1: \documentclass[pre,final,twocolumn,superscriptaddress,showpacs,footinbib,floatfix]
2: {revtex4}
3: %\documentclass[pra,draft,groupedaddress,showpacs,nofootinbib]{revtex4}
4: \usepackage{amsmath}
5: %\usepackage{graphics}
6: \usepackage{epsfig}
7: \newcommand{\ket}[1]{\left| #1 \right\rangle}
8: \newcommand{\bra}[1]{\left\langle #1 \right|}
9: \newcommand{\iprod}[2]{\left\langle #1|#2\right\rangle}
10: \newcommand{\Eq}[1]{Eq.~(\ref{#1})}
11: \newcommand{\sinc}{\mathrm{sinc}}
12: \newcommand{\tr}{\mathrm{Tr}}
13:
14: \begin{document}
15: \bibliographystyle{revtex}
16:
17: \title{Efficiency of Free Energy Calculations of Spin Lattices by Spectral
18: Quantum Algorithms}
19: \author{Cyrus P. Master}
20: \email{cpmaster@stanford.edu}
21: \affiliation{Quantum Entanglement Project, ICORP, JST,
22: Stanford University, Stanford, CA 94305-4085}
23: \author{Fumiko Yamaguchi}
24: \affiliation{Quantum Entanglement Project, ICORP, JST,
25: Stanford University, Stanford, CA 94305-4085}
26: \author{Yoshihisa Yamamoto}
27: \affiliation{Quantum Entanglement Project, ICORP, JST,
28: Stanford University, Stanford, CA 94305-4085}
29: \affiliation{NTT Basic Research Laboratories,
30: 3-1 Morinosato-Wakamiya, Atsugi, Kanagawa 243-0198, Japan}
31: \date{\today}
32: \begin{abstract}
33: Quantum algorithms are well-suited to calculate estimates of the
34: energy spectra for spin lattice systems. These algorithms are
35: based on the efficient calculation of discrete Fourier components
36: of the density of states. The efficiency of these algorithms in
37: calculating the free energy per spin of general spin lattices to
38: bounded error is examined. We find that the number of Fourier
39: components required to bound the error in the free energy due to
40: the broadening of the density of states scales polynomially with the
41: number of spins in the lattice. However, the precision with which
42: the Fourier components must be calculated is found to be an
43: exponential function of the system size.
44: \end{abstract}
45: \pacs{03.67-a, 05.30.-d, 05.50.+q}
46: \maketitle
47:
48: \section{Introduction}
49:
50: Spin lattice models are useful for the study of magnetic ordering
51: in real materials. The dynamics of these models are specified by
52: a Hamiltonian $\hat{\mathcal{H}}$ involving spin operators for
53: each of the $n$ lattice sites. Of particular interest is the
54: behavior of thermodynamic functions -- such as the magnetization,
55: specific heat capacity, and magnetic susceptibility -- across
56: phase transitions. These functions are encapsulated in the
57: dependence of the Helmholtz free energy per spin, $F$, on system
58: parameters such as the temperature or applied magnetic field;
59: partial derivatives of $F$ yield the thermodynamic functions.
60: Thus, calculation of $F$ over a wide parameter space suffices for
61: the determination of the finite temperature behavior of the spin
62: lattice model.
63:
64: Calculation of the free energy for a general spin lattice by
65: conventional means is difficult. A naive approach is to enumerate
66: the eigenenergies $\{E_m\}$ of $\hat{\mathcal{H}}$, since
67: \begin{equation}
68: F = -n^{-1}k_B\theta\ln Z = -n^{-1}k_B\theta\ln
69: \left(\sum_m e^{-\beta E_m}\right),
70: \label{Eq:I.1}
71: \end{equation}
72: where $k_B\theta = \beta^{-1}$ is the thermal energy, and $Z$ is
73: the partition function. However, as the number of eigenstates
74: grows exponentially with the number of spins in the lattice, the
75: time required to perform the calculation is prohibitively large. A
76: variety of quantum Monte Carlo methods exist to calculate the free
77: energy, including thermodynamic integration
78: \cite{Frenkel1,deKoning1}, histogram methods
79: \cite{Ferrenberg1,Alves1} and cumulant expansion
80: \cite{Rickman1,Phillpot1} techniques. However, the ``sign
81: problem'' (see, for example, Ref. \onlinecite{Landau1}) prevents
82: application of these methods to arbitrary lattice Hamiltonians.
83:
84: %As the state space of a quantum computer grows exponentially with
85: %its physical size, there have been several inquiries as to whether
86: %a quantum computer may be more successful at simulating a quantum
87: %system than conventional techniques
88: %\cite{Manin1,Feynman1,Wiesner1,Boghosian1,Zalka1}. However,
89: %whereas the mapping of the unitary dynamics of test systems onto
90: %the state space of a quantum computer has been investigated in
91: %detail \cite{Abrams1,Ortiz1,Somma1}, these approaches are not
92: %readily applicable to simulating systems at finite temperature.
93:
94: An alternate approach is available if one can efficiently generate
95: an estimate of the density of states $\rho(E)$. As Eq.
96: (\ref{Eq:I.1}) may be written in the form
97: \begin{equation}
98: F = -n^{-1}k_B\theta\ln\left(\int_{-\infty}^\infty\rho(E)e^{-\beta
99: E}dE\right),
100: \end{equation}
101: an approximation for the density of states $\rho(E)$ directly
102: translates into an estimate $\tilde{F}$ for the free energy per
103: spin.
104:
105: Algorithms for quantum computers have been proposed to determine
106: information about the spectra of Hermitian operators
107: \cite{Shor2,Kitaev1,Cleve1,Abrams1,DeRaedt1,Somma1}. We focus on
108: algorithms \cite{DeRaedt1,Somma1} that efficiently generate
109: estimates of individual Fourier components $f_\ell$ of $\rho(E)$;
110: they will be reviewed in detail in Section II. $N$ iterations of
111: the algorithms yield $N$ Fourier components, from
112: which an estimate of the density of states can be calculated.
113:
114: An important issue that has not been addressed is the efficiency
115: of these algorithms for calculating thermodynamic functions as a
116: function of $n$. For the calculation to be deemed efficient, it
117: must be shown that the computation time -- and, thus, the number
118: of Fourier components -- required to calculate an estimate $\tilde{F}$
119: to bounded error scales polynomially with $n$. The bounded error criterion
120: we adopt is
121: \begin{equation}
122: \mathrm{Prob}\left(|\tilde{F}-F| <
123: \gamma k_B\theta\right) > 1-\epsilon, \label{Eq:I.3}
124: \end{equation}
125: where $\gamma$ and $\epsilon$ are small constants. Thus, the
126: absolute error in the estimated free energy per spin must be
127: smaller than a fraction of the thermal energy with probability
128: arbitrarily close to one.
129:
130: We examine the primary sources of error involved in the
131: calculation of $F$ to determine the efficiency of the
132: spectral algorithms. First, as only a finite number of Fourier
133: components $f_\ell$ of the density of states are calculated, the
134: estimated density of states is broadened relative to the actual
135: function. This \emph{deterministic} source of error (\emph{i.e.},
136: it is unchanged if the calculation is repeated) is reduced by
137: increasing the number of components $N$, and thus the computation
138: time. Second, there is an inherent \emph{stochastic} source of
139: error reflected in deviations in the estimated $f_\ell$ from the
140: actual values. This error could arise from imprecise
141: implementation of logic gates or noise in the measurement process.
142:
143: In this paper, we will show that if all of the $f_\ell$ are known
144: exactly, the bound in Eq. (\ref{Eq:I.3}) may be met by a number of
145: Fourier components that scales polynomially with $n$. Thus, the
146: error due to the broadening of the density of states does not
147: prevent efficient estimation of the free energy per spin. However,
148: the free energy becomes increasingly sensitive to random errors in
149: each of the $f_\ell$ as the number of spins is increased. We will
150: show that the precision of the output of the quantum algorithm
151: must improve exponentially with $n$ in order to sustain the
152: condition in Eq. (\ref{Eq:I.3}). Thus, it is questionable as to
153: whether spectral algorithms can be applied to the calculation of
154: thermodynamic functions.
155:
156: The paper is organized as follows: Section II reviews the quantum
157: algorithms used to generate the components $f_\ell$, and discusses
158: assumptions and expected properties of the spin Hamiltonian.
159: Section III describes the calculation of $\tilde{F}$ from the
160: Fourier components, and discusses the influence of sampling and
161: window functions. In Section IV, we analyze the deterministic
162: error due to broadening and determine the number of samples
163: required to meet Eq. (\ref{Eq:I.3}). In Section V, we analyze the
164: impact of random deviations in the components $f_\ell$ on the
165: estimated free energy.
