\
151: \end{eqnarray}
152: \noindent
153: then adiabatic approximation corresponds to $\psi(t) \propto \Psi_0(t)$ (up to the oscillating
154: phase factor). Coefficients $C_k(t)$ with $k>0$ correspond to nonadiabatic corrections.
155: Using perturbation theory in the basis of eigenfunctions $\Psi_k(t)$ the total probability $p_{n-ad}(t)$
156: of {\itshape not\/} finding the system at the instant $t$ in its adiabatic ground state equals
157: \begin{eqnarray}
158: && p_{n-ad}(t)=\left(\frac{\dot {s}(t)}{s(t)}\right)^2\,
159: \sum_{k=1}^{2^n-1}\, \frac{|V_{0k}|^{2}}{[g_k(t)-g_0(t)]^4}
160: \label{nonad} \\
161: && V_{0k}=\langle \Psi_0|V|\Psi_k\rangle.\nonumber
162: \end{eqnarray}
163: \noindent Here we used the explicit form of coefficients
164: $\alpha,\,\beta$ given in (\ref{alpha}). It is seen that during
165: the quantum evolution coefficients $C_k(t)\sim \dot s/s \sim 1/T$
166: and the largest admixture of the exited states into the total
167: superposition occurs at the instant of time when one of
168: the exited levels $g_k(t)$ closely approaches the ground state
169: (avoided-crossing). From here the overall criterion for the
170: adiabatic evolution can be expressed in the well-known form
171: \begin{equation}
172: \eta = {\tilde{V}\over T \Delta g_{min}^2} \ll 1, \label{criterion}
173: \end{equation}
174: \noindent where $\Delta g_{min}$ is the closest approach of the
175: ground state to one of the excited states during the evolution - a
176: minimum gap- and $\tilde V$ is the characteristic energy scale for
177: the matrix elements of $V$. We note that although instantaneous
178: nonadiabatic corrections (\ref{nonad}) are quadratic in the
179: parameter $\eta$ (near the avoided crossing) the probability of
180: nonadiabatic transitions $W_{n-ad}$ away from the ground state is
181: exponentially small in $\eta$ \cite{Landau}. This probability is
182: defined on an infinite time axis and its logarithm is proportional
183: to the imaginary part of the integral along the contour in the
184: complex plane of $t$ that begins and ends on the real time axis and
185: loops around the complex branching point $t^*$
186: \begin{equation}
187: W_{n-ad}~\exp\left(-|{\rm Im} \oint\, g_0(t) dt|\right).
188: \label{imagin}
189: \end{equation}
190: \noindent Here $t^*$ correspond to one of the roots of the
191: equation
192: \begin{equation}
193: g_k(t^*)=g_0(t^*)\label{roots}
194: \end{equation}
195: \noindent that provides the smallest value for the exponential in
196: (\ref{imagin}) (out of all possible complex solutions of
197: (\ref{imagin}) for different excited states $\Psi_k$). In a
198: standard (Landau-Zener) theory of nonadiabatic transitions the value
199: of the exponent is approximately of the order of the parameter
200: $\eta$ in (\ref{criterion}), and therefore it is the size of the
201: minimum gap $\Delta g_{min}$ that determines the condition for {\itshape T\/}
202: and hence the complexity of the quantum adiabatic search algorithm
203: according to \cite{Farhi}. We note finally that, as pointed out in
204: \cite{Cerf}, the improved complexity of the adiabatic algorithm is
205: determined by the instantaneous rate $\dot s(t)/s(t)$ of the variation
206: of the control parameter $s(t)$ near the avoided crossing. We will
207: not discuss in this paper such modifications and focus primarily on
208: intrinsic properties of the quantum system in question.
