Quantum Computers
Quantum Computers
Abstract
Algorithms are described for efficiently simulating quantum mechanical systems on quantum computers. A class of al-
gorithms for simulating the Schr6dinger equation for interacting many-body systems are presented in some detail. These
algorithms would make it possible to simulate nonrelativistic quantum systems on a quantum computer with an exponential
speedup compared to simulations on classical computers. Issues involved in simulating relativistic systems of Dirac or gauge
particles are discussed. © 1998 Elsevier Science B.V. All rights reserved.
1. I n t r o d u c t i o n
There has been a great deal of excitement in the field of quantum computing over the last few years, largely
due to the result of Shor [1] showing that large numbers can be factored on a quantum computer in a time that
scales polynomially with the number of digits. Shor's work was based on earlier results showing that certain oracle
problems could be solved on a quantum computer exponentially faster than on a classical computer [2]. Since
it is believed that factoring is an exponentially hard problem on a classical computer, Shot's result provided for
the first time a concrete demonstration that quantum computers can achieve useful results much more efficiently
than classical computers. Despite a great deal of work in this field, however, there has been little progress toward
discovering other algorithms which give the quantum computer an exponential advantage in performance.
Early in the development of quantum computing, it was suggested by Feynman [3] that it might be possible to
simulate quantum mechanical systems exponentially faster on quantum computers than on classical computers.
Feynman argued that understanding this issue might help to quantify the differences between the computational
capacities of classical and quantum computers. Recently, it was argued by Lloyd [4] that for a fairly general class of
quantum systems, particularly for discrete systems, it should be possible to achieve an exponential speedup when
* Expanded version of a talk given by Washington Taylor IV at the PhysComp '96 conference, Boston University, Boston, November
1996.
* Corresponding author. E-marl: [email protected].
I E-mail: [email protected].
simulating quantum mechanical systems on a quantum computer. In this paper we will discuss algorithms which
are concrete realizations of these general arguments.
There are three particular types of quantum theories which we would like to try to simulate on a quantum
computer:
- Nonrelativistic many-body systems (the Schr6dinger equation).
- Relativistic many-fermion systems (the many-body Dirac equation).
- Gauge field theories (quantum Yang-Mills theories).
In this paper we describe in some detail a class of algorithms for simulating the many-body Schr6dinger equation.
These algorithms were presented from a different point of view in [5,6]. Based on simple quantum cellular automata
(QCA) [7] and quantum lattice-gas automata (QLGA) [8] models, these algorithms can be used to simulate systems
of interacting nonrelativistic quantum particles with speedup exponential in the number of particles in the system.
We give a short discussion of progress which has been made toward developing algorithms for simulating systems
of many Dirac particles, and we discuss prospects for simulating gauge field theories on a quantum computer.
2. Quantum computers
We begin with a brief description of a universal quantum computer, in order to fix notation. For recent reviews of
quantum computing, in general, see [9]. The state space of a quantum computer is defined to be the Hilbert space
associated with a finite number (N) of two-state quantum bits (q-bits). A natural basis for this Hilbert space is given
by the set of 2 N states
A single unit of computation is defined to be the action of an arbitrary two-bit gate on the state of the system. Mathe-
matically, this corresponds to acting on the state of the system with a 2 N x 2 N matrix which is a tensor product of an
arbitrary 4 x 4 unitary matrix (acting on the Hilbert space associated with a pair of q-bits ai, aj), with an identity oper-
ator of dimension 2 N - 2 x 2 N-2" It has been shown [ 10] that such two-bit gates are universal for quantum computation.
A measurement in a quantum computer is performed by measuring the states of some of the q-bits. It is sufficient
to restrict attention to measurements of the states of individual spins with respect to the canonical basis. As dictated
by quantum mechanics, a single q-bit in the quantum state (TQ'] 1" ) + ~+l $ )) will be measured to be in the state I"
(or $) with probability I~t 12 (or I~+ ]2).
The definition of a universal quantum computer is closely modeled on the conceptual framework of classical
computation. It has been suggested [ 11 ] that a more general definition of a quantum computer may be desirable, as
there may be physical systems which exhibit the capacity for performing useful quantum computation which do not
fit into this axiomatic framework. We will discuss briefly in the following section an example of a computation which
is difficult on a standard universal quantum computer, which might be more easily implemented on a generalized
quantum computer. However, we will otherwise remain within the context of the standard definition of a universal
quantum computer in terms of q-bits and two q-bit operations.
In this section we describe a class of algorithms for simulating the many-body Schr6dinger equation on a quantum
computer with exponential speedup over simulations on a classical computer. To begin, let us recall the form of the
Schr6dinger equation for a system of interacting particles moving in d dimensions.