166:
167: \section{Quantum Algorithms}
168: In this section, we review quantum algorithms for the calculation
169: of the Fourier transform of the density of states. We
170: describe a simple algorithm applicable only to Hamiltonians that
171: are diagonal in the computational basis, and then discuss a more
172: general algorithm \cite{Knill1} applicable to ensemble quantum computers.
173:
174: In regards to notation, we use the standard model for quantum
175: computation, assuming our $p$ qubits to be two-level systems with
176: logical states $\ket{0}_j$ and $\ket{1}_j$,
177: $j\in\{0,1,\ldots,p-1\}$, corresponding to eigenstates of the
178: $\hat{\sigma}_z^{(j)}$ Pauli spin operator with eigenvalues $\pm
179: 1$. The computational basis states for the quantum computer are
180: denoted as $\ket{x}=\ket{x_1}_1\ket{x_2}_2\ldots\ket{x_p}_p$,
181: where $\{x_j\}$ are the binary digits of the integer $x$. It is
182: assumed that the quantum computer is capable of implementing a
183: universal set of elementary single-qubit and two-qubit gates. The
184: evolution time of these gates is an implementation-dependent
185: constant, such that the overall computation time is reflected by
186: the number of gates used in the algorithm.
187:
188: We will restrict our discussion to lattices of spin-$1/2$ particles, as it
189: is straightforward to map the eigenstates of $\hat{\sigma}_z^{(j)}$ in
190: the spin system to the logical $\ket{0}_j$ and $\ket{1}_j$ states
191: of qubit $j$ of the quantum computer. It should be noted that this restriction
192: does not preclude the treatment of lattices of particles with spins larger
193: than $1/2$. Generalized Jordan-Wigner transformations exist
194: \cite{Batista1,Jordan1} to represent the dynamics of such lattices by
195: a collection of spin-1/2 particles \emph{via} intermediate fermionization.
196:
197: Prior to the discussion of individual algorithms, we state three
198: assumptions regarding the nature of the spin lattice Hamiltonian.
199: First, we assume that the energy bandwidth $\Delta E$ -- the
200: energy difference between the ground state and the highest excited
201: state -- is bounded by a polynomial function of $n$. This
202: assumption is likely to be valid for models of interest. As an
203: example, consider a lattice of particles interacting by an
204: nearest-neighbor XXZ interaction:
205: \begin{equation}
206: \hat{\mathcal{H}} = \sum_{\left<i,j\right>}\left[J_x\left(\hat{\sigma}_x^{(i)}
207: \hat{\sigma}_x^{(j)}+\hat{\sigma}_y^{(i)}
208: \hat{\sigma}_y^{(j)}\right)+J_z\hat{\sigma}_z^{(i)}
209: \hat{\sigma}_z^{(j)}\right]. \label{Eq:II.1}
210: \end{equation}
211: The expectation value of the summand in Eq. (\ref{Eq:II.1}) must
212: lie between $-(2|J_x|+|J_z|)$ and $(2|J_x|+|J_z|)$. The number of
213: terms in the summation is equal to $n/2$ times the coordination
214: number for the lattice, implying that the energy bandwidth is
215: bounded by a function linear in $n$. In general, for any
216: Hamiltonian involving only pairwise interactions \footnote{This
217: assumption is more restrictive if a pair-wise Hamiltonian for a
218: lattice of particles with spin larger than $1/2$ is transformed to
219: an equivalent spin-1/2 lattice, as the transformed Hamiltonian
220: will not necessarily consist of only pair-wise interactions.}
221: between spins (of $n$-independent interaction energy), it is
222: evident that the energy bandwidth is $O(n^2)$.
223:
224: Second, we assume that the time-evolution operator
225: $\hat{U}(t)\equiv\exp(-i\hat{\mathcal{H}}t)$ can be implemented as a
226: sequence of elementary single-qubit and two-qubit gates, where the
227: number of gates is a polynomial function of $n$. In cases where
228: the Hamiltonian consists of commuting pair-wise interactions
229: (\emph{e.g.}, the Ising model), this decomposition is trivial.
230: Otherwise, one may use a Trotter-Suzuki expansion of non-commuting
231: terms \cite{Suzuki1,Lloyd1} to implement $\hat{U}(t)$ to
232: arbitrarily small error.
233:
234: Finally, we assume that the energy scale is shifted such that the
235: eigenenergies fall between $E=0$ and $E=\Delta E$. This last
236: assumption is made for mathematical convenience, and does not
237: affect the results of our analysis.
238:
239: The following algorithms are based on the fact that the Fourier transform
240: of the density of states $\rho(E)$ is equal to the trace of the
241: time evolution operator:
242: \begin{align}
243: f(t) &\equiv
244: \int_{-\infty}^{\infty}\rho(E)e^{-iEt}dE =
245: \mathrm{Tr}\left(e^{-i\hat{\mathcal{H}}t}\right).
246: \end{align}
247: As $|f(t)|\leq 2^n$, it is convenient to define a function
248: \begin{equation}
249: g(t) \equiv \frac{1}{2^n}f(t) = \frac{1}{2^n}\mathrm{Tr}
250: \left(e^{-i\hat{\mathcal{H}}t}\right),
251: \label{eq:def g(t)}
252: \end{equation}
253: such that $|g(t)|\leq 1$. The algorithms described in this section
254: calculate samples of $g(t)$ at discrete times $t_\ell$.
255:
256: Before discussing the general case, it is illuminating to examine
257: a simple algorithm restricted to spin lattices for which $\hat{\mathcal{H}}$
258: is diagonal in the computational basis. As an example, one might consider
259: a nearest-neighbor Ising model in a longitudinal magnetic field:
260: \begin{equation}
261: \hat{\mathcal{H}} =
262: J_z\sum_{\{i,j\}}\left(1-\hat{\sigma}_z^{(i)}\hat{\sigma}_z^{(j)}\right)
263: +h\sum_i\left(1-\hat{\sigma}_z^{(i)}\right). \label{eq: Ising
264: model}
265: \end{equation}
266: The gates shown in Fig. 1 for an $n$-qubit computer can be used to
267: calculate the magnitude of $g({t_\ell})\equiv g_\ell$. The
268: quantum computer is initialized to the $\ket{0}$ state. The gate
269: $W$ corresponds to a sequence of Walsh-Hadamard gates $W_j$ for
270: each qubit $j$:
271: \begin{equation}
272: W_j: \left\{
273: \begin{array}{ll}
274: \ket{0}_j \rightarrow \frac{1}{\sqrt{2}}
275: \left(\ket{0}_j+\ket{1}_j\right) \\
276: \ket{1}_j \rightarrow \frac{1}{\sqrt{2}}
277: \left(\ket{0}_j-\ket{1}_j\right)
278: \end{array}.
279: \right.
280: \end{equation}
281: The gate $U({t_\ell})$ corresponds to the time evolution operator.
282: \begin{figure}
283: \begin{center}
284: \epsfxsize = 3in \epsfbox{block1.eps}
285: %\includegraphics{block1.eps}
286: \caption{Logic diagram of an elementary algorithm to estimate $|g_\ell|$.}
287: \end{center}
288: \label{fig: 1}
289: \end{figure}
290:
291: As a final step, a projective measurement is performed in the computational
292: basis. It is straightforward to show that the probability of observing all
293: qubits in the logical $\ket{0}$ state is equal to $|g_\ell|^2$:
294: \begin{align}
295: \ket{0}&\xrightarrow{W}\frac{1}{\sqrt{2^n}}\sum_{m=0}^{2^n-1}\ket{m}
296: \xrightarrow{U({t_\ell})}\frac{1}{\sqrt{2^n}}\sum_{m=0}^{2^n-1}
297: e^{-iE_m {t_\ell}}\ket{m} \nonumber \\ &\xrightarrow{W}
298: g_\ell\ket{0}+\text{orthogonal components}.
299: \end{align}
300: By assumption, the computational basis states $\ket{m}$ are
301: eigenstates of the Hamiltonian, and the time-evolution operator
302: appends a phase proportional to the eigenvalue $E_m$ to each term.
303: An unbiased estimator for $|g_\ell|$ can be derived by
304: performing $R$ repetitions of the algorithm $R$, and counting the number
305: of times all qubits are found in the logical $\ket{0}$ state.