209:
210: \section{Set Partition Problem}
211: \label{sec:spp}
212:
213: In this paper, we will analyze the complexity of the adiabatic quantum
214: optimization for the \emph{set partition problem}
215: (SPP), which is one of the basic NP-complete problems of
216: theoretical computer science \cite{Garey}. The optimization
217: version of SPP is to partition a set of $n$ positive integers
218: $\{\alpha_0, a_1, \ldots, \alpha_{n-1}\}$ into two disjoint
219: subsets ${\mathcal A}_1$ and ${\mathcal A}_2$ such that the
220: ``residue'' $|\sum_{\alpha_j \in {\mathcal A}_1} \alpha_j -
221: \sum_{\alpha_j \in {\mathcal A}_2} \alpha_j|$ is minimized. The
222: complexity of the problem substantially depends on the size of the
223: integers $\alpha_j$ (see below). It is often customary for the
224: analysis of the random instances of the problem to introduce
225: finite-precision rational numbers $a_j$ that are independently and
226: identically distributed (i.i.d.) in the unit interval $(0,1]$.
227: \begin{equation}
228: \alpha_{j} \leq 2^{b} \,\,\,\,\forall j, \quad a_{j} = 2^{-b}
229: \alpha_{j} \in (0, 1]
230: \end{equation}
231: \noindent Here $b$ is the total number of bits used to represent the
232: numbers $a_j$. The values of the partition can be encoded in
233: binaries by attaching ``sign'' bits $s_j$ to the numbers $a_j$. The
234: partition residue can be defined as $|\Omega_{\textsf{z}}|$ where
235: \begin{equation}
236: \Omega_{\textsf{z}} = \sum_{j=0}^{n-1} s_j a_j,\quad s_j = 1 - 2
237: z_j = \pm 1 \quad (z_j =0,1)\label{omega}
238: \end{equation}
239: \noindent Here $\Omega_{\textsf{z}}$ is a {\em signed} partition
240: residue. We note that by definition the problem is symmetric: two
241: bit strings that can be obtained from each other by flipping all
242: the bits ($s_j \rightarrow -s_j$) correspond to two values of
243: $\Omega_{\textsf{z}}$ that differ only in sign. We note that the
244: minimum-residue partition(s) may be thought of as the ground
245: state(s) of the following spin Hamiltonian \cite{Mertens}
246: \begin{equation}
247: \Omega_{\textsf{z}}^{2} = \sum_{i, \, j=0}^{n-1} a_i a_j s_i s_j.
248: \label{cost0}
249: \end{equation}
250: \noindent This is an infinite range Ising spin glass with Mattis
251: type antiferromagnetic coupling, $J_{ij}=-a_i\, a_j$. Infinite
252: range coupling clearly represents a major problem with direct
253: (`analog') physical implementation of this Hamiltonian on a quantum
254: computer. Therefore one can consider using an oracle-type cost
255: function $E(\textsf{z})=|\Omega_{\textsf{z}}|$ to implement the
256: problem Hamiltonian in (\ref{farhi}) for SPP. The corresponding
257: unitary transformation will multiply the basis states $|z\rangle$
258: by phase factors $\exp\left(-i\Delta t \,E(\textsf{z})\right)$
259: during the elementary discrete steps of the `continuous-time'
260: adabatic quantum optimization (\ref{farhi}). Although this approach
261: is natural for the satisfiability problem \cite{Farhi} it has a
262: serious limitation for SPP (as well as some other NP-complete
263: problems like integer programming, where the precision of
264: integers is of central importance). To demonstrate this point
265: we need to consider the density of states of the partition residues.
