You are on page 1of 4

week ending

PRL 106, 170501 (2011) PHYSICAL REVIEW LETTERS 29 APRIL 2011

Quantum Simulation of Time-Dependent Hamiltonians


and the Convenient Illusion of Hilbert Space
David Poulin,1 Angie Qarry,2,3 Rolando Somma,4 and Frank Verstraete2
1
Département de Physique, Université de Sherbrooke, Sherbrooke, Québec, Canada
2
Faculty of Physics, University of Vienna, Boltzmanngasse 5, 1090 Vienna, Austria
3
Centre for Quantum Technologies, National University of Singapore, Singapore 117543, Singapore
4
Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA
(Received 20 February 2011; published 29 April 2011)
We consider the manifold of all quantum many-body states that can be generated by arbitrary time-
dependent local Hamiltonians in a time that scales polynomially in the system size, and show that it occupies
an exponentially small volume in Hilbert space. This implies that the overwhelming majority of states in
Hilbert space are not physical as they can only be produced after an exponentially long time. We establish
this fact by making use of a time-dependent generalization of the Suzuki-Trotter expansion, followed by a
well-known counting argument. This also demonstrates that a computational model based on arbitrarily
rapidly changing Hamiltonians is no more powerful than the standard quantum circuit model.

DOI: 10.1103/PhysRevLett.106.170501 PACS numbers: 03.67.Ac, 03.65.Ud, 89.70.Eg

The Hilbert space of a quantum system is big—its Hamiltonians obeying constraints (1) and (2), and would
dimension grows exponentially with the number of reach the same conclusions.
particles it contains. Thus, parametrizing a generic The second constraint is very much reminiscent of the
quantum state of N particles requires an exponential way complexity classes are defined in theoretical computer
number of real parameters. Fortunately, the states of science, where the central object of study is the scaling of
many physical systems of interest appear to occupy a the time required to solve a problem as a function of its
tiny submanifold of this gigantic space. Indeed, the input size. The classical analogue for the problem that we
essential physical features of many systems can be address is a well-known counting argument of Shannon [4]
explained by variational states specified with a small demonstrating that the number of Boolean functions of N
bits scales doubly exponentially (as 22 ), with the conse-
N
number of parameters. Well-known examples include
the BCS state for superconductivity [1], Laughlin’s state quence that no efficient (i.e., polynomial) algorithm can
for fractional quantum Hall liquids [2], and tensor net- exist to compute the overwhelming majority of those
work states occurring in real-space renormalization functions. Indeed, the number of different functions that
methods [3]. In these cases, the number of parameters can be encoded by all classical circuits of polynomial
scales only polynomially with N. depth scale as 2polyðNÞ , which is exponentially smaller
In this Letter, we attempt to define the class of physical than the total number of Boolean functions.
states of a many-body quantum system with local Hilbert Our contribution is a quantum generalization of this
spaces of bounded dimensions and prove that they repre- result. The crux of our argument is to demonstrate that
sent an exponentially small submanifold of the Hilbert the dynamics generated by any local Hamiltonian, without
space. We say that a state is physical if it can be reached, any assumptions on its time dependence, can be simulated
starting in some fiducial state (e.g., a ferromagnetic state, by a quantum circuit of polynomial size. All previously
or the vacuum), by an evolution generated by any time- known simulation methods [5–10] produced a quantum
dependent quantum many-body Hamiltonian, with the con- circuit of complexity that depends on the smoothness of
straint that (1) the Hamiltonian is local in the sense that it is the Hamiltonian, scaling, e.g., with k@H=@tk or some
the sum of terms each acting on at most k bodies for some higher derivatives. Using the results of Huyghebaert and
constant k independent of N, and (2) the duration of the De Raedt [11], we show how these conditions can be
evolution scales at most as a polynomial in the number of overcome. We then use a well-known counting argument
particles in the system. The duration of the evolution has a [12] for quantum circuits that involves the Solovay-Kitaev
meaning only if there is a well-defined time or energy theorem [13] to arrive at the conclusion that most states
scale. We do this by setting the strength of the local terms in the Hilbert space are not physical: they can only be
in the Hamiltonian to some constant E. Such bounded reached after an exponentially long time. Note that a direct
interaction strength is crucial to our derivation. The as- parameter counting would not produce this result because
sumption about the initial fiducial state is artificial; we we impose no restriction on the time dependence of
could alternatively define the class of physical evolutions the Hamiltonian. The complete description of a rapidly
for quantum many-body systems as the ones generated by changing Hamiltonian requires lots of information, so