306:
307: The magnitude of $g_\ell$ is insufficient to reconstruct the density of
308: states. By adding an ancilla qubit $a$, as shown in Fig. 2, one may extract
309: estimates of both the real and imaginary parts of $g_\ell$. The $X
310: \equiv\exp\left(i\pi\hat{\sigma}_x^{(a)}/4\right)$ and $Y \equiv
311: \exp\left(i\pi\hat{\sigma}_y^{(a)}/4\right)$ gates correspond to
312: $\pi/2$ rotations of the ancilla qubit. If the $X$ gate is
313: used, the probabilities of observing the
314: $\ket{\phi_0}\equiv\ket{0}_a\ket{0}_{q_1}\ldots\ket{0}_{q_n}$ or
315: $\ket{\phi_1}\equiv\ket{1}_a\ket{0}_{q_1}\ldots\ket{0}_{q_n}$ states
316: are
317: \begin{align}
318: p_{X0} = \left|\frac{1+ig_\ell}{2}\right|^2, \
319: p_{X1} = \left|\frac{1-ig_\ell}{2}\right|^2,
320: \end{align}
321: respectively. The $Y$ gate leads to probabilities
322: \begin{align}
323: p_{Y0} = \left|\frac{1+g_\ell}{2}\right|^2, \
324: p_{Y1} = \left|\frac{1-g_\ell}{2}\right|^2.
325: \end{align}
326: By executing $R$ iterations of the algorithm with the $X$ gate and $R$
327: iterations with the $Y$ gate, one can derive estimators $\tilde{p}$
328: for the probabilities. Unbiased estimates of the real and imaginary
329: parts of $g_\ell$ are
330: \begin{align}
331: \mathrm{Re}\left(\tilde{g}_\ell\right) &=
332: \left(\tilde{p}_{Y0}-\tilde{p}_{Y1}\right), \\
333: \mathrm{Im}\left(\tilde{g}_\ell\right) &=
334: \left(\tilde{p}_{X1}-\tilde{p}_{X0}\right).
335: \end{align}
336: We use the tilde to distinguish estimates of the Fourier components
337: obtained from the quantum algorithm from the exact values.
338: \begin{figure}
339: \begin{center}
340: %\includegraphics{block2.eps}
341: \epsfxsize = 3in \epsfbox{block2.eps}
342: \caption{Logic diagram of an algorithm to estimate the real and imaginary
343: parts of $g_\ell$ for diagonal $\hat{\mathcal{H}}$.}
344: \end{center}
345: \label{fig: 2}
346: \end{figure}
347:
348: The algorithm described above depends on our ability
349: to construct an equally-weighted coherent superposition of eigenstates of
350: $\hat{\mathcal{H}}$; hence, the restriction to Hamiltonians that are
351: diagonal in the computational basis. It can be generalized
352: \cite{Knill1} by considering an algorithm involving an ensemble of quantum
353: computers, such that the ancilla qubit is still initialized to
354: the $\ket{0}_a$ state, but the remaining $n$ qubits are in a completely
355: random mixed state. The initial density matrix for the system
356: is
357: \begin{equation}
358: \hat{\rho}_i = \frac{1}{2^{n+1}}\left(\hat{I}^{(a)}+{\hat{\sigma}_z^{(a)}}
359: \right)\hat{I}^{(q_1)}\hat{I}^{(q_2)}\ldots\hat{I}^{(q_n)},
360: \end{equation}
361: where $\hat{I}^{(\ell)}$ is the identity operator for qubit $\ell$. The
362: operator $\hat{I}^{(q_1)}\hat{I}^{(q_2)}\ldots\hat{I}^{(q_n)}$
363: is equal to the resolution of the identity
364: $\sum_{\ell}\ket{\psi_\ell}\bra{\psi_\ell}$, where $\{\ket{\psi_\ell}\}$ is
365: an orthogonal set of states in the subspace spanned by qubits $q_1$ through
366: $q_n$. One could use as $\{\ket{\psi_\ell}\}$ the eigenstates of the
367: Hamiltonian $\hat{\mathcal{H}}$. We do not need to explicitly solve for the
368: eigenstates of $\hat{\mathcal{H}}$; the initial density matrix can be thought of
369: as an incoherent mixture of eigenstates for any Hamiltonian $\hat{\mathcal{H}}$.
370:
371: If the coherent superposition created by the Walsh-Hadamard gates
372: is replaced by such an incoherent mixture, then an algorithm
373: nearly identical to the one shown above works for any choice of
374: $\hat{\mathcal{H}}$, as shown in Fig. 3. The final measurement is
375: the expected value of $\hat{\sigma}_z^{(a)}$ averaged over the
376: ensemble.
377: \begin{figure}
378: \label{fig: 3}
379: \begin{center}
380: \epsfxsize = 3in \epsfbox{block3.eps}
381: %\includegraphics{block3.eps}
382: \caption{Logic diagram of an ensemble algorithm to determine $g_\ell$ for
383: arbitrary $\hat{\mathcal{H}}$.}
384: \end{center}
385: \end{figure}
386:
387: If the $X$ gate is used for the ancilla qubit, the expected value of
388: $\hat{\sigma}_z^{(a)}$ is
389: \begin{equation}
390: \left<\hat{\sigma}_z^{(a)}\right> =
391: \mathrm{Im}\left(g_\ell\right),
392: \end{equation}
393: whereas the $Y$ gate leads to
394: \begin{equation}
395: \left<\hat{\sigma}_z^{(a)}\right> =
396: \mathrm{Re}\left(g_\ell\right).
397: \end{equation}
398: Thus, the real and imaginary components of the estimator
399: $\tilde{g}_\ell$ can be calculated from two iterations of the
400: ensemble quantum algorithm.
401:
402: The ensemble algorithm is attractive for two reasons. First, it
403: is applicable to any spin-1/2 lattice Hamiltonian
404: $\hat{\mathcal{H}}$, provided that the time-evolution operator can
405: be decomposed into a sufficiently small number of elementary
406: gates. Second, initial state preparation lends itself to ensemble
407: quantum computation proposals involving spin resonance, where the
408: Z\'{e}eman splitting between qubit spin states is small compared
409: to the thermal energy. In equilibrium, the initial density matrix
410: of the system is well-approximated by the identity operator. The
411: single pseudopure state qubit can be created from two thermal
412: spins \cite{Gershenfeld1}.
413:
414: \section{Free Energy Estimation}
415: In this section, we discuss how an estimate $\tilde{F}$ of the
416: free energy is generated from the Fourier components of the
417: density of states, and we examine the effects of discretization on the
418: estimated density of states.
419:
420: A conceptual overview of the free energy calculation including
421: post-processing is shown in Fig. 4. $N$ samples of $f(t)$ are
422: estimated \emph{via} the quantum algorithm \footnote{As this section
423: exploits the Fourier relationship between $f(t)$ and the density of states,
424: it is more convenient notation-wise to commence with the former than $g(t)$.},
425: and are weighted by a windowing function $b_\Theta(t)$, described below. Fourier
426: transformation yields an estimate $\tilde{\rho}(E)$ for the
427: density of states. The density of states may be integrated to compute
428: the partition function, and, thus, the free energy.
429:
430: As iterations of the quantum algorithm yield discrete samples of
431: $f(t)$, the reconstructed estimate $\tilde{\rho}(E)$ is distorted
432: relative to the exact function $\rho(E)$. This distortion
433: translates into error in $\tilde{F}$. It is convenient to
434: view this error in the context of windowing and sampling of the
435: exact Fourier transform $f(t)$ of the density of states.
436: Truncation of $f(t)$ to a window of width $T_0$ centered about
437: $t=0$ (\emph{i.e.}, multiplication of $f(t)$ by a windowing function
438: $b_1(t)$ that is constant for $|t| \leq T_0/2$ and zero elsewhere)
439: leads to a convolution of the density of states by a broadening
440: function $b_1(E) \equiv \alpha_1\sinc\left(\frac{\pi E}{\Delta
441: e}\right) =\alpha_1\left[\sin\left(\frac{\pi E}{\Delta
442: e}\right)\right] /\left(\frac{\pi E}{\Delta e}\right)$, where the
443: energy resolution $\Delta e$ is given by
444: \begin{equation}
445: \Delta e = \frac{2\pi}{T_0}. \label{E3:energy resolution window
446: size}
447: \end{equation}
448: The window is scaled such that the broadening function is
449: normalized to unit area; \emph{i.e.,} $\alpha_1 = 1/\Delta e$.
450: Increasing the window size $T_0$ reduces the width of the broadening
451: function, and thus the error in the estimate of $\tilde{F}$.
452: \begin{figure}
453: \begin{center}
454: \epsfxsize = 3in \epsfbox{block4.eps}
455: %\includegraphics{block4.eps}
456: \caption{Block diagram of the calculation of $\tilde{F}$.}
457: \end{center}
458: \label{fig: 4}
459: \end{figure}
460:
461: The effect of sampling on the estimated density of states can be
462: determined by multiplying $f(t)b_1(t)$ by an impulse train $s(t)$
463: of spacing $\Delta t$:
464: \begin{equation}
465: s(t) = \Delta t\sum_{\ell=-\infty}^\infty
466: \delta(t-\ell\Delta t).