32 B.M. Boghosian, W. Taylor 1V/Physica D 120 (1998) 3 0 4 2
The Schr6dinger equation for a single free particle of mass m moving in d dimensions is
• 0 1 02
l~lp(x, t) - 2m Z ~(x,t), (2)
!
where we have chosen units with the Planck constant set to be h = 1. If the particle is moving in the presence of an
external potential U (x), the Schr6dinger equation becomes
• 0 1 02
1~ lp(X, t) -- 2m Z ~p(x, t) + U(x)Tt(x, t). (3)
l
For a system of n identical particles interacting via a (symmetric) pairwise potential U(xi, xj), the Schr6dinger
equation is
0 1 , ~ 02
i~-~P(xL . . . . . Xn, t ) - 2m ~ Z ~ ~ ( x ' . . . . . X n ' t ) + Z U ( x j ' x k ) ~ P ( x ' . . . . . Xn, t). (4)
k=l i=1 j<k
If we wish to simulate the n-particle Schr6dinger equation in d-dimensions, one approach is to discretize space.
If we discretize so that the particles move on a spatial lattice with l lattice sites in each direction, the number of
independent components of the n-particle wavefunction grows as 16". Even for d = 3, for reasonably large values
of 1 and n, this number becomes extremely large• If we wish to simulate the system on a classical computer, the
number of independent components in the wavefunction is a lower bound both on the amount of memory needed to
store the state of the system at a fixed point in time, and also for the amount of computation needed to take a single
time step in the simulation. For a lattice size of I = 20 and a system of 20 particles, we have l 3" = 2 0 6 0 ~ 1078,
which is clearly far beyond the memory and computational resources of any imaginable classical computer. In the
following sections we will describe algorithms with which this system can be simulated on a quantum computer
with memory and computational requirements per time step on the order of Id, independent of n (so long as n << l J).
We begin our discussion of simulations on a quantum computer by considering the simplest case: a free
Schr6dinger particle moving on a lattice in one dimension. Let us consider a one-dimensional lattice with l vertices.
We will associate a single q-bit ai with each of the vertices i. Let us associate a q-bit o-i in the state ]" with the presence
of a "particle" at the ith lattice site, and a q-bit in the state $ with the absence of a particle. The canonical basis for
the Hilbert space contains 2 l states, with anywhere between 0 and I particles in the available states. Let us restrict
our attention to the single-particle subspace of the Hilbert space. This subspace has a basis containing l states; we
denote by Ik) the state where the single particle is contained at lattice site k. Graphically, this state would look like
... ~ ,~ $ ~ $ $ ,~ ...
•. . k-1 k k+l . . .
We now wish to define a discrete time dynamics on the Hilbert space which gives a Schr6dinger equation in the
continuum limit• Writing the wave function describing the single-particle state at time t in the form
the equation of motion for evolution over a single time step will be
where M is a unitary matrix. One way of getting a Schr6dinger equation would be to find an operation on the full
Hilbert space which had the effect in the single-particle Hilbert space of acting by the operator
/b a 0 0 0 0 0 a
a b a 0 0 0 0 0
0 a b a 0 0 0 0
0 0 a b 0 0 0 0
M= .. (7)
0 0 0 0 .. b a 0 0
0 0 0 0 .. a b a 0
0 0 0 0 .. 0 a b a
a 0 0 0 .. 0 0 a b
where a, b are complex numbers satisfying Ib[2 + 21a21 : 1 and a/~ + fib = 0. Unfortunately, this operator cannot
be implemented by a simple set of two q-bit operations [7,8]. One way of seeing that this is not possible is to take
the inverse of the matrix M. The matrix M - l is a dense matrix for generic lattice size l, indicating that M cannot
be implemented by performing some sequence of local two q-bit operations at each lattice site in a way which is
independent of l.
It is worth noting at this point that although the operation M cannot be simply realized in the framework of
universal quantum computers using q-bits and two q-bit operations, there are simple physical systems which behave
in a fashion very similar to the dynamics which this operation would define. In particular, a lattice of s p i n - l / 2
particles with a Hamiltonian proportional to the sum of exchange operations, as discussed in [12], would lead to a
very similar dynamics. Thus, it may be advantageous to consider a wider range of quantum computational devices
in searching for algorithms for simulating generic quantum systems.