266:
267: \subsubsection{Density of states}
268:
269: We define the density of states for a given instance of SPP as
270: follows
271: \begin{equation}
272: \rho(\Omega) = \sum_{\textsf{z}} \delta(\Omega -
273: \Omega_{\textsf{z}}).\label{density}
274: \end{equation}
275: \noindent The exact form of $\rho(\Omega)$ depends on a given
276: instance of SPP (i.e., a particular set of numbers $a_j$). However
277: we introduce a {\em coarse-grained} density of states
278: \begin{equation}
279: {\bar \rho}(\Omega) = \frac{1}{\Delta \Omega}
280: \int_{\Omega}^{\Omega + \Delta \Omega} \rho (\zeta) \mathrm{d}
281: \zeta \label{cg}
282: \end{equation}
283: \noindent where averaging is over an interval of $\Delta \Omega$
284: whose size will be determined below. Using (\ref{omega}) and
285: (\ref{density}) we can rewrite this expression in the form
286: \begin{equation}
287: {\bar \rho}(\Omega) = \frac{2^n}{2 \pi} \int_{-\infty}^{\infty}
288: {e^{i w (\Omega+\Delta\Omega)}-e^{i w \Omega} \over i w}\,I(w)\,
289: \mathrm{d} w, \quad\quad I(w)= \prod_{j=0}^{n-1} \cos(a_j w).
290: \label{cos}
291: \end{equation}
292: \noindent
293: Note that $I(\pi k 2^b)=(-1)^k,\,k=0,\pm1,\ldots$ and $I(w)$
294: has very sharp maxima (minima) at those points. In their
295: vicinities the integral in (\ref{cos}) can be evaluated by
296: steepest descent method for any given problem instance. The sum over
297: the contributions from different saddle points was obtained by
298: Mertens \cite{Mertens} in his derivation of the partition function
299: for the corresponding spin glass model. We emphasize however that
300: $I(w)$ can have multiple sharp resonances at the intermediate
301: points $|w|<2^{-b}$. The positions of these resonances are at the
302: multiples of $\pi/q$ where $q$ is an {\it approximate} greatest
303: common divisor (g.c.d.) of
304: the set of $n$ numbers $a_j$ such that $a_j = f_j q+ r_j$ where
305: $f_j$ are integers and $r_j$ are residues of the division. Provided
306: that most of the residues are sufficiently small
307: \[
308: {\pi^{2}\over 2} \sum_{j=1}^{p} r_{j}^{2} \leq 1, \quad p \sim n,
309: \]
310: the function $I(w)$ will have steep peaks at those points. It can
311: be shown that in the general case the value of the approximate g.c.d.
312: for a set of $n$ $b-bit$ numbers inside the unit interval
313: scales as $2^{-n}$ for $n<b$. Obviously it equals $2^{-b}$ for
314: $n>b$. In what follows we will be interested in the
315: high-precision case $n<b$. We choose the size of the averaging
316: window $\Delta \Omega \gg 2^{-n}$ and this introduces a cut-off
317: in the integral (\ref{cost0}) at
318: \[|w| \ll \pi/\Delta\Omega \ll 2^{n}\]
319: It follows from above that in this case the values of the g.c.d. will
320: lie outside the cutoff and corresponding resonances will not
321: contribute to the integral. The value of the integral can be
322: estimated near the single remaining maximum at $w=0$. The width of
323: the maximum near that point is ~ $\delta w ~ n^{-1/2} \ll 1$ and
324: therefore the window function in (\ref{omega}) works as a step
325: function in that region. Finally we obtain
326: \begin{equation}
327: {\bar \rho}(\Omega) = {2^n \over \sqrt{2 \pi n \sigma^2}} \exp
328: \left(-{\Omega^2 \over 2 n \sigma^2} \right),\qquad
329: \sigma^2 = \frac{1}{n} \sum_{j=0}^{n-1} a_j^2.