467: \label{Eq:sdef}
468: \end{equation}
469: Sampling leads to periodic replication of the broadened $\rho(E)$. The
470: resultant density of states is given by the Fourier transform of
471: $f(t)b_1(t)s(t)$:
472: \begin{equation}
473: \rho'(E) \equiv \rho(E) \ast b_1(E) \ast \sum_{k=-\infty}^\infty
474: \delta\left(E+\frac{2\pi k}{\Delta t}\right).
475: \label{eq: estimated DOS as convolution}
476: \end{equation}
477: To avoid aliasing in the estimated density of states, the Nyquist
478: sampling condition requires that
479: \begin{equation}
480: \Delta t \leq \frac{2\pi}{\Delta E}. \label{E3: energy bandwidth
481: sampling time}
482: \end{equation}
483: The spacing between samples of $f(t)$ is determined solely by
484: the estimate of the energy bandwidth, $\Delta E$. We
485: assume that sampling is performed at the Nyquist rate, and the
486: equality holds in Eq. (\ref{E3: energy bandwidth sampling time}).
487:
488: As the number of samples is equal to the ratio of the windowing
489: function width $T_0$ to the sampling time, one could determine the
490: minimum value of $T_0$ required to satisfy Eq. (\ref{Eq:I.3}) as a
491: function of $n$. However, the rectangular windowing function
492: leads to poor results. The envelope of the associated broadening
493: function $b_1(E)$ falls off weakly as $1/E$; the oscillating side
494: lobes are amplified at low energies by the Boltzmann
495: factor in the calculation of the free energy. The window width
496: required to mitigate the resultant error scales poorly with $n$.
497: In contrast to using wider rectangular windows, one may
498: adopt more elaborate window shapes whose corresponding
499: broadening functions exhibit envelopes that are more sharply
500: peaked. We consider the functions $b_\Theta(t)$ formed by the
501: successive convolution of $\Theta$ rectangular windows, each of
502: width $T_0$. $\Theta$ is referred to as the order of the windowing
503: function. For $\Theta=2$, the window is triangular and of width $2
504: T_0$. With increasing order, the window approaches a Gaussian
505: shape, and is of width $\Theta T_0$. The resulting broadening
506: function is then
507: \begin{equation}
508: b_\Theta(E) = \alpha_\Theta \left[\sinc\left(\frac{\pi E}{\Delta
509: e}\right)\right]^\Theta, \label{E3: defn of b_Theta}
510: \end{equation}
511: which exhibits a $1/E^\Theta$ envelope. The value of
512: $\alpha_\Theta$ is determined, as the area under
513: $b_\Theta(E)$ is one. In practice, a given window shape is
514: constructed by obtaining samples $\tilde{f}_\ell$ within the window
515: width $\Theta T_0$ centered at $t=0$, and weighting each sample by
516: $b_{\Theta,\ell}\equiv b_\Theta({t_\ell})$.
517:
518: As the envelope of the side lobes of $b_\Theta(E)$ falls off
519: exponentially with $\Theta$, windowing functions of large order
520: significantly reduce the error in the calculated free energy.
521: However, the tradeoff is a wider window, leading to more Fourier
522: components, and thus more iterations of the quantum algorithm:
523: \begin{equation}
524: N=\frac{\Theta T_0}{\Delta t}.
525: \label{eq: N samples}
526: \end{equation}
527: Therefore, the question of how $N$ scales with the number of spins, $n$,
528: translates into the determination of the minimum values of $\Theta$ and $T_0$
529: required to satisfy Eq. (\ref{Eq:I.3}).
530:
531: An estimate $\tilde{Z}$ for the partition function can be
532: calculated directly from the estimated Fourier components without
533: intermediate calculation of the density of states. We denote
534: quantities obtained from the quantum algorithm with a tilde, in
535: contrast to their exact values. First, note that the Fourier
536: transform of $\tilde{f}(t)b_\Theta(t)s(t)$ may be evaluated explicitly
537: via Eq. (\ref{Eq:sdef}) to give an estimate of the broadened, periodically
538: replicated density of states $\rho'(E)$ in terms of the components
539: $\tilde{f}_\ell$:
540: \begin{align}
541: \tilde{\rho}'(E) &= \frac{1}{2\pi}\int_{-\infty}^\infty \tilde{f}(t)b_\Theta(t)
542: \left[\Delta t \sum_{\ell=-\infty}^\infty
543: \delta(t-\ell\Delta t)\right]e^{iEt}dt \nonumber \\
544: &= \frac{\Delta t}{2\pi}\sum_{\{t_\ell\}} \tilde{f}_\ell b_{\Theta,\ell}
545: e^{iEt_\ell},
546: \label{rho from f samples}
547: \end{align}
548: where we have defined $t_\ell\equiv\ell\Delta t$, and the sum is performed
549: over all $t_\ell$ within the window described by $b_{\Theta,\ell}$.
550: Integrating Eq. (\ref{rho from f samples}) over the
551: energy bandwidth and using Eqs. (\ref{eq:def g(t)}) and
552: (\ref{E3: energy bandwidth sampling time}), one finds
553: \begin{widetext}
554: \begin{align}
555: \tilde{Z} &= \int_0^{\Delta E}\tilde{\rho}'(E)e^{-\beta E} dE
556: %\nonumber\\&
557: = \frac{2^n\Delta t}{2\pi\beta}\left(1-e^{-\beta\Delta E}\right)
558: \left\{b_{\Theta,0}+2\sum_{\ell>0}^{N/2}b_{\Theta,\ell}
559: \left[\frac{1}{1+(t_\ell/\beta)^2}\mathrm{Re}(\tilde{g}_\ell)
560: -\frac{t_\ell/\beta}{1+(t_\ell/\beta)^2}\mathrm{Im}(\tilde{g}_\ell)\right]
561: \right\}.
562: \label{partition function from f samples}
563: \end{align}
564: \end{widetext}
565: Note that an estimate $\tilde{F}$ of the free energy may be obtained from the
566: logarithm of Eq. (\ref{partition function from f samples}).
567:
568: In addition to describing how an estimate of the free energy per spin is
569: calculated from the Fourier components, Eq.
570: (\ref{partition function from f samples}) will serve as a starting point to
571: determine the probabilistic error in $\tilde{F}$ due to imprecise values
572: of $\tilde{g}_\ell$.
573:
574: \section{Error Analysis: Broadening}
575:
576: In this section, we determine an upper bound on the number of
577: samples $N$ of $g(t)$ required to calculate the free energy to the tolerance
578: prescribed by Eq. (\ref{Eq:I.3}). At this point, we consider the individual
579: samples of $g_\ell$ to be known exactly, and only consider the error in
580: $\tilde{F}$ due to the finite number of Fourier components --
581: \emph{i.e.}, due to broadening of the density of states. With this
582: restriction, we can show that $N$ is a polynomial function of the number of
583: spins $n$.
584:
585: As it is more convenient to work with the partition function than the free
586: energy, we use a more stringent bound based upon the relative error
587: in the calculated partition function $\tilde{Z}$. As
588: \begin{equation}
589: \left|\tilde{F}-F\right| < \gamma k_B \theta \Longleftrightarrow
590: e^{-\gamma n}-1 < \frac{\tilde{Z}-Z}{Z} < e^{\gamma n}-1,
591: \end{equation}
592: it is sufficient to demand that
593: \begin{equation}
594: \mathrm{Prob}\left(r\equiv\left|\frac{\tilde{Z}-Z}{Z}\right| < \xi\right)
595: > 1-\epsilon, \label{eq: Z error criterion}
596: \end{equation}
597: where $\xi\equiv 1-\exp(-\gamma n)$. Note that $\xi<1$, and in the limit
598: $\gamma n \ll 1$, $\xi\rightarrow\gamma n$.
599:
600: By Eqs. (\ref{E3:energy resolution window size}), (\ref{E3: energy bandwidth
601: sampling time}) and (\ref{eq: N samples}), if the Nyquist sampling condition
602: is satisfied, then
603: \begin{equation}
604: N = \frac{\Theta\Delta E}{\Delta e}.
605: \label{eq: N}
606: \end{equation}
607: It has been asserted that $\Delta E$ is a polynomial function of $n$.
608: In the remainder of this section, we examine the dependence of
609: $\Theta$ and $\Delta e$ on $n$ such that Eq. (\ref{eq: Z error criterion})
610: is satisfied. We require a pair of intermediate results:
611:
612: \boxed{\emph{Lemma 1:}} If $b_\Theta(E)$ (as defined in Eq.
613: (\ref{E3: defn of b_Theta})) is subject to the normalization
614: condition $1 = \int_{-\infty}^\infty b_\Theta(E)dE$, then
615: \begin{equation}
616: \alpha_\Theta < \frac{c\pi}{\Delta e}\sqrt{\frac{\Theta}{6\pi}},
617: \label{E3: Lemma 1}
618: \end{equation}
619: where $c \approx 2.0367$.