Returning to the framework of universal quantum computers, we now define a slightly more complicated dynamics
which can be implemented using two q-bit operations, and which leads to the desired Schr6dinger equation. Consider
the two operators acting on the single-particle Hilbert space
/b a 0 0 .-- 0 0 0 0~ (b 0 0 0 ••- 0 0 0 a ~
a b 0 0 ..- 0 0 0 0 0 b a 0 ••- 0 0 0 0
0 0 b a ... 0 0 0 0 0 a b ... 0 0 0 0
0 0 a b .. 0 0 0 0 0 0 0 b ... 0 0 0 0
MI = : .. , m2= .° (8)
0 0 0 0 .. b a 0 0 0 0 0 0 '' b 0 0 0
0 0 0 0 .. a b 0 0 0 0 0 0 '' 0 b a 0
0 0 0 0 .- 0 0 b a 0 0 0 0 -" 0 a b 0
0 0 0 0 -.- 0 0 a b a 0 0 0 -" 0 0 0 b
where [al 2 + [bl 2 = 1 and a/7, + b(~ = O. We can now describe a quantum algorithm which gives a time-evolution
equation for the single-particle sector
0 b a
s = 0 a b ' (10)
0 0 0
Graphically, we can describe the implementation of the dynamics (9) through the following sequence of applications
of the operator s to pairs of q-bits
t+l
t+2
t+3
We will now take the continuum limit of the dynamics (9) as the lattice spacing 6 becomes small. In the continuum
limit, we are interested in the behavior of wavefunctions which are continuous and which therefore vary only by
order e between adjacent lattice sites. We can analyze the dynamics of these wavefunctions in the continuum limit
by performing a mode analysis. Assume that at time t = 0, we act with the matrix M2 which acts on the pairs of
q-bits (2j, 2j + 1). If at time t = 0, the even and odd lattice sites, respectively, have amplitudes given by
For a fixed value of the angular wave vector x the matrix which appears in this equation has two eigenvalues. In the
continuum limit, where to6 --~ 0, the eigenvectors of this matrix approach the limiting values (1, 1) and (1, - 1 ) .
Since we are interested in the behavior of wavefunctions which are smooth to order ~ at t = 0, only the first of these
eigenvectors will be relevant for describing time-development in the continuum limit. The eigenvalue associated
with the first of these eigenvectors is
we see that after scaling the time variable as 62, the wavefunction ~ obeys the dispersion relation
iw = aK2. (16)
2b
B.M. Boghosian, W. Taylor lV/Physica D 120 (1998) 30-42 35
Thus, in the continuum limit, to within an error of order E, the wavefunction ~ (x, t) obeys the Schr6dinger equation
1 2
iOt q/ (x, t) = --~m OXq/ (x, t ) , (17)
bi
m = --. (18)
a
Note that this mass is always real, due to the condition a/~ ÷ fib = 0, necessary for unitarity.
We have now seen that the dynamics (9) gives rise in the continuum limit to a Schrrdinger equation for a single
free particle moving in one dimension. A number of similar models have been considered previously by other
authors. Essentially the same model was studied by Succi and Benzi [ 13]; a number of numerical experiments were
performed on this model by Succi [ 14] to study the emergent Schrrdinger behavior in the continuum limit. A closely
related model was described by Feynman and Gibbs [15] in the context of the single-particle Dirac equation in one
spatial dimension. In Feynman's model, however, the amplitude b was taken to scale as the lattice spacing E. This
model was the basis for the recent work of Meyer [8,16,17] on the many-particle one-dimensional Dirac system;
we will discuss this system in Section 4.
We now have an algorithm for simulating a free quantum particle in one dimension. The dynamics of this system,
which are described by the dispersion relation (16), are essentially trivial. We shall now add various features to
the model to give a system of more physical interest. First, we describe how an external potential can be easily
incorporated into the system.
Given an external potential function U(x), the effect of this potential can be brought into the dynamics by acting
on each q-bit at each time step with the matrix
Ur ----
(,
0
0 )
e -i~2U(x) ' (19)
where x is the position of the q-bit. The inclusion of this operation changes the time development of the system so
that the differential equation satisfied by q* becomes
1 2
iOtql(x, t) = -~mOX q,'(x, t) + U(x)q/(x, t) (20)
in the continuum limit, precisely the Schrrdinger equation for a particle in an external potential.
It is worth noting that the continuum limit of the theory with an external potential may not be well-behaved
when the external potential is not smooth. If the potential is singular or has a discontinuity, it is possible that the
eigenvectors of the matrix in (12) with limiting form (1, - 1 ) can be excited. This issue is discussed by Meyer in
[ 16]. As long as the potential function is smooth, however, the only eigenstates of the system which are excited will
be of the (1, 1) form, and the Schr6dinger equation in an external potential will be satisfied to order E. We have
performed numerical analyses of the spectrum for simple systems and have verified that for smooth potentials the
behavior of the model closely matches the continuum system being modeled [6].