330: \label{density1}
331: \end{equation}
332: \noindent Here the variance $\sigma$ is
333: a ``self-averaging'' quantity, and the coarse-graining is
334: performed over an interval much larger than the characteristic
335: separation between neighboring partition residue values
336: \begin{equation}
337: \Delta E\, \sim \, \sqrt{n}\,\, 2^{-n} \end{equation} \noindent
338: yet much smaller than the scale of variation of ${\rho}(\Omega)$:
339: $\Delta E
340: \ll \Delta \Omega \ll \sqrt{n}$. We note that in the
341: high-precision regime ($n < b$), partition residues
342: $\Omega_{\textsf{z}}$ are irregularly spaced and well separated
343: from each other (on the scale of $2^{-b}$). However this structure
344: is being averaged out in (\ref{density1}) and the result indicates
345: that, in general, no more structure exists on a scale
346: $\gg 2^{-n}$ other than that given by the Gaussian distribution in
347: (\ref{density1}). We note that this distribution is usually
348: obtained for the SPP using averaging over different instances of
349: the problem (cf. \cite{Fu,Ferreira}, \cite{Mertens}(\emph{b}));
350: here we recovered it as a coarse-grained distribution for a {\em
351: given} instance which is more consistent with our goal of studying
352: the complexity of the adiabatic quantum optimization algorithm \cite{note2}.
353:
354: \subsection{Cost function}
355: Computational complexity of SPP depends critically on the number
356: of bits $b$: numerical simulations with independent identically
357: distributed (i.i.d.) random $b$-bit numbers $a_j$ show
358: \cite{Walsh,Korf} that the solution time grows exponentially with
359: $n$ for $2^{n - b} \ll 1$ (high-precision, computationally `hard
360: phase'), and polynomially for $2^{n-b} \gg 1$ (low-precision,
361: computationally `easy phase'), exhibiting a behavior similar to a
362: phase transition \cite{Mertens}(\emph{a}).
363:
364: In the low-precision phase, values of $\Omega_{\textsf{z}}$ are
365: equally spaced (in $2^{-b}$) and strongly degenerate each
366: corresponding to (roughly) $2^{b-n} \gg 1$ number of
367: bit-strings. This degeneracy grows exponentially with n if $b$
368: remains fixed. The total number of solutions with zero residues
369: accumulate correspondingly and this is why the complexity
370: eventually becomes polynomial in $n$. The quantum algorithm suggested
371: in \cite{Raedt} directly computes the density of states (\ref{density}) of the SPP
372: and is efficient in finding the number of
373: solutions in the low-precision case. In this case it is also
374: feasible to use a cost function
375: $E(\textsf{z})=|\Omega_{\textsf{z}}|$ (provided the number of
376: possible values does not grow exponentially with $n$).
377:
378: The situation is qualitatively different in the high-precision
379: case. Implementation of the approach based on the above cost
380: function will require a quantum computer using exponentially high
381: precision physical parameters (external fields, etc) to control
382: small differences {\em in the phases} of unitary transformations on
383: the scale at least $~2^{-n}$. This is a technical difference from
384: the constraint satisfaction problem in which the cost function
385: generally takes only the set of values that scales polynomially
386: with $n$; the size of the set equals the total number of
387: constraints $m$ (the computationally most difficult case
388: corresponds to $m \sim n$ and the case of $m \sim 2^{n}$ is not of
389: general interest there). To avoid the above restriction in
390: the implementation of the adabatic quantum optimization algorithm for SPP, we
391: introduce a cost function $E(\Omega)$ based on a logarithmic
392: scale of the partition residue values:
393: \begin{eqnarray}
394: &&E(\Omega) = 0 , \quad {\rm for} \quad 0\leq |\Omega|/\Delta < 1,
395: \label{cost2}\\ && E(\Omega) = k , \quad {\rm for}
396: \quad 2^{k-1}\leq |\Omega|/\Delta < 2^{k}, \quad k=0,1,\ldots
397: L,\nonumber\\ &&2^{L-1} \leq A/\Delta < 2^{L}, \quad A=\sum_{j=0}^{n-1}
398: a_{j}.\nonumber
399: \end{eqnarray}
400: \noindent Since the density of states is linear at $\Omega \ll
401: \sqrt{ n}$ number of states $d_k$ per energy level
402: $E(\Omega_{\textsf z})=k$ will grow exponentially with $k$ in that
403: range ($\sim 2^k$). The total number of levels $L$ depends on the
404: value of $\Delta$. Using the density of states (\ref{density1})
405: for $\Omega_{\textsf{z}}\ll \sqrt n$ one can set
406: \begin{equation}
407: \Delta =\sqrt{n} \, 2^{-n} \,K.\label{seed}
408: \end{equation}
409: where $K$ is some fixed number (a few dozen) independent of $n$.