0031-9007=11=106(17)=170501(4) 170501-1 Ó 2011 American Physical Society


week ending
PRL 106, 170501 (2011) PHYSICAL REVIEW LETTERS 29 APRIL 2011

from this perspective there are in principle enough parame- intervals, thus increasing the overall complexity of the
ters to reach all states in the Hilbert space. simulation.
Our demonstration that arbitrary local time-dependent Time-dependent Trotter-Suzuki expansion.—Somewhat
Hamiltonians can be efficiently simulated on quantum surprisingly, it is possible to generalize the Trotter-
computers is of interest in its own right in the context of Suzuki formula to time-dependent Hamiltonians without
quantum computation. More precisely, we are concerned compromising the error, where the Hamiltonian may ex-
with Hamiltonians acting on N particles of the form hibit fluctuations much faster than the time step t. We
X begin by breaking the total time evolution into short seg-
HðtÞ ¼ HX ðtÞ; (1) ments Uð0;tÞ ¼ Uðtn ;tn þ tÞ...Uðt2 ;t2 þ tÞUð0;0 þ tÞ,
Xf1;2;...;Ng
each of duration t
where X labels subsets of the N particles, each term has  Z t þt X 
j
bounded normkHX ðtÞk  E, and each term acts on no Uðtj ; tj þ tÞ ¼ T exp i ds HX ðsÞ :
more than k particles, i.e., HX ðtÞ ¼ 0 if jXj > k, and k tj X
is fixed, independent of the system size. We make no In the simple case where the sum over X contains only two
assumption on the geometry of the system, and the cou- terms, say H1 and H2 , it has been shown [11] that the
pling can be arbitrarily long ranged. The time-evolution generalized Trotter-Suzuki expansion
operator Uð0; tÞ from time 0 to t is governed by  Z t þt 
Schrödinger’s equation dtd Uð0; tÞ ¼ iHðtÞUð0; tÞ, with UTS ðtj ; tj þ tÞ ¼ T exp i
j
dsH1 ðsÞ
solutionRgiven in terms of a time-ordered integral Uð0; tÞ ¼ tj
T expf t0 HðsÞdsg.  Z t þt 
j
Starting with Feynman’s exploration of quantum com-  T exp i dsH2 ðsÞ
tj
puters [14], it has been well established that the time-
evolution operator generated by Hamiltonians of the form gives an error in terms of the operator norm that is
Eq. (1) can be decomposed into short quantum circuits,
provided that HðtÞ varies slowly enough [5–10]. In all kUðtj ; tj þ tÞ  UTS ðtj ; tj þ tÞk  c12 ðtÞ2 ;
cases, this is achieved by approximating the evolution
with c12 of the order of 1 and given by
operator by a product formula
1 Z tj þt Zv
Y
NP c12 ¼ dv duk½H1 ðuÞ; H2 ðvÞk:
Uð0; tÞ  expfiHXp ðtp Þtp g; (2) ðtÞ2 tj tj
p¼1
c12 is upper bounded by c2max =2 with cmax ¼ maxX kHX k,
where the sequences Xp , tp , and tp are set by specific with kHX k ¼ sup0st kHX ðsÞk. Note that this bound does
approximation schemes, such as the Trotter formula [15] or not depend on the derivative of the Hamiltonian (and is
the Lie-Suzuki-Trotter formula [16]. Because each term therefore also valid for nonanalytic time dependence).
HX acts on at most k particles, this last expression repre- Note also that the bound reduces to the usual Trotter error
sents a sequence of k-body unitaries. A standard quantum for the time-independent case and is therefore equally
circuit is obtained by simulating each of these k-body strong, and that it can straightforwardly be generalized to
operators as a sequence of one- and two-qubit gates using higher order decompositions.
the result of Solovay-Kitaev [13,17]. For our present application, the Hamiltonian is the sum
Perhaps the simplest example of a product formula of L 2 polyðNÞ k-particle terms, cf. Eq. (1). We can there-
decomposition of Uð0; tÞ is given by fore iterate the above procedure log2 ðLÞ times; at the nth
Y
n Y iteration, there are 2n terms, each of strength upper
Uð0; tÞ  expfiHX ðjtÞtg; (3) bounded by cmax L=2n . The total error for approximating
j¼1 X the exact time evolution of the Hamiltonian with L terms
by a product of L time-ordered terms is
where the product over X can be carried in any given order.
This decomposition makes use of two approximations.
1 2 X2 L  L 2 1
log
cmax ðtÞ 2 2m m  c2max L2 ðtÞ2 ;
First, the time dependence of the HðtÞ is ignored on time
scales lower than t: the Hamiltonian is approximated by a 2 m¼1 2 2
piecewise constant function taking the values HðjtÞ on
the time interval ½ðj  1Þt; jt. Second, each matrix which can be made arbitrary small by choosing a t that
exponential is decomposed using the Trotter formula scales as an inverse polynomial in N. Approximating the
Q time-evolution operator over a total time t with a product of
expfiHðtÞtg  X expfiHX ðtÞtg. Clearly, the size
t of the time intervals must be shorter than the fluctuation k-body unitaries, such that the total error is =2, can there-
time scale of HðtÞ for the first approximation to be valid, fore be achieved by choosing t ¼ tc2  L2 . The total num-
max