36 B.M. Boghosian, W. Taylor IV/Physica D 120 (1998) 30-42
Thus far, we have really not used the power of the quantum computer at all. The number of states available in
the single-particle sector of the Hilbert space is essentially the same as the number of q-bits, so the algorithm for a
single particle could just as well be implemented on a classical computer. Now, however, let us consider the behavior
of the multiple-particle sectors of the Hilbert space with the same algorithm described above. Keeping the notation
that a q-bit in the state 1" corresponds to the presence of a particle, we can for example graphically describe a state
with three particles at positions k - 4, k, k + 2 by
• .. ~, ¢ ¢, ~, ~, $ ¢ ~, ¢ ~, ~, + ¢ ...
• . . k-1 k k+l . . .
Because the operator s in (10) preserves particle number, the number of particles in the system is invariant under
the dynamics. If we have a lattice of size l, as long as the number of particles present n is small compared to the
lattice size (n << 1), generically particles will propagate for a fairly large number of time steps without encountering
other particles. In the limit where the lattice spacing gets arbitrarily small but the number of particles remains fixed,
we will obtain a Schr6dinger equation for n particles, which will be noninteracting except when a pair of particles
are at the same lattice site. Thus, we have described an algorithm for simulating many Schrrdinger particles in one
dimension interacting with a local (8 function) interaction potential.
By incorporating an external potential using single-particle operators like (19) we can simulate a system of n
Schrrdinger particles moving in an external potential• It is perhaps more interesting, however, to include interactions
between the particles through a pairwise potential function U(x, y). A local interaction corresponding to a potential
U (x, y) ~ 6 (x - y) with an arbitrary phase can be incorporated by including a phase ~bin the two-particle component
of the matrix s in (10). In general, however, we would like to incorporate nonlocal potential functions. In order
(00 0)
to implement the effects of a general interparticle potential, we can act at every time step on the four-dimensional
Hilbert space associated with every pair of q-bits at positions x and y with the operator
1 0 0 (21)
Ux.,, = 0 1 0 "
0 0 e-i~ZU(x v)
This operator acts on the state 1"1" with a nontrivial phase, and therefore affects only the basis states where there are
particles at positions x and y. Just as in the case of the external potential, incorporating the effects of these operators
turns the dynamical equation for qJ into a Schr6dinger equation for many interacting particles (4).
So far we have considered systems of particles moving only in a single spatial dimension. Of course, most problems
of physical interest involve particles moving in three (or more) spatial directions. We will now discuss the gener-
alization of the algorithms discussed above to higher dimensions. One simple way of simulating a d-dimensional
system is to put the q-bits on a d-dimensional lattice, and to implement the kinetic part of the time-development rule
by repeatedly using the algorithm described in Section 3.2 along each of the dimensions successively. This leads to
a fairly simple set of algorithms for simulating the many-particle interacting Schrrdinger equation in d-dimensions.
There are several aspects of this algorithm, however, which may be somewhat problematic. For one thing, acting
in each of the directions in turn with the kinetic operator (9) breaks the underlying lattice symmetry. For another
B.M. Boghosian, W. Taylor lV/Physica D 120 (1998) 30-42 37
Fig. 1. A typical basis state in the Hilbert space of a two-dimensional QLGA. Filled circles correspond to occupied particle states (q-bits
in the state I"). Arrows correspond to directions of propagation.
thing, the number of q-bits which behave differently under the time-development rule scales as 2 d, which for large
d may complicate the numerical interpretation of simulations.
A nice way of packaging the q-bits and kinetic time-development rule in d-dimensions is given by the general
class of systems called Q L G A [5,8]. These models are defined by analogy with classical lattice-gas automata [ 18] to
be simple dynamical systems in which q-bits on a lattice undergo a time-development which consists of alternating
propagation and collision steps. At each lattice site there are a set of q-bits corresponding to particle occupation
sites for particles with velocity vectors lying on the lattice. On a cartesian lattice in d-dimensions, for example, we
could have 2d q-bits at each lattice site, corresponding to sites for particles moving along each of the basis vectors.
Just as in the previous models, we associate a q-bit in state t ($) with the presence (absence) of a particle. These
models have the advantage that they can be defined in a fashion which is completely symmetric with respect to the
discrete symmetry group of the underlying lattice. Also, the number of distinct types of q-bits goes as 2d instead of
2 d as in the iterative model mentioned above. An example of a basis state for the Hilbert space of a Q L G A in two
dimensions is given in Fig. 1. The filled circles represent occupied particle sites, and the empty circles represent
unoccupied sites.
In the propagation stage of the time-development of a QLGA, the state of the system is transformed by a
permutation on the q-bits. This permutation takes the q-bit associated with each position and velocity to a q-bit at
the lattice site reached by adding the velocity vector to the position. As indicated by the arrows in the example in
Fig. 1, this transformation can be achieved by exchanging the q-bit states in pairs. For each pair ot' q-bits which
must be exchanged, the operator acting on the Hilbert space is given in the basis {I ~,$ >, I 1"$ ), I S t ), 1"1")} by
the matrix
0 1 (22)
E---- 1 0 "
0 0
In the collision stage of the time-development, the state associated with all the q-bits at each lattice site is
acted on by a unitary "collision operator". This operator is chosen to conserve particle number. In general, we
would like to choose a collision operator which is invariant under the group of discrete symmetries of the lattice.