410: The number of ground states of the problem Hamiltonian
411: \begin{equation}
412: H_{P} = \sum_{k=0}^{L-1}E(\Omega_{z})|z\rangle \langle z|
413: \label{Hp}
414: \end{equation}
415: \noindent approximately equals $K$, and the total number of
416: energy levels $L$ is close to $n$. It can be estimated from (\ref{cost2}) that
417: $n-L \sim log_{2}{K}$. The distribution of the
418: low-lying states $d_k$ with cost function (\ref{cost2}) is
419: somewhat similar to that in the slightly underconstrained cases of
420: the satisfiability problem.
421:
422:
423:
424: \section{Results}
425: To study the complexity of the adiabatic quantum optimization algorithm
426: for SPP, we numerically integrate the time dependent
427: Schr{\" o}dinger equation with the Hamiltonian $H(t)$
428: (\ref{farhi}), (\ref{driver}) in which we set $B_{i}=1$ and
429: $J_{ij}=0$. We start from the symmetric initial state (\ref{ini})
430: and integrate the Schr{\" o}dinger equation in the
431: interval $0 \leq t \leq T$. Unlike the approach adopted in
432: \cite{Farhiapp} we do not set {\em a-priori} a value of success
433: probability. Instead we introduce a complexity metric for the algorithm
434: \begin{equation}
435: {\cal C}(T) = {T+1\over p_0(T)} d_0.\label{complexity}
436: \end{equation}
437: \noindent Here $p_0(T)$ is the total probability of finding the system in
438: its ground level (with $E(\Omega_{\textsf{z}})=0$) at the end of the algorithm, $t=T$,
439: and $d_0$ is the number of states at the ground level. The
440: algorithm has to be repeated on average $d_0/p_0(T)$ number of
441: times to reach success probability $\approx$ 1.
442: A typical plot of
443: ${\cal C}(T)$ for an instance of SPP with $n$=15 numbers is shown in
444: Fig. \ref{fig:complexity15}. At very small $T$ the wavefunction is
445: close to the symmetric initial state and the complexity is $\sim 2^{n}$.
446: The extremely sharp decrease in ${\cal C}(T)$ with $T$ is due to the
447: buildup of the population $p_0(T)$ in the ground level as quantum
448: evolution approaches adiabatic limit. At certain $t=T^{*}$ the function
449: ${\cal C}(T)$ goes through the minimum: for $T > T^{*}$ the
450: decrease in the number of trials $d_0/p_0(T)$ does not compensate
451: anymore for the overall increase in the runtime $T$ for each trial.
452: The minimal complexity $C^*=C(T^*)$ is
453: defined via one dimensional minimization over $T$ for a given
454: problem instance \cite{adnote}.
455: \begin{figure}
456: \centerline{\psfig{figure=Complexity_N1501.eps,width=4in,height=3in}}
457: \caption{$C(T)$ {\em vs} $T$ for n=15, precision b=25 bits. Point 1 on the figure corresponds
458: to the minimal value of complexity;
459: the corresponding values are $T^*=22.67$, $p_0(T^*)=0.15$ and $d_0=22$. At Point 2 the
460: total population of the ground level has already reached $p_0(T)=70\%$.