t  k@H=@tk1 . Higher frequency fluctuations would ber G of k-body unitaries to achieve this accuracy is then
2
therefore require breaking the time evolution into shorter equal to Gð; L; tÞ ¼ L tt ¼ cmax 2 3
 t L .
170501-2
week ending
PRL 106, 170501 (2011) PHYSICAL REVIEW LETTERS 29 APRIL 2011

The value of each such k-body unitary can be obtained The second step to obtain a product formula for
by solving the corresponding finite time-ordered integral Uð0; tÞ—one that does not require any integrals—makes
av
on a classical computer. The Solovay-Kitaev (SK) use of randomness. The average Hamiltonian HX;j on
theorem [13,17] shows that each of these k-body unitary the interval ½ðj  1Þt; jt can be estimated using
transformations can be simulated with standard one- Monte Carlo integration. For every j, we can pick m
av
and two-qubit gates chosen from a fixed discrete set (e.g., random times kj 2 ½tj ; tj þ t and approximate HX;j 
CNOT’s between any pair of qubits supplemented by a local 1 Pm
m k¼1 H X ð k
j Þ. Because the variance of the Hamiltonian is
=8 rotation gate). To achieve an accuracy G per unitary bounded by kHX k2 , the sum converges to HX;j av
with error
transformation, we need dSK ½logc2SK ð1=G Þ standard gates pffiffiffiffi
estimate tkHX k= m. Using this Monte Carlo average,
with cSK and dSK constants; we choose G such that G ¼