620:
621: \boxed{\emph{Lemma 2:}}
622: \begin{equation}
623: A_{\text{side}} \equiv 1-\int_{-\Delta e}^{\Delta e} b_\Theta(E)dE <
624: \frac{c}{\pi^{\Theta-3}} \sqrt{\frac{\Theta}{6\pi}}, \label{E3:
625: Lemma 2}
626: \end{equation}
627: where $\Theta$ is an even integer.
628:
629: Lemma 1 places an upper bound on $\alpha_\Theta$ such
630: that $b_\Theta(E)$ is normalized. Lemma 2 states an upper bound on
631: the area of $b_\Theta(E)$ that is outside of the
632: interval $[-\Delta e,\Delta e]$ (\emph{i.e.},
633: outside the main lobe of the broadening function) that decreases
634: exponentially with $\Theta$. Both are proved in the Appendix. Note that
635: the proof of Lemma 2 applies to the case where $\Theta$ is even.
636:
637: One can relate the relative error $r$ in the
638: calculated partition function to the parameters $\Theta$ and
639: $\Delta e$ \emph{via} Lemma 2. As the exact density of states $\rho(E)$ may
640: be expressed as a sum of delta functions for each eigenenergy $E_m$,
641: Eqs. (\ref{eq: estimated DOS as convolution})-(\ref{E3: energy bandwidth
642: sampling time}) at the Nyquist condition yield
643: \begin{align}
644: \tilde{Z} &= \int_0^{\Delta E}\tilde{\rho}(E)
645: e^{-\beta E}dE \nonumber \\
646: &= \sum_m \sum_{k=-\infty}^\infty \int_0^{\Delta E}
647: b_\Theta(E-E_m+k\Delta E)e^{-\beta E}dE \nonumber \\
648: &= \sum_m \left[\sum_{k=-\infty}^\infty \int_{k\Delta E}^{(k+1)\Delta E}
649: b_\Theta(E-E_m)e^{-\beta(E-k\Delta E)}dE\right] \nonumber \\
650: &\equiv \sum_m \tilde{Z}_m.
651: \end{align}
652: The change of variables allows one to view $\tilde{Z}_m$ as an integral of
653: the broadening function, centered at $E_m$, and weighted by
654: periodically-replicated segments of an exponential function. $\tilde{Z}$
655: is found by summing over all eigenenergies.
656:
657: The maximum relative error $r$ in the partition function is bounded by the
658: largest contribution $r_m \equiv \max_m \left|\frac{\tilde{Z}_m-Z_m}{Z_m}
659: \right|$ from any single eigenenergy, where $Z_m \equiv e^{-\beta E_m}.$
660: Define $\gamma_m = \tilde{Z}_m/Z_m$. Then,
661: \begin{align}
662: r &= \left|\frac{\tilde{Z}-Z}{Z}\right| = \frac{\left|\sum_m
663: \left(\tilde{Z}_m-Z_m\right)\right|}{\sum_m Z_m} =
664: \frac{\left|\sum_m(\gamma_m-1)Z_m\right|}{\sum_m Z_m}
665: \nonumber \\
666: &\leq \max_m \left|\gamma_m - 1\right| = \max_m
667: \left|\frac{\tilde{Z}_m-Z_m}{Z_m}\right|=r_m.
668: \end{align}
669:
670: This argument shows that one may consider a simplified system with just one
671: eigenstate at an energy ${E_m}$ somewhere in the energy bandwidth. An upper
672: bound on the error $r$ for this simplified system for any ${E_m}$ suffices to
673: bounded the error for an arbitrary energy spectrum over the same
674: bandwidth.
675:
676: Lower and upper bounds on $\tilde{Z}_m$ ($\tilde{Z}_{m,\text{min}}$
677: and $\tilde{Z}_{m,\text{max}}$, respectively) are now derived to bound
678: $r_m$, since
679: \begin{equation}
680: r_m < \max\left(\left|\frac{\tilde{Z}_{m,\text{min}}-Z_m}{Z_m}\right|,
681: \left|\frac{\tilde{Z}_{m,\text{max}}-Z_m}{Z_m}\right|\right). \label{E3:
682: ri}
683: \end{equation}
684: In the main lobe, the minimum value of the Boltzmann factor
685: is $e^{-\beta({E_m}+\Delta e)}$. Outside
686: of the main lobe, the minimum value is $e^{-\beta\Delta E}$.
687: Thus,
688: \begin{align}
689: \tilde{Z}_m &= \sum_{k=-\infty}^\infty \int_{k\Delta E}^{(k+1)\Delta E}
690: b_\Theta(E-E_m)e^{-\beta(E-k\Delta E)}dE \nonumber \\
691: &\geq (1-A_{\text{side}})e^{-\beta({E_m}+\Delta e)}+A_{\text{side}}
692: e^{-\beta\Delta E} \equiv \tilde{Z}_{m,\text{min}}. \label{E3: Zimin}
693: \end{align}
694: Similarly, as the maximum value of the Boltzmann factor is
695: $e^{-\beta({E_m}-\Delta e)}$ inside the main lobe and one
696: outside,
697: \begin{align}
698: \tilde{Z}_m \leq (1-A_{\text{side}})e^{-\beta({E_m}-\Delta
699: e)}+A_{\text{side}} \equiv \tilde{Z}_{m,\text{max}}. \label{E3: Zimax}
700: \end{align}
701: Substituting Eqs. (\ref{E3: Zimin}) and (\ref{E3: Zimax}) into Eq.
702: (\ref{E3: ri}), we see that
703: \begin{align}
704: r_m < \max[&1-(1-A_{\text{side}})e^{-\beta\Delta e}
705: -A_{\text{side}}e^{-\beta(\Delta E-{E_m})}, \nonumber \\
706: &(1-A_{\text{side}})e^{\beta\Delta e}+
707: A_{\text{side}}e^{\beta{E_m}}-1]. \label{E3: ri2}
708: \end{align}
709: It is difficult to invert Eq. (\ref{E3: ri2}) explicitly to find
710: optimal conditions on $A_{\text{side}}(\Theta)$ and $\Delta e$ that ensure
711: that $r_m < \xi$. However, one can show that the following
712: conditions are sufficient:
713: \begin{align}
714: \beta\Delta e &= \ln(1+\xi/2), \label{E3: Delta e condition}\\
715: A_{\text{side}} &< \frac{\xi}{2}e^{-\beta\Delta E}. \label{E3: Aside
716: condition}
717: \end{align}
718: As proof of their sufficiency, note that
719: \begin{align}
720: &1-(1-A_{\text{side}})e^{-\beta\Delta e}-A_{\text{side}}e^{-\beta(\Delta
721: E-{E_m})} \nonumber \\
722: &< \frac{\xi}{2}+\frac{\xi}{2}e^{-\beta\Delta E}\left(1-\frac{\xi}{2}\right)
723: \nonumber \\
724: &< \xi,
725: \end{align}
726: and,
727: \begin{align}
728: &(1-A_{\text{side}})e^{\beta\Delta e}+A_{\text{side}}e^{\beta{E_m}}-1
729: \nonumber \\
730: &< \frac{\xi}{2}+\frac{\xi}{2}e^{-\beta(\Delta E-E_m)}
731: \nonumber\\
732: &< \xi.
733: \end{align}
734: Therefore, the conditions in Eqs. (\ref{E3: Delta e
735: condition}) and (\ref{E3: Aside condition}) guarantee that $r < r_m <
736: \xi$, as desired.
737:
738: Using Lemma 2, one can manipulate Eqs. (\ref{E3: Delta e
739: condition}) and (\ref{E3: Aside condition}) to show that $N$ scales
740: polynomially with $n$.
741: \begin{align}
742: \Delta e &= \frac{\ln(1+\xi/2)}{\beta}, \label{E3: final condition 1}\\
743: \Theta -
744: \frac{\ln\Theta}{2\ln\pi}&> \frac{\beta\Delta E}{\ln \pi}+
745: \frac{\ln(1/\xi)}{\ln\pi}+\kappa, \label{E3: final condition 2a}
746: \end{align}
747: where $\kappa = 5/2+\frac{\ln(2c/\sqrt{6})}{\ln\pi}
748: \approx 2.9443$ . As $\ln\Theta < \Theta$, a sufficient condition to
749: satisfy Eq. (\ref{E3: final condition 2a}) is
750: \begin{equation}
751: \Theta/2 = \lceil\mu\beta\Delta E + \mu\ln(1/\xi) + \kappa'\rceil,
752: \label{eq: final condition 2b}
753: \end{equation}
754: where $\mu \equiv 1/(2\ln\pi-1)$ and $\kappa' \equiv \mu\kappa\ln\pi$.