As in the one-dimensional model discussed above, for systems where the particle number is fixed and the lattice
spacing becomes small, the part of the collision operator in the single-particle sector determines the propagation
of the individual particles. On a cartesian lattice in d dimensions, the collision operator in the single-particle
sector is parametrized by three complex phases/~, v, ,k. These phases correspond to the eigenvalues of the collision
matrix for vectors which lie in the three irreducible representations of the discrete rotation group into which the
2d-dimensional single-particle space decomposes. Specifically, /~ is the eigenvalue associated with the constant
vector (1, 1. . . . . 1) and v is the eigenvalue of vectors which are antisymmetric under a parity transformation. It
38 B.M. Boghosian, W. Taylor lV/Physica D 120 (1998) 30-42
is shown in [5] that in the generic case where the three eigenvalues are distinct, the behavior of the individual
particles in the system is governed by a Schrfdinger equation for a nonrelativistic particle with mass m given by
the relation
i v 1
-- - - + - . (23)
2m # - v 2
By combining the propagation and collision processes into a single time-development rule, we have a QLGA model
which has a continuum limit described by a many-particle Schrfdinger equation in d dimensions. As described so far,
the particles in this system interact only through local (3-function) interactions, which are parametrized by the part
of the collision matrix acting in the multiple-particle space. Just as in the one-dimensional model discussed above,
it is straightforward to include external and interparticle potentials, so that an arbitrary system of nonrelativistic
interacting quantum particles can be simulated using this type of model.
In fact, the one-dimensional model discussed in Sections 3.2-3.4 is precisely equivalent to a simple one-
dimensional QLGA. We can associate the even and odd lattice sites with right- and left-moving particle sites
at even time steps, and with left- and fight-moving particle sites at odd time steps. If we define a one-dimensional
QLGA with a collision matrix given by s as in (10), it is easy to see that the QLGA decomposes into two non-
interacting copies of the one-dimensional model defined through (9). For this choice of collision matrix we have
= a + b and v = a - b (the sign on v arises because the collision operator s implicitly changes the direction of
velocity for each particle). Inserting these values for #, v into (23) reproduces (18) as we would expect.
It should be noted that the algorithms we have discussed here are most easily used to simulate systems of interacting
bosons. The microscopic quantum system contains only a single state in the Hilbert space corresponding to each
configuration of particles. Two paths in which particles move to positions x, y and y, x will contribute amplitudes
which add with no change of sign, as is appropriate for Bose statistics. It is possible to modify the algorithm to
incorporate Fermi statistics, however this requires somewhat more bookkeeping. This issue was discussed in [19].
3.6. Computational c o m p l e x i ~
We have now described a class of algorithms for simulating the nonrelativistic many-body Schrfdinger equation
on a quantum computer. As discussed in Section 3.1, simulating n quantum particles on a lattice of size I d requires a
computation of complexity O(1 dn ) on a classical computer. We now briefly discuss the time and memory requirements
for such a simulation on a quantum computer. Using the quantum lattice-gas automaton formalism, we can simulate
such a system with 2d. I d q-bits. The propagation step takes on the order o f d l d two q-bit operations, since the q-bits
are simply exchanged pairwise. The free kinetic collision step requires on the order of d2l d operations (the factor of
d 2 is needed to construct the collision operator at each vertex from two q-bit operations). Incorporating an arbitrary
interparticle interaction potential would require on the order of d212d operations, since we must act on each pair of
q-bits independently. This operation will clearly be the most computationally expensive part of the simulation. We
see, however, that the memory and time requirements for the algorithms we have discussed are only polynomial in
the lattice size, and are completely independent of the number of particles as long as n << I d. Thus, we can achieve
a speedup for many-particle quantum simulations which is exponential in the number of particles being simulated.
For example, for a lattice of size I = 20 in three spatial dimensions, the number of operations needed in a simulation
would be on the order of 36 × 206 = 2.3 × 109 compared with 1078 operations for a simulation on a classical
computer.
In discussing the computational complexity of these simulation algorithms, it is worth noting that if we simulate
nonrelativistic fermions extra work must be done to keep track of the phase of the state. Using the bookkeeping
method suggested in [19], for example, we must define a canonical ordering for all the particles, and we must check
B.M. Boghosian, W. Taylor IV/Physica D 120 (1998) 30-42 39
to see when the propagation changes the ordering of any pair of particles. This will take roughly on the order of
d212d operations per time step,just as for the interparticle interaction. In fact, however, we can incorporate this phase
into the interparticle interaction matrix, so that the simulation of an interacting system of nonrelativistic fermions
can be achieved in precisely the same number of quantum operations as the simulation of interacting "hard" bosons
(particles with bosonic statistics which cannot occupy the same lattice site).