we can approximate the evolution operator of the time
2Gð;L;tÞ . The total number of quantum gates as chosen from interval ½tj ; tj þ t by
a discrete set of gates needed to approximate the complete  
time evolution with an error  is therefore upper bounded by av 1 Xm
UX ðtj ; tj þ tÞ  exp i H ð Þk
(5)
Gtot ð; L; tÞ ¼ dSK Gð; L; tÞlogc2SK ½Gð; L; tÞ=, which is m k¼1 X j
polynomial in the number of qubits; it roughly scales quad-
ratically in t and as the cube of the total number of (non-
Y  
m
t
commuting) local terms in the Hamiltonian.  exp i HX ðkj Þ ; (6)
Average Hamiltonians and randomized evolution.— k¼1
m
Note that the time-dependent Trotter-Suzuki decom-
position described in the previous section does not where the order of the product can be chosen according to
increasing values of kj . The error in the first approximation
lead to a product formula because each term appearing pffiffiffiffi
in it involves a time-ordered integral UX ðtj ; tj þ tÞ ¼ Eq. (5) is set by the Monte Carlo estimate tkHX k= m
Rt þt while the second approximation Eq. (6) is the usual
T expfi tjj HX ðsÞdsg rather than the exponential of Trotter-Suzuki formula. Summarizing, we can decompose
a term of the Hamiltonian at a given time expfitHX ðtÞg the total evolution operator from time 0 to t as Uð0; tÞ 
as in Eq. (2). Although this does not affect the conclusions Q t
j;k;X expfi m HX ðj Þg, where the product should be
k
reached in the next section on the counting of possible
taken in increasing order of kj and any order of X. This
quantum states, it is unsatisfactory from the point of quan-
tum simulation. In this section, we demonstrate how to is a standard product formula like Eq. (2)—identical to the
recover a product formula by making use of randomness. usual decomposition explained in the introduction [Eq. (3)]
In Ref. [18], product formula decompositions Eq. (2) were and the one presented in [18]—except that the times kj at
found for any local Hamiltonian, where the number of which the Hamiltonian is sampled are random. Thus, we
terms NP in the product depends on a smoothness parame- see that by sampling the Hamiltonian at random times, we
P completely circumvent any smoothness requirements [20].
ter P ¼ sup0pP;0st X ðk@ps HX ðsÞkÞ1=ðpþ1Þ . In par-
We note that the proof of Eq. (4) is an illustration
ticular, these decompositions are inefficient when the
of the decoupling principle that tells us that the high-
fluctuation time scale of the Hamiltonian becomes too
frequency fluctuations of the Hamiltonian should not
small. Our methods circumvent these requirements by
affect the low-energy physics. As a consequence, it is
using randomness. Further, randomization avoids the com-
possible to largely ignore these fluctuations—by replacing
plexity of integration.
the time-dependent Hamiltonian by its average value on
We do this in two steps. First, we replace the time-
each time bin—without significantly modifying the dy-
ordered exponential integral with the exponential of an
namics of the system. This is the working principle behind
ordinary integral without introducing a significant error.
renormalization group methods of quantum field theory
Indeed, we show in Appendix A in the supplemental
and quantum many-body physics. The rotating wave ap-
material [19] that
  Z t þt   Z t þt  proximation [21] and effective Hamiltonian theory [22] are
 j j  simple examples illustrating this principle in the case of
T exp i tj dsHX ðsÞ  exp i dsHX ðsÞ 
tj time-dependent Hamiltonians.
2 More generally, we show in Appendix B of the supple-
 kHX k2 t2 : (4) mental material [19] that we can replace the time-
3
dependent Hamiltonian HðtÞ with a smoothed version
Using this result, we obtain the approximate decomposition ~ with fluctuation time scale bounded by  without
HðtÞ
n Y
Y  Z t þt  significantly affecting the resulting time-evolution opera-
j
Uð0; tÞ  exp i HX ðsÞds : tor. More precisely, we show that the time-evolution op-
tj
j¼1 X
erators from time 0 to t differ by at most kHk2 t.
Note that this is still not a product formula because it Counting states.—Let us now consider the set of all
involves integrals. This first step has nevertheless elimi- quantum states that can be reached starting from some
nated the need of a time-order operator. fiducial state j0i and evolving for some polynomial amount
170501-3
week ending
PRL 106, 170501 (2011) PHYSICAL REVIEW LETTERS 29 APRIL 2011

of time under any time-dependent Hamiltonian. A direct circuits. This raises the question of whether it makes sense
counting argument appears difficult because there are to describe many-body quantum systems as vectors in a
infinitely many distinct time-dependent Hamiltonians, linear Hilbert space. The recent advances in real-space
and therefore there can a priori be infinitely many states renormalization group methods [3,23,24] indeed seem to
in that set. However, we just established that the time- suggest that a viable approach consists of parametrizing
evolution operator generated by any one of these quantum many-body states using tensor networks and quan-
Hamiltonians can be well approximated by a polynomial- tum circuits.
size quantum circuit built from a fixed set of M discrete We thank I. Cirac, G. Vidal, and A. Winter for discus-
gates for some constant M. sions. We acknowledge funding of the ERC grant Querg,
Thus, to count the number of states that can be produced the European grant Quevadis, and the FWF SFB grants
by arbitrary time-dependent Hamiltonians, it suffices to Foqus and Vicom. D. P. is partially funded by NSERC and
count the number of polynomial-size quantum circuits FQRNT. R. S. is partially funded by NSF through the CCF
constructed from a universal discrete set of one- and program and the Laboratory Directed Research and
two-qubit gates, and to consider an " ball around the output Development program at LANL.
of each of these circuits, i.e., the set of states within a
distance " of the outputs of these circuits. Surely, the states
reached in polynomial time by arbitrary time-dependent
Hamiltonians are contained in the union of these balls. It is [1] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev.
well known [12] that such circuits can only reach expo- A 108, 1175 (1957).
nentially small subset of states. [2] R. B. Laughlin, Phys. Rev. Lett. 50, 1395 (1983).
[3] J. I. Cirac and F. Verstraete, J. Phys. A 42, 504004 (2009).
Since we limit the evolution to polynomial time, there
[4] C. Shannon, Bell Syst. Tech. J. 28, 656 (1949).
exists a constant  such that the total number of gates in the [5] S. Lloyd, Science 273, 1073 (1996).
simulation circuit is bounded by K , where K / N is the [6] C. Zalka, Proc. R. Soc. A 454, 313 (1998).
number of qubits required for the simulation. There are no [7] D. Aharonov and A. Ta-Shma, in Proceedings of the 35th
more than Ncircuits ¼ ðMK2 ÞK distinct ways of arranging