755:
756: In summary, the error bound on the partition function is satisfied if the
757: energy resolution scales linearly with temperature, and if $\Theta$
758: scales linearly with $\beta\Delta E$.
759:
760: As a final step, we substitute the conditions in Eqs.
761: (\ref{E3: final condition 1}) and (\ref{eq: final condition 2b}) into
762: Eq. (\ref{eq: N}), disregarding the weak logarithmic dependence of $\Theta$
763: and $\Delta e$ on $n$:
764: \begin{equation}
765: N = \frac{\Theta\Delta E}{\Delta e} \propto \frac{(\beta\Delta E)
766: (\Delta E)}{1/\beta} = \beta^2(\Delta E)^2\propto
767: \mathrm{poly}(n),
768: \end{equation}
769: by the assertion that the energy bandwidth is a polynomial
770: function of the number of spins in our system.
771: This result shows that in the absence of error in the calculated Fourier
772: components of the density of states, the free energy per spin can be
773: determined efficiently to bounded error.
774:
775: \section{Error Analysis: Fourier Components}
776:
777: We now consider random errors in the individual values of
778: $\tilde{g}_\ell$, which may arise from imprecise implementation of
779: logic gates, or noise in the measurement process. Treating
780: $\mathrm{Re}(\tilde{g}_\ell)$ and $\mathrm{Im}(\tilde{g}_\ell)$ as
781: random variables, these fluctuations are modelled by their
782: variances. We assume that the variances $\sigma^2_g$ are
783: independent \footnote{The assumption is
784: reasonable if we consider the error for the ensemble quantum
785: algorithm to arise from noise in the measurement process, as
786: discussed below.} of $\ell$. In this section, the dependence
787: of the maximum allowable value of $\sigma^2_g$ on $n$ such that
788: Eq. (\ref{Eq:I.3}) is maintained is derived.
789:
790: As the estimate for the partition function $\tilde{Z}$ is a linear
791: combination of the independent random variables
792: $\mathrm{Re}(\tilde{g}_\ell)$ and $\mathrm{Im}(\tilde{g}_\ell)$,
793: the variance of $\tilde{Z}$ can be calculated from Eq.
794: (\ref{partition function from f samples}):
795: \begin{equation}
796: \sigma_{\tilde{Z}}^2 = 4^{n+1}\left(\frac{\Delta
797: t}{2\pi\beta}\right)^2(1-e^{-\beta\Delta E})^2
798: \sigma_g^2\sum_{\ell>0}^{N/2}\frac{b_{\Theta,\ell}^2}{1+(t_\ell/\beta)^2}.
799: \label{Eq:V.1}
800: \end{equation}
801: If we assume that $\tilde{Z}$ is Gaussian-distributed \footnote{
802: The validity of this assumption is dependent on the nature of the
803: noise source. It is exact if the probability distribution
804: functions for each of the $\mathrm{Re}(\tilde{g_\ell})$ and
805: $\mathrm{Im}(\tilde{g_\ell})$ are themselves Gaussian.}, then the
806: probability of $\tilde{Z}$ deviating from its exact value $Z$ can
807: be related to the variance. Thus, the sum in Eq. (\ref{Eq:V.1}) is
808: evaluated by making two simplifications. First, we model the
809: windowing function $b_\Theta(t)$ as a Gaussian. Recall that $b_\Theta(t)$
810: is constructed by the convolution of $\Theta$ rectangular windows of
811: width $T_0$. In the
812: limit of large $\Theta$, $b_\Theta(t)$ may be approximated
813: \footnote{As the Fourier transform of $b_\Theta(t)$ is the broadening
814: function $b_\Theta(E)$, which is of unit area, $b_\Theta(t=0)=1$.} by
815: \begin{equation}
816: b_\Theta(t) \approx e^{-t^2/2\nu^2},
817: \end{equation}
818: where $\nu^2 = \Theta T_0^2/12$. Although this approximation
819: overestimates $b_\Theta(t)$ away from $t=0$, the fractional error in Eq.
820: (\ref{Eq:V.1}) incurred by the approximation is less than $5\times
821: 10^{-3}$ for $\Theta>40$. Second, it is assumed that $\beta/\Delta
822: t = \beta\Delta E/2\pi \gg 1$. The energy bandwidth is thus much
823: larger than the thermal energy. This condition assures that the
824: sum can be well-approximated by the integral
825: \begin{align}
826: \sigma_{\tilde{Z}}^2&\approx 4^{n+1}\left(\frac{\Delta
827: t}{2\pi\beta}\right)^2\sigma_g^2
828: \int_0^\infty\frac{e^{-t^2/\nu^2}}{1+t^2/\beta^2}
829: \frac{dt}{\Delta t} \nonumber \\
830: &=\frac{4^n\sigma_g^2}{\beta\Delta E}
831: e^{\beta^2/\nu^2}[1-\mathrm{erf}(\beta/\nu)]. \label{eq:V.2}
832: \end{align}
833:
834: Eq. (\ref{eq:V.2}) indicates that the standard deviation of
835: $\tilde{Z}$ scales exponentially \footnote{Although the term
836: $e^{\beta^2/\nu^2}[1-\mathrm{erf}(\beta/\nu)]$ is a weakly
837: decreasing function of $n$, the $4^n$ dependence dominates.} with
838: $n$; \emph{i.e.}, as $2^n$. Note that the exact partition function
839: $Z$ will typically be a more slowly increasing function of $n$. If
840: the energy eigenvalues are limited to the domain $[0,\Delta E]$,
841: then $2^n$ is an upper bound for the value of the partition
842: function (achieved at infinite temperature, or if all eigenstates
843: are degenerate with zero energy). Consider two simple examples. For
844: the case of $n$ non-interacting spins in a magnetic field with
845: Zeeman energy $h$, $Z=(1+e^{-\beta h})^n<2^n$; for a linear chain Ising
846: model in zero magnetic field, described by Eq. (\ref{eq: Ising model}),
847: $Z=(1+e^{-2\beta J})^n$ for
848: periodic boundary conditions. Thus, if the distribution function
849: for $\tilde{Z}$ is Gaussian, one expects that the standard deviation increases
850: exponentially faster \footnote{The mean of $\tilde{Z}$ is not strictly $Z$
851: due to the broadening error, but may be bounded to an arbitrarily
852: small region about $Z$ by the techniques of the previous
853: section.} than the mean $Z$.
854:
855: The above result may be used to derive a condition on $\sigma_g^
856: 2$ such that the error bound on the free energy per spin is fulfilled.
857: By Eq. (\ref{Eq:I.1}), the condition $|\tilde{F}-F|<\gamma k_B\Theta$ in
858: Eq. (\ref{Eq:I.3}) is equivalent to
859: \begin{equation}
860: Ze^{-\gamma n} < \tilde{Z} < Z e^{\gamma n}.
861: \end{equation}
862: Assuming a Gaussian distribution for $\tilde{Z}$ centered about
863: $Z$,
864: \begin{align}
865: \epsilon&=1-\mathrm{Prob}(Ze^{-\gamma n} < \tilde{Z} < Z e^{\gamma n}) \\
866: &=\frac{1}{2}\left\{
867: \mathrm{erfc}\left[\frac{Z(e^{\gamma n}-1)}{\sqrt{2}\sigma_{\tilde{Z}}}\right]
868: +\mathrm{erfc}\left[\frac{Z(1-e^{-\gamma n})}{\sqrt{2}\sigma_{\tilde{Z}}}\right]
869: \right\}. \nonumber
870: \end{align}
871: This result can be simplified if we consider the limit $\gamma n \ll 1$, such
872: that $e^{\pm\gamma n}\approx 1\pm\gamma n$; \emph{i.e.}, for small desired
873: absolute error in the free energy relative to the number of spins:
874: \begin{equation}
875: \epsilon = \mathrm{erfc}\left(\frac{Z\gamma n}{\sqrt{2}\sigma_{\tilde{Z}}}\right)
876: \approx \mathrm{erfc}\left(\sqrt{\frac{\beta\Delta E}{2}}
877: \frac{Z\gamma n}{2^n\sigma_g}\right).
878: \end{equation}
879: The argument of the $\mathrm{erfc}(\cdot)$ function must be order unity or
880: larger for $\epsilon < 0.1$, so
881: \begin{equation}
882: \sigma_g^2 = O\left(\frac{Z^2\mathrm{poly}(n)}{4^n}\right).
883: \label{eq:V.3}
884: \end{equation}
885: By the above argument, the variance in the measured Fourier components must
886: decrease exponentially with $n$.