In the bulk of this paper we have discussed the simulation of nonrelativistic systems of many Schr6dinger
particles. Many systems of physical interest are described to a high degree of precision by the nonrelativistic many-
body Schr6dinger equation. However, there are many physical systems for which the Schr6dinger equation is not
a sufficiently accurate model. In particular, it is desirable to have a way of simulating relativistic quantum systems
such as gauge fields or Dirac fermions.
Let us first consider the problem of simulating free Dirac fermions in d dimensions on a quantum computer. As
mentioned in Section 3.2, it was found by Feynman and Gibbs [ 15] that if the parameter b is taken to approach 0 at the
same rate as the lattice spacing, the one-dimensional model described above satisfies the first-order two-component
Dirac equation. By generalizing this single-particle model to the QLGA framework, it was shown by Meyer [8] that
a free system of many noninteracting Dirac particles in 1 + 1 dimensions could be simulated on a quantum computer.
The problem of simulating Dirac fermions on a lattice in more than one spatial dimension has presented difficulties
to physicists for decades now. A full discussion of the approaches which have been used to try to solve this problem
is beyond the scope of this paper. The best-known obstacle to a sensible discretization of a field theory of Dirac
particles is the fermion doubling problem. If we try naively to generalize Feynman's model for the one-dimensional
Dirac system to higher dimensions, we immediately encounter several difficulties. First, even for a single particle
it is not straightforward to find a lattice rule which gives the propagator for the Dirac equation. Within the QLGA
framework, in fact, it is not possible to construct a single-particle Dirac equation in any dimension d > 1. A
model suggested several years ago by Bialynicki-Birula [20], however, fits very closely into the framework we have
been discussing, and may provide a way of simulating systems of many Dirac particles without encountering the
difficulties which have plagued previous efforts.
The basic approach of Bialynicki-Birula is to construct a quantum cellular automaton model with a single time-
development rule which depends only on local variables. He finds that in three spatial dimensions, on the body-
centered cubic lattice where every vertex has eight neighbors, it is possible to find such a rule whose continuous
limit corresponds to a two-component Weyl spinor. Remarkably, it seems that such a model cannot exist on the
three-dimensional cartesian lattice. A Dirac particle is formed by taking two Weyl spinors and including a mass
term by hand. As described by Bialynicki-Birula, the model avoids the no-go theorem of Nielsen and Ninomiya
[21] because it has discrete time evolution. The model of Bialynicki-Birula suffers from the same difficulty as the
model with time-development operator (7) which we discussed in Section 3.2, and thus this algorithm cannot be
implemented naturally on a universal quantum computer. However, the existence of the model suggests that by
using a multi-step rule on the body-centered cubic lattice, it might be possible to efficiently simulate free Dirac
fermions in three dimensions on a quantum computer.
If one could indeed simulate many free Dirac particles on the lattice, it is natural to try to simulate a full second-
quantized field theory of Dirac particles. To do this, one would need to introduce antiparticles corresponding to
negative energy states. Incorporating a potential directly into a Dirac theory is liable to be problematic, however,
since the simplest interaction term, the four-fermi interaction, is nonrenormalizable and would destroy the continuum
limit of the theory [22].
40 B.M. Boghosian, V~ Taylor lV/Physica D 120 (1998) 30-42
Of course, if one wishes to simulate interacting Dirac fermions, one is led naturally to the problem of simu-
lating quantum gauge field theories. Finding an algorithm with which one could simulate a quantum field the-
ory of interacting gauge fields and fermions on a quantum computer with exponential performance enhancement
would be a very interesting result. The QLGA approach is naturally suited to describing second-quantized systems
in a Hamiltonian framework, so such a model is probably best discussed in the context of Hamiltonian lattice
methods [23].
There are several difficulties inherent in constructing a QLGA model of a quantum gauge theory. First, it is
necessary to find a lattice rule which describes the propagation of a single particle according to Maxwell's equations.
This is a challenge, because it is difficult to have an isotropic propagator on a lattice when the scaling of the spatial
and temporal lattice spacing is the same. Bialynicki-Birula has suggested that his model for the Weyl spinor can
be used to describe a propagating photon by writing the Maxwell equations in spinor notation. It is unclear how to
formulate this model in QLGA language, however it does not seem that there is any fundamental obstacle to such a
formulation. Another general difficulty with simulating gauge theories is the discretization of the gauge group. In
the QLGA formalism, however, where we are essentially dealing with a second-quantized field theory, this should
not be a problem; the fermions will carry a gauge index, and the gauge particles will carry an index in the adjoint
representation of the gauge group. Thus, for example, to simulate the theory of QCD with an SU(3) gauge theory,
we would include three colors of fermions and eight species of gluons, just as in the physical particle theory we are
simulating.