Annual ACM Symposium on Theory of Computing, 2003,
these gates into a quantum circuit (M possibility for each p. 20
gate and K2 possible pairs of qubits between which it can [8] D. W. Berry, G. Ahokas, R. Cleve, and B. C. Sanders,
be applied), and therefore no more than Ncircuits distinct Commun. Math. Phys. 270, 359 (2006).
states that can be produced. On the other hand, the states of [9] D. W. Berry and A. M. Childs, arXiv:0910.4157.
K qubits live on a (2Kþ1  1)-dimensional hypersphere, [10] A. M. Childs and R. Kothari, arXiv:1003.3683.
[11] J. Huyghebaert and H. De Raedt, J. Phys. A 23, 5777
whose surface area is S ¼ 22 =ð2K Þ, and an " ball
K

(1990).
around a given state is a (2 Kþ1
 2)-dimensional hyper- [12] M. A. Nielsen and I. L. Chuang, Quantum Computation
sphere of volume V ¼ 22 1 "2 2 =ð2K Þ. Combining,
K Kþ1
and Quantum Information (Cambridge University Press,
we see that the " balls of physical states occupy only an Cambridge, England, 2000).
¼ OðKK 2 Þ of the
K
exponentially small fraction Ncircuits
S
V
[13] A. Kitaev, A. H. Shen, and M. N. Vyalyi, Classical and
total volume in Hilbert space. Thus, the overwhelming Quantum Information (AMS, Providence, 2002).
majority of states in the Hilbert space of a quantum [14] R. P. Feynman, Int. J. Theor. Phys. 21, 467 (1982).
many-body system can only be reached after a time scaling [15] H. Trotter, Proc. Am. Math. Soc. 10, 545 (1959).
[16] M. Suzuki, Proc. Jpn. Acad. 69, 161 (1993).
exponentially with the number of particles.
[17] C. M. Dawson and M. A. Nielsen, Quantum Inf. Comput.
Conclusion—We demonstrated that any time-dependent 6, 81 (2006).
local Hamiltonian can be simulated efficiently using a [18] N. Wiebe, D. Berry, P. Høyer, and B. C. Sanders, J. Phys.
quantum computer, independent of the frequencies in- A 43, 065203 (2010).
volved. As an application, we showed that the set of quan- [19] See supplemental material at http://link.aps.org/
tum states that can be reached from a product state with a supplemental/10.1103/PhysRevLett.106.170501 for de-
polynomial-time evolution of an arbitrary time-dependent tails of the proof.
quantum Hamiltonian is an exponentially small fraction of [20] Some discretization may be unavoidable when sampling
the Hilbert space. This means that the vast majority of HðtÞ on a digital computer or an oracle, but this computa-
quantum states in a many-body system are unphysical, as tional cost grows only logarithmically with the inverse
they cannot be reached in any reasonable time. As a con- step size.
[21] C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg,
sequence, all physical states live on a tiny submanifold, and
Atom-Photon Interactions: Basic Processes and
that manifold can easily be parametrized by all poly-sized Applications (John Wiley and Sons, New York, 1992).
quantum circuits. Although quantum circuits are unitary [22] D. F. V. James and J. Jerke, Can. J. Phys. 85, 625 (2007).
and do not directly simulate imaginary time evolution, the [23] G. Vidal, Phys. Rev. Lett. 99, 220405 (2007); Phys. Rev.
counting argument we presented holds equally well in this Lett. 101, 110501 (2008).
setting, and hence thermal and ground states of local [24] F. Verstraete, J. I. Cirac, and V. Murg, Adv. Phys. 57, 143
Hamiltonians are also efficiently parametrized by short (2008).

170501-4

You might also like