887:
888: The condition on $\sigma_g^2$ is likely to translate into an exponentially
889: scaling computation time for the overall calculation. For example, consider as
890: a quantum computer an ensemble of spin-1/2 nuclei, where readout is performed by
891: measuring the voltage induced in a pickup coil by free induction. A source of
892: error in the measured Fourier components is the Johnson-Nyquist voltage noise
893: due to the resistance of the coil \cite{Hoult1}. The variance in the observed
894: voltage -- and thus in the estimates for $\mathrm{Re}(g_\ell)$ and
895: $\mathrm{Im}(g_\ell)$ -- is inversely proportional to the measurement time.
896: Thus, Eq. (\ref{eq:V.3}) implies that an exponentially long measurement time is
897: required to satisfy the condition in Eq. (\ref{Eq:I.3}).
898:
899: \section{Conclusion}
900:
901: We examined the applicability of spectral quantum algorithms for
902: the calculation of the free energy of spin lattice models.
903: Provided that the time-evolution operator for the system is
904: decomposable into an efficient number of elementary gates, an
905: ensemble quantum algorithm exists to generate estimates of the
906: density of states by calculating individual Fourier components of
907: $\rho(E)$. We analyzed the efficiency of this algorithm in
908: calculating the free energy per spin of the system to bounded
909: absolute error.
910:
911: The error in the calculated free energy arises from the
912: calculation of only a discrete number of Fourier components
913: $f_\ell$, as well as from deviations in the measured values of
914: $f_\ell$ due to statistical errors. The first source of error,
915: attributable to broadening in the estimated density of states, was
916: shown to lead to bounded error with a number of Fourier components
917: that is polynomial in $n$. Thus, if the components $f_\ell$ are
918: known exactly, the spectral algorithm is an efficient means to
919: calculate the free energy per spin. However, the effect of random
920: deviations in the calculated values of $f_\ell$ grows with
921: increasing $n$. As the size of the system increases, the maximum
922: tolerable variance in measured Fourier components decreases as
923: $Z^2/4^n$ for large $n$ and small absolute error. As an upper
924: bound for the partition function is $2^n$, the spectral algorithms
925: are not an efficient method to determine $F$ in the presence of
926: statistical errors in $f_\ell$.
927:
928: \begin{acknowledgments}
929: This work is partially supported by the DARPA QuIST program. CPM
930: acknowledges the support of the PACCAR Inc. Stanford Graduate Fellowship.
931: \end{acknowledgments}
932: %\bibliography{bibdata}
933:
934: \begin{thebibliography}{10}
935: \providecommand*{\bibinfo}[2]{#2}
936: \providecommand*{\eprint}[1]{#1}
937: \providecommand*{\url}[1]{#1}
938: \bibitem{Frenkel1}
939: \bibinfo{author}{D.~Frenkel} and \bibinfo{author}{A.~Ladd},
940: \bibinfo{journal}{J. Chem. Phys} \bibinfo{volume}{\textbf{81}},
941: \bibinfo{pages}{3188} (\bibinfo{date}{1984}).
942: \bibitem{deKoning1}
943: \bibinfo{author}{M.~de~Koning}, \bibinfo{author}{A.~Antonelli}, and
944: \bibinfo{author}{S.~Yip}, \bibinfo{journal}{Phys. Rev. Lett.}
945: \bibinfo{volume}{\textbf{83}}, \bibinfo{pages}{3973} (\bibinfo{date}{1999}).
946: \bibitem{Ferrenberg1}
947: \bibinfo{author}{A.~M. Ferrenberg} and \bibinfo{author}{R.~H. Swendsen},
948: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{61}},
949: \bibinfo{pages}{2635} (\bibinfo{date}{1988}).
950: \bibitem{Alves1}
951: \bibinfo{author}{N.~A. Alves}, \bibinfo{author}{B.~A. Berg}, and
952: \bibinfo{author}{R.~Villanova}, \bibinfo{journal}{Phys. Rev. B}
953: \bibinfo{volume}{\textbf{41}}, \bibinfo{pages}{383} (\bibinfo{date}{1990}).
954: \bibitem{Rickman1}
955: \bibinfo{author}{J.~Rickman} and \bibinfo{author}{S.~Phillpot},
956: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{66}},
957: \bibinfo{pages}{349} (\bibinfo{date}{1991}).
958: \bibitem{Phillpot1}
959: \bibinfo{author}{S.~Phillpot} and \bibinfo{author}{J.~Rickman},
960: \bibinfo{journal}{J. Chem. Phys.} \bibinfo{volume}{\textbf{94}},
961: \bibinfo{pages}{1454} (\bibinfo{date}{1991}).
962: \bibitem{Landau1}
963: \bibinfo{author}{D.~P. Landau} and \bibinfo{author}{K.~Binder},
964: \bibinfo{title}{\emph{A Guide to Monte Carlo Simulations in Statistical
965: Physics}} (\bibinfo{publisher}{Cambridge University Press}, U.K.,
966: \bibinfo{year}{2000}).
967: \bibitem{Shor2}
968: \bibinfo{author}{P.~Shor} (\bibinfo{date}{1995}), e-print, quantum-ph/9508027.
969: \bibitem{Kitaev1}
970: \bibinfo{author}{A.~Y. Kitaev} (\bibinfo{date}{1995}), e-print,
971: quantum-ph/9511026.
972: \bibitem{Cleve1}
973: \bibinfo{author}{R.~Cleve}, \bibinfo{author}{A.~Ekert},
974: \bibinfo{author}{C.~Macchiavello}, and \bibinfo{author}{M.~Mosca},
975: \bibinfo{journal}{Proc. R. Soc. London A} \bibinfo{volume}{\textbf{454}},
976: \bibinfo{pages}{339} (\bibinfo{date}{1998}).
977: \bibitem{Abrams1}
978: \bibinfo{author}{D.~S. Abrams} and \bibinfo{author}{S.~Lloyd},
979: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{79}},
980: \bibinfo{pages}{2586} (\bibinfo{date}{1997}).
981: \bibitem{DeRaedt1}
982: \bibinfo{author}{H.~{De~Raedt}}, \bibinfo{author}{A.~Hams},
983: \bibinfo{author}{K.~Michielsen}, \bibinfo{author}{S.~Miyashita}, and
984: \bibinfo{author}{K.~Saito}, \bibinfo{journal}{Prog. Theor. Phys. Supplement}
985: \bibinfo{volume}{\textbf{138}}, \bibinfo{pages}{489} (\bibinfo{date}{2000}).
986: \bibitem{Somma1}
987: \bibinfo{author}{R.~Somma}, \bibinfo{author}{G.~Ortiz},
988: \bibinfo{author}{J.~Gubernatis}, \bibinfo{author}{E.~Knill}, and
989: \bibinfo{author}{R.~Laflamme}, \bibinfo{journal}{Phys. Rev. A}
990: \bibinfo{volume}{\textbf{65}}, \bibinfo{pages}{042323/1}
991: (\bibinfo{date}{2002}).
992: \bibitem{Knill1}
993: \bibinfo{author}{E.~Knill} and \bibinfo{author}{R.~Laflamme},
994: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{81}},
995: \bibinfo{pages}{5672} (\bibinfo{date}{1998}).
996: \bibitem{Batista1}
997: \bibinfo{author}{C.~Batista} and \bibinfo{author}{G.~Ortiz},
998: \bibinfo{journal}{Phys. Rev. Lett.} \bibinfo{volume}{\textbf{86}},
999: \bibinfo{pages}{1082} (\bibinfo{date}{2001}).
1000: \bibitem{Jordan1}
1001: \bibinfo{author}{P.~Jordan} and \bibinfo{author}{E.~Wigner},
1002: \bibinfo{journal}{Z. Phys.} \bibinfo{volume}{\textbf{47}},
1003: \bibinfo{pages}{631} (\bibinfo{date}{1928}).
1004: \bibitem{Suzuki1}
1005: \bibinfo{author}{M.~Suzuki}, in \bibinfo{editors}{M.~Suzuki}, ed.,
1006: \emph{Quantum Monte Carlo Methods in Condensed Matter Physics}
1007: (\bibinfo{publisher}{World Scientific}, Singapore, \bibinfo{year}{1993}),
1008: \bibinfo{pages}{pp. 13--48}.
1009: \bibitem{Lloyd1}
1010: \bibinfo{author}{S.~Lloyd}, \bibinfo{journal}{Science}
1011: \bibinfo{volume}{\textbf{273}}, \bibinfo{pages}{1073} (\bibinfo{date}{1996}).
1012: \bibitem{Gershenfeld1}
1013: \bibinfo{author}{N.~Gershenfeld} and \bibinfo{author}{I.~Chuang},
1014: \bibinfo{journal}{Science} \bibinfo{volume}{\textbf{275}},
1015: \bibinfo{pages}{350} (\bibinfo{date}{1997}).