461: }
462: \label{fig:complexity15}
463: \end{figure}
464: \noindent
465: In Fig.\ref{fig:complexity} we plotted the data for optimal complexities $C^*$ at different
466: values of $n$ on a logarithmic scale. Vertical sets of points
467: on the plot indicate the results for {\em all} simulation data we currently have
468: for each $n$. The results indicate that the
469: median value of complexity $C^*$ scales {\em exponentially} with
470: $n$; linear fit to the graph gives $\log C^* \approx 0.56\, n$. This
471: corresponds to the scaling law $C^* \sim 2^{0.8\, n}$. The
472: exponential behavior of the algorithm clearly manifests itself for
473: the larger values of $n\geq$ 11. The scatter in the values of
474: $\log C^*$ appears to decreases with $n$ however this result is probably due to the
475: smaller number of data points available for larger
476: $n$ values.
477: \begin{figure}
478: \centerline{\psfig{figure=complexity06.eps,width=6in}}
479: \caption{$C^{*}$ {\em vs} $n$, precision b=25 bits. Percentage figures correspond to the
480: total population at the ground level at minimal complexity for a given n
481: (e.g. 25$\%$ for n=9). Numbers below the vertical sets of points for each $n$ show
482: the number of trials (e.g., 97 for n=8, 6 for n=16, etc). Numbers
483: on the top indicate the average values of $d_0$ for all trials at a given $n$.
484: Median values for the complexity for each $n$ are shown with red squares.
485: The line is the least squares fit of an exponential function to the median values
486: between n=11 and n=17.
487: }
488: \label{fig:complexity}
489: \end{figure}
490:
491:
492: In Fig. \ref{fig:probabilities} we show the distribution of the
493: probabilities $|\langle z |\psi(T)|^2$ for different values of $T$
494: for an instance of SPP with $n$=15 (plots for different $T$ shown with different
495: colors)and precision b=25 bits. Values of $z$ are ordered with
496: respect to the corresponding values of the partition residues
497: $|\Omega_{\textsf z}|$. It is clearly seen on logarithmic scale
498: that probability distribution forms 'steps' corresponding to
499: different values of the cost function $E(\Omega_{\textsf z})$ defined
500: in (\ref{cost2}). Within each step, the distribution of probabilities
501: does not reveal any structure. The same property holds also for
502: intermediate times ($t < T$). Detailed analytical results \cite{VUD}
503: indicate that it is this absence of structure in
504: $\psi_z(t)$ that is responsible for the exponential
505: complexity of the algorithm.
506: \begin{figure}
507: \centerline{\psfig{figure=Probability_Distribution01.eps,width=4in,height=3in}}
508: \caption{$|\langle z|\psi(T)|^2$ {\em vs } $z$ for one instance of SPP with n=15.
509: Note that the values of the index number on the horizontal axis correspond to the positions of
510: different bit-strings
511: ${\textsf{z}}$
512: sorted with respect to the
513: partition residue values
514: $|\Omega_{\textsf{z}}|$
515: (in increasing order). Index number 0 corresponds to the smallest partition
516: residue.
517: The number of states at the ground
518: level is $d_0$=22. Curve shown in red corresponds to the value of $T=T^*=32$ (minimal complexity).
519: Curve shown in green corresponds to the value of $T$=8 and black color curve corresponds to $T$=90.