To end this section, we briefly discuss a completely different approach to simulating quantum field theories on a
quantum computer which does not rely on the QLGA framework. It is always possible to describe a field theory with
a momentum cutoff by writing the Hamiltonian in terms of creation and annihilation operators for the modes of the
fields with momentum below the cutoff. In a discrete time formulation, the evolution of the state of the system in a
single time step is given by the exponential of the Hamiltonian times the time step (times i). If this time-development
operator can be reasonably approximated by a polynomial number of two-particle interactions, then such a system
could be simulated on a quantum computer with an exponential performance increase using the methods discussed
in [4]. Whether such a nonlocal approach to the simulation is preferable to the local QLGA method is not clear;
further study of both approaches will probably be needed to resolve this issue.
5. Conclusions
We have discussed algorithms for simulating quantum systems on quantum computers. In particular, we have
described in detail algorithms which enable the simulation of interacting many-body quantum systems with ex-
ponential performance enhancement over classical computers. There are many obstacles to building the sort of
large-scale quantum computers which would be necessary for the algorithms discussed here to be useful. In addi-
tion to the technical difficulties involved with precisely controlling systems of many quantum elements, there are the
theoretical problems of making algorithms robust against imprecision and loss of coherence which must appear in
any physical system. Many of the other speakers at this conference have discussed these and related issues in great
detail, so we have concentrated here on describing algorithms under the assumption that eventually the technical
problems will be solved and we will have a programmable universal quantum computer at our disposal. Certainly,
recent progress in quantum error correcting codes [24] gives some indication that there is room for optimism in this
direction, Even if a true "universal quantum computer" can never be built, however, it may be possible to construct
quantum devices which have some computational capacity [11 ]. Because the algorithms we have discussed here are
based only on simple local quantum interactions, it seems plausible that many different types of quantum systems
may be useful in simulating other quantum systems in a precise and relatively controlled fashion.
B.M. Boghosian, W. Taylor lV/Physica D 120 (1998) 30-42 41
One interesting aspect of algorithms which simulate quantum systems on a quantum computer is the issue of
measurement. In a classical computation, all the bits in the system are available at all times to the programmer, so that
the state of the system can be completely known. In a quantum computer, on the other hand, the state of the system
can only be known through the measurement of certain q-bits, whose values are determined prohabilistically through
the usual rules of quantum mechanics. This means that some information in the state of a quantum computer is not
directly accessible to the programmer. In terms of the algorithms discussed here for simulating quantum systems,
this means that only certain questions can be asked of the simulation. For example, imagine simulating a system of
a single electron in an external potential well. In principle, given the exact time-development of the wavefunction
in the quantum computer doing the simulation, a spectral analysis would immediately enable one to calculate the
entire energy spectrum of the system. However, this complete information is not available. Instead, we can only ask
the types of questions which we could ask in the real physical system - for example, we could ask for the probability
that a particle which starts at a fixed point in space moves to a given region in a fixed time. This limitation on the
set of questions which can be meaningfully asked of a quantum computer may seem at first to make the process of
extracting meaningful results from such simulations more difficult. However, since the questions which can be asked
of the simulation are precisely those questions which can be asked of the system being simulated, the connection
to results of physical experiments is very direct. Any difficulty in connecting the results of the simulation with
predictions about physical systems must be the consequence of theoretical prejudices.
Acknowledgements
We would like to acknowledge helpful conversations with Dan Abrams, Seth Lloyd, Vipul Periwal and Jeffrey
Yepez. We would like to thank the organizers of the PhysComp '96 conference for providing a stimulating environ-
ment, and Boston University for their hospitality during the conference. BMB was supported in part by an IPA from
Phillips Laboratories and in part by the United States Air Force Office of Scientific Research under grant number
F49620-95-1-0285. W T was supported by the National Science Foundation (NSF) under contract PHY96-00258.
References
[I] P.W. Shor, Polynomial-time algorithms for prime factorization and discrete logarithms on a quantum computer, AT&T preprint
q ~ a n t - p h / 9 5 0 8 0 2 7 , August 1995.
[2] D. Simon, On the power of quantum computation, in: Proceedings of the 35th Annual Symposium on Foundations of Computer
Science, IEEE Computer Society Press, Los Alamitos, CA, 1994, pp. 116-123.
[3] R. Feynman, Simulating physics with computers, Int. J. Theoret. Phys. 21 (1982) 467-488.