1016: \bibitem{Hoult1}
1017: \bibinfo{author}{D.~Hoult} and \bibinfo{author}{R.~Richards},
1018: \bibinfo{journal}{J. Magn. Resonance} \bibinfo{volume}{\textbf{24}},
1019: \bibinfo{pages}{71} (\bibinfo{date}{1976}).
1020: \bibitem{Jeffrey1}
1021: \bibinfo{author}{A.~Jeffrey}, \bibinfo{title}{\emph{Handbook of Mathematical
1022: Formulas and Integrals}} (\bibinfo{publisher}{Academic Press}, U.K.,
1023: \bibinfo{year}{1995}).
1024:
1025: \end{thebibliography}
1026:
1027:
1028: \appendix
1029: \section{Proofs of Lemmas 1 and 2}
1030: \emph{Proof of Lemma 1:} A lower bound is first derived for
1031: \begin{equation}
1032: I \equiv \int_{-\infty}^{\infty} \left[\sinc(x)\right]^\Theta dx =
1033: \int_{-\infty}^{\infty} e^{\Theta\ln\left[\sinc(x)\right]}dx.
1034: \end{equation}
1035: We exclude infinitesimal regions around $x=m\pi$ ($m \in \mathcal{Z}$)
1036: from the integral to avoid divergence of the logarithm; as $\mathrm{sinc}(x)$
1037: approaches a finite value in these regions, the contribution of these
1038: regions to the integral can be made arbitrarily small.
1039:
1040: Using a series expansion for $\ln\left[\sinc(x)\right]$
1041: \cite{Jeffrey1},
1042: \begin{align}
1043: \ln\left[\sinc(x)\right] &= -\frac{x^2}{6} - \sum_{k=2}^\infty
1044: \frac{x^{2k}}
1045: {k\pi^{2k}}\left(\sum_{n=1}^\infty\frac{1}{n^{2k}}\right) \nonumber \\
1046: &> -\frac{x^2}{6} - \left(\frac{\pi^2}{6}\right)\sum_{k=2}^\infty
1047: \frac{x^{2k}} {k\pi^{2k}}.
1048: \end{align}
1049: Thus,
1050: \begin{align}
1051: I > \int_{-\infty}^{\infty} e^{-\Theta x^2/6}
1052: e^{-\frac{\Theta\pi^2}{6}\sum_{k=2}^\infty
1053: \frac{x^{2k}}{k\pi^{2k}}}dx.
1054: \end{align}
1055: The integrand is positive over the entire domain of $x$,
1056: and both exponential factors monotonically decrease with
1057: $|x|$. Thus, one may place a lower bound on $I$ by reducing the
1058: limits of integration to any finite interval, such as
1059: $|x|<\sqrt{6/\Theta}$. Thus,
1060: \begin{align}
1061: I > e^{-\frac{\Theta\pi^2}{6}\sum_{k=2}^\infty
1062: \frac{(6/\Theta)^{k}}{k\pi^{2k}}}
1063: \int_{-\sqrt{6/\Theta}}^{\sqrt{6/\Theta}}e^{-\Theta x^2/6}dx.
1064: \end{align}
1065: The integral is $\sqrt{6\pi/\Theta}\ \mathrm{erf}(1)$. The
1066: summation can be performed explicitly to yield
1067: \begin{align}
1068: I &>
1069: e^{1+\pi^2\Theta\ln(1-6/\pi^2\Theta)/6}\sqrt{\frac{6\pi}{\Theta}}\mathrm{erf}(1)
1070: \nonumber \\
1071: &=
1072: e\left(1-\frac{6}{\Theta\pi^2}\right)^{\Theta\pi^2/6}\mathrm{erf}(1)
1073: \sqrt{\frac{6\pi}{\Theta}}
1074: \nonumber\\&>e\left(1-\frac{6}{\pi^2}\right)^{\pi^2/6}\mathrm{erf}(1)
1075: \sqrt{\frac{6\pi}{\Theta}}.
1076: \end{align}
1077: where we make use of the fact that $(1-1/x)^x$ is a monotonically
1078: increasing function for $x>1$.
1079:
1080: This lower bound for $I$ is used to establish an upper bound
1081: for $\alpha_\Theta$.
1082: \begin{align}
1083: \alpha_\Theta &= \frac{1}{\int_{-\infty}^{\infty}\left[\sinc
1084: \left(\frac{\pi E}{\Delta e}\right)\right]^\Theta dE} \nonumber \\
1085: &= \frac{\pi}{\Delta e \ I} \nonumber \\
1086: &< \frac{\pi}{\Delta e} \left(c \sqrt{\frac{\Theta}{6\pi}}\right),
1087: \end{align}
1088: where $c$ is defined as
1089: \begin{equation}
1090: c \equiv \frac{1}{e}\left(\frac{1}{1-6/\pi^2}\right)^{\pi^2/6}
1091: \frac{1}{\mathrm{erf}(1)} \approx 2.0367.
1092: \end{equation}
1093:
1094: \emph{Proof of Lemma 2:}
1095: %[NOTE: The proof I've laid out here is by
1096: %analogy to probability distributions, which I think is intuitive
1097: %to engineering folk, but probably confusing to mathematicians. I
1098: %have a more formal mathematical proof which I could substitute for
1099: %this one.]
1100: For $\Theta$ even, $b_\Theta(E)$ is a non-negative function with
1101: unit area. If one treats $b_\Theta(E)$ as a probability density
1102: function, one can use the Markov inequality to bound the area
1103: outside of the main lobe.
1104:
1105: Consider a random variable $Y$ with support $y \geq 0$;
1106: \emph{i.e.,} $Y$ only takes non-negative values. Markov's
1107: inequality bounds the probability of deviations from the mean:
1108: \begin{equation}
1109: \mathrm{Pr}\left(Y\geq\delta\right) \leq \frac{E(Y)}{\delta},
1110: \end{equation}
1111: where $E(Y)$ is the expectation value of $Y$. Define a second
1112: random variable $X$, such that $Y = [X-E(X)]^m$, where $m$ is an
1113: even integer. Then,
1114: \begin{align}
1115: \mathrm{Pr}\left\{[X-E(X)]^m \geq \delta\right\} &\leq \frac
1116: {E\left\{[X-E(X)]^m\right\}}{\delta}\nonumber\\
1117: \Rightarrow\mathrm{Pr}\left\{|X-E(X)| \geq \epsilon\right\} &\leq
1118: \frac{E\left\{[X-E(X)]^m\right\}}{\epsilon^m} \label{A: Chebyshev
1119: inequality}
1120: \end{align}
1121: This bound is expressed in terms of the $m^{\mathrm{th}}$ central
1122: moment of $X$, if it exists. The result reduces to Chebyshev's
1123: inequality for $m=2$.
1124:
1125: Note that if one treats $b_\Theta(E)$ as a probability distribution
1126: function for a zero-mean random variable $E$, the above inequality
1127: provides a bound for the area outside the main lobe (\emph{i.e.,}
1128: $\epsilon = \Delta e$). The central moment is evaluated for $m =
1129: \Theta-2$.
1130: \begin{align}
1131: E\left\{[X-E(X)]^m\right\} &= \int_{-\infty}^{\infty} E^{\Theta-2}
1132: b_\Theta(E)dE \nonumber \\
1133: &= \alpha_\Theta\left(\frac{\Delta e}{\pi}\right)^{\Theta-1}
1134: \int_{-\infty}^{\infty} \frac{\sin^\Theta x}{x^2}dx \nonumber \\
1135: &\leq \alpha_\Theta\left(\frac{\Delta e}{\pi}\right)^{\Theta-1}
1136: \int_{-\infty}^{\infty} \frac{\sin^2 x}{x^2}dx \nonumber \\
1137: &= \alpha_\Theta\left(\frac{\Delta e}{\pi}\right)^{\Theta-1}\pi
1138: \end{align}
1139: If we define the area outside the main lobe as
1140: \begin{equation}
1141: A_{\text{side}} \equiv 1-\int_{-\Delta e}^{\Delta e}b_\Theta(E)dE,
1142: \end{equation}
1143: then,
1144: \begin{equation}
1145: A_{\text{side}} = \mathrm{Pr}\left\{|X-E(X)| \geq \Delta e\right\} \leq
1146: \frac{\alpha_\Theta\Delta e}{\pi^{\Theta-2}}. \
1147: \label{A: lemma 2 almost done}
1148: \end{equation}
1149: Combining Eq. [\ref{A: lemma 2 almost done}] with Lemma 1, we find
1150: \begin{equation}
1151: A_{\text{side}} < \frac{c}{\pi^{\Theta-3}} \sqrt{\frac{\Theta}{6\pi}}.
1152: \end{equation}
1153:
1154: \end{document}
1155: