Sei sulla pagina 1di 14

Entanglement guides quantum computation

Arun K. Pati

Citation: AIP Conference Proceedings 864, 114 (2006); doi: 10.1063/1.2400883


View online: http://dx.doi.org/10.1063/1.2400883
View Table of Contents: http://aip.scitation.org/toc/apc/864/1
Published by the American Institute of Physics
Entanglement guides quantum computation
Arun K. Pati
Institute of Physics, Bhubaneswar-751005, Orissa, India

Abstract. Origin of the speed-up in quantum algorithms is not yet fully understood. But there are
indications that entanglement does play an important role in quantum computation. Algorithms that
do not involve entanglement can be efficiently simulated on a classical computer. Here I make a
simple observation. I show that infinitesimal change in multi-particle pure product state always
gives rise to an entangled state. During quantum computation even though at each time instant the
state is not entangled, it does pass through entangled states at infinitesimal time steps. This applies
to multi-particle pure, pseudo-pure and general mixed states as well. This suggests that even though
unitary operators do not produce any entanglement, quantum entanglement guides the process of
quantum computation.
Keywords: Quantum entanglement, quantum computing, speed-up
PACS: <Missing classification>

INTRODUCTION:
Quantum entanglement is one of the much studied subject in recent years due to its
potential application in information processing. In early days, the notion of quantum
entanglement was much debated concept. Einstein was not in favor of such a notion as
it leads to spooky-action at a distance [1]. But, now we know that the spooky-action
is not so disturbing as it cannot be used on its own for faster than light communica-
tion. However, supplemented by classical communication, quantum entanglement can
become a resource for very useful and exotic information processing tasks. Many im-
portant tasks like super dense coding, quantum teleportation, remote state preparation,
quantum cryptography etc require quantum entanglement. It is also argued that quantum
entanglement may play an important role in quantum algorithms [2] and in giving extra
power to quantum computers [3].
Is it also at the heart of quantum computation? Yes, indeed. I make a simple yet
an important observation that could throw some light on the role of entanglement in
quantum evolution and this in turn may answer the question where from the extra power
comes for quantum computation.
Usual quantum computation paradigm involves preparation of initial logical states
and application of sequence of unitary evolution operators (prescribed by a particular
quantum mechanical algorithm) and then finally reading out the desired answer. In this
context an important question has been whether linear superposition alone is sufficient
to have the required speed-up or we need quantum entanglementthe weirdest feature
of quantum world. Though the existing quantum algorithms such as Deustch-Jozsa [4],
Grover [5] and Shor [6] require quantum entanglement it is not clear whether in general

CP864, Quantum Computing, edited by D. Goswami


2006 American Institute of Physics 978-0-7354-0362-8/06/$23.00
114
entanglement is the key for quantum speed-up . In particular, there has been debates in
NMR implementation of quantum algorithms as to what gives the power to quantum
computers if there is no entanglement generated during computation [7].
Origin of the speed-up in quantum algorithms is not fully understood. But there are
indications that entanglement does play an important role. Algorithms that do not in-
volve entanglement can be efficiently simulated on a classical computer. Deutsch-Joza
(DJ) algorithm for single qubit and two qubit case does not involve entanglement [8].
For three qubit or more it involves entanglement