520: %Total probabilities at the ground level $p(T)$ equal 0.5 for $T$=90; 0.1 for $T=T^*=$=32;
521: %0.05 for $T=8$
522: }
523: \label{fig:probabilities}
524: \end{figure}
525: \noindent
526: \begin{figure}
527: \centerline{\psfig{figure=eigen_10_13.eps, width=4in,height=3in}}
528: \caption{Adiabatic eigenvalues $g_k$ {\em vs} $s=t/T$ (k=0,1,2,$\ldots$).}
529: \label{fig:Eigenvalues1}
530: \end{figure}
531: \noindent
532: \begin{figure}
533: \centerline{\psfig{figure=eigen_10_13a.eps,width=4in,height=3in}}
534: \caption{Magnified version of the Fig. \ref{fig:Eigenvalues1} in the avoided -crossing region.}
535: \label{fig:Eigenvalues2}
536: \end{figure}
537: \noindent
538: \subsection*{The Stationary Schr{\"o}dinger equation}
539: In addition to solving the time-dependent Schr{\" o}dinger equation
540: we also analyzed the adiabatic solutions of the stationary Schr{\"
541: o}dinger equation with the same form of the Hamiltonian $H(t)$
542: (\ref{farhi}) as above. Our preliminary results were obtained using
543: Mathematica for modest values of $n \leq$ 10. The results for n=10
544: are shown in Figs. \ref{fig:Eigenvalues1} and
545: \ref{fig:Eigenvalues2}. Figure \ref{fig:Eigenvalues2} represents the magnified
546: part of Fig. \ref{fig:Eigenvalues1} near the avoided crossing region.
547: %e used the same form of the Hamiltonian
548: Adiabatic eigenvalues were computed for different
549: values of the scaled time parameter $s=t/T \in (0,1)$. The solid line
550: represents the evolution of the ground state eigenvalue between $s=0$ and $s=1$.
551: The vertical sets of points
552: correspond to excited adiabatic levels for a given $s$. At the
553: beginning ($s=0$) eigenvalues correspond to those of the
554: Hamiltonian $V$: equally spaced levels -n,-n+2, $\ldots$, n,
555: corresponding to different number of spin excitations along the $x$
556: quantization axis. The first excited state is $n-$fold degenerate,
557: the second is $n(n-1)$, the k-th exited state is $\binom{n}{k}$-fold
558: degenerate, etc. For $s>0$ the degeneracy is removed. For
559: $s\rightarrow 1$ the eigenvalue spectrum is the one for the problem
560: Hamiltonian $H_{P}$ (in \ref{Hp}). We have shifted the energy reference in the Hamiltonian
561: $H_P$ by $-n$ (cf. also (\ref{farhi})) to match the energy scale for the symmetric case
562: which emphasizes the avoided crossing region.
563: In our case the ground state was 13-fold degenerate and
564: the corresponding eigenvalues merge at $s\rightarrow 1$. The close
565: approach of these eigenvalues is not relevant for the minimum-gap
566: analysis since they all end up in the same final level. However
567: the minimum separation of the instantaneous adiabatic ground state
568: eigenvalue from the excited state eigenvalues that {\em do not} end up
569: on the same ground level at $s=1$ is clearly seen in the figures.
570: Note that the size of this separation is much greater than the
571: separations between the excited states. This behavior clearly
572: departs from the standard 2-level avoided crossing picture and is
573: due to the contributions from the exponential number of terms in
574: (\ref{nonad}) as will be analyzed elsewhere \cite{VUD}.
575: We also note that the value n=10 does not correspond to the
576: exponential scaling regime for the algorithmic complexity that appears to
577: start for greater $n$ values as follows from the discussion above.
578:
579: In conclusion, we have performed numerical simulations of the
580: adiabatic quantum optimization for SPP using a step-like density
581: of states defined on a logarithmic scale of partition residues.
582: The results indicate an exponential scaling of the algorithmic
583: complexity as a function of the problem size. The apparent reason
584: is the loss of structure in SPP during the effective
585: coarse-graining over the intervals of partition residues
586: corresponding to the same cost function values.