[4] S. Lloyd, Universal quantum simulators, Science 273 (1996) 1073.
[5} B. Boghosian, W. Taylor, A quantum lattice-gas model for the many-body Schr6dinger equation in d dimensions, BU-CCS/PUPT
preprint quant-ph/9604035, April 1996.
[6] B. Boghosian, W. Taylor, Quantum lattice-gas models for the many-body Schr6dinger equation, talk presented at the Sixth
International Conference on Discrete Fluid Mechanics, Boston University, Boston, MA, August 1996, BU-CCS/PUPT preprint
quant-ph/9701016, January 1997.
[7] G. Gr6ssing, A. Zeilinger, Quantum cellular automata, Complex Systems 2 (1988) 197-208.
[8] D. Meyer, From quantum cellular automata to quantum lattice gases, UCSD preprint quant-ph/9604003, March 1996.
[9] A. Ekert, R. Jozsa, Quantum computation and Shor's factoring algorithm, Rev. Modern. Phys. 68 (3) (1996) 733; S. Lloyd, Quantum-
mechanical computers, Sci. Amr. (Oct. 1995) 140; D.P. DiVincenzo, Quantum computation, Science 270 (1995) 255.
[10] D.E DiVincenzo, Two-bit gates are universal for quantum computation, Phys. Rev. A 51 (1995) 1015-1022; S. Lloyd, Almost any
quantum logic gate is universal, Phys. Rev. Lett. 75 (1995) 346-349; D. Deutsch, A. Barenco, A. Ekert, Universality of quantum
computation, Proc. Roy. Soc. London Ser. A 449 (1995) 669~77; A. Barenco, C.H. Bennett, R. Cleve, D. DiVincenzo, N.H. Margolus,
E Shor, T. Sleator, J. Smolin, H. Weinfurter, Elementary gates for quantum computation, Phys. Rev. A 52 (1995) 3457-3467.
[ 11] R. Feynman, Quantum mechanical computers. Found. Phys. 16, (1986) 507-531; D. Deutsch, Quantum computational networks,
Proc. Roy. Soc. London, Ser. A 425, 73-90; N. Margolus, Quantum computation, Ann. NY Acad. Sci. 480 (1986) 487-497.
42 B.M. Boghosian, W. Taylor lV/Physica D 120 (1998) 30-42
[12] B. Nachtergaele, V. Periwal, Quantum logic as a sum over classical logic gates, PUPT preprint quant-ph/9606018, June 1996.
[13] S. Succi, R. Benzi, Lattice Boltzmann equation for quantum mechanics, Physica D 69 (1993) 327-332.
[14] S. Succi, Numerical solution of the Schr6dinger equation using discrete kinetic theory, IBM ECSEC preprint, 1995.
[15] R.E Feynman, A.R. Hibbs, Quantum Mechanics and Path Integrals, McGraw-Hill, New York, 1965, pp. 35-36; see also Feynman's
unpublished notes as reproduced in S.S. Schweber, Rev. Modern Phys. 58 (1986) 449.
[16] D. Meyer, Quantum lattice gases and their invariants, Talk presented at the Sixth International Conference on Discrete Fluid
Mechanics, Boston University, Boston, MA, August 1996.
[17] D. Meyer, Quantum mechanics of lattice gas automata I. One particle plane waves and potentials, UCSD preprint
quant-ph/9611005, October 1996.
[18] U. Frisch, B. Hasslacher, Y. Pomeau, Lattice-gas automata for the Navier-Stokes equation, Phys. Rev. Lett. 56 (1986) 1505; U.
Frisch, D. d' Humibres, B. Hasslacher, E Lallemand, Y. Pomeau, J.-E Rivet, Lattice gas hydrodynamics in two and three dimensions,
Complex Systems 1 (1987) 648-707; S. Wolfram, Cellular automaton fluids 1: basic theory, J. Stat. Phys. 45 (1986) 471-526.
[19] D.S. Abrams, S. Lloyd, Simulation of Many-Body Fermi Systems on a Universal Quantum Computer, MIT preprint, November
1996.
[20] I. Bialynicki-Birula, Weyl, Dirac and Maxwell equations on a lattice as unitary cellular automata, Phys. Rev. D 49 (1994) 6920-6927.
[21] H.B. Nielsen, M. Ninomiya, A no-go theorem for regularizing chiral fermions, Phys. Lett. B 105 (1981) 219-223.
[22] L. Susskind, Lattice fermions, Phys. Rev. D 16 (1977) 3031-3039.
[23] J. Kogut, L. Susskind, Hamiltonian formulation of Wilson's lattice gauge theories, Phys. Rev. D l 1 (1975) 395.
[24] EW. Shor, Fault-tolerant quantum computation, AT&T preprint quant-ph/9605011, May 1996, and references therein.