587:
588:
589:
590: %
591: %\begin{figure}
592: %\centerline{\psfig{figure=T_distribution01.eps,width=4in,height=3in}}
593: %\caption{}
594: %\label{fig:Tdistribution}
595: %\end{figure}
596:
597: \begin{thebibliography}{99}
598: \bibitem{Garey}M.R. Garey and D.S. Johnson, {\em Computers and Intractability. A Guide to the Theory
599: of NP-Completeness} (W.H. Freeman, New York, 1997)
600:
601: \bibitem{Farhi} (\emph{a}) E. Farhi, J. Goldstone, S. Gutmann, and M. Sipser, ``Quantum computation by
602: adiabatic evolution,'' arXiv:quant-ph/0001106; (\emph{b})
603: E. Farhi, J. Goldstone, S. Gutmann, J. Lapan, A. Lundgren, and D. Preda, ``A quantum adiabatic evolution
604: algorithm applied to random instances of an NP-complete problem'', {\it Science} {\bf 292}, 472 (2001).
605:
606: \bibitem{Farhiapp} E. Farhi, J. Goldstone, and S. Gutmann,
607: ``A numerical study of the performance of a quantum adiabatic evolution algorithm for satisfiability,''
608: arXiv:quant-ph/0007071; A. M. Childs, E. Farhi, J. Goldstone, and
609: S. Gutmann, ``Finding cliques by quantum adiabatic evolution'', arXiv:quant-ph/0012104.
610:
611: \bibitem{Cerf} J. Roland and N. Cerf, "Quantum Search by local adiabatic evolution",
612: arXiv:quant-ph/0107015
613:
614: \bibitem{Grover} L. K. Grover, ``Quantum mechanics helps in searching for a needle in a haystack,''
615: Physical Review Letters {\bf 79}, 325--328 (1997).
616:
617: \bibitem{Landau} L.D. Landau and E.M. Lifschitz , "Quantum
618: Mechanics", Pergamon, London (1959).
619:
620: \bibitem{Mertens} S. Mertens, (\emph{a}) ``Phase transition in the number partitioning problem,''
621: Physical Review Letters {\bf 81}, 4281--4284 (1998); (\emph{b})
622: ``Random costs in combinatorial optimization,'' {\em Physical
623: Review Letters} {\bf 84}, 1347--1350 (2000).
624:
625: \bibitem{Fu} Y. Fu, in {\em Lectures in the Sciences of Complexity}, ed. by D.L. Stein (Addison-Wesley Publishing Company, Reading, Massachusetts, 1989).
626:
627: \bibitem{Ferreira} F. Ferreira and J. Fontanari, {\em Journal of Physics A} {\bf 31}, 3417 (1998).
628:
629: \bibitem{note2}
630: We also note that occasionally for certain "singular"
631: instances of the partition problem values of approximate g.c.d.
632: can be rather large (e.g. $~1/n$), i.e. most of the numbers are
633: nearly commensurate with each other. In this case intermediate
634: resonances will be of interest and the density of states will have
635: an additional structure at low 'frequencies' $\delta w ~ /pi/d$.
636: We do not consider those instances in a present paper.
637:
638: \bibitem{Walsh} I.P.Gent and T. Walsh, {\em Comp. Intell.} {\bf
639: 14}, 430 (Blackwell, Cambridge MA, 1998).
640:
641: \bibitem{Korf} R.E. Korf, {\em Artif. Intell.} {\bf 106}, 181
642: (1998).
643:
644: \bibitem{Raedt} H. De Raedt et al,{\em Phys. Lett. A} {\bf 290},5-6,
645: p. 227-233 (2001).
646:
647: \bibitem{adnote} We note that an additional optimization can be
648: done if one defines a complexity $C$ similar to (\ref{complexity})
649: but uses an intermediate time instance $t$ instead of $T$ there.
650: In this case quantum algorithm is terminated at
651: the instance $t=t^*(T)\leq T$ for each $T$ when minimal complexity
652: is reached and then additional minimization over $T$
653: is performed. Our results indicate that this method does
654: provide further improvement for the overall complexity but takes a
655: prohibitely long time to perform numerical optimization of $C$.
656:
657: \bibitem{VUD} V.N. Smelyanskiy, U.V. Toussaint, D.A. Timucin, to
658: be submitted.
659: \end{thebibliography}
660: \end{document}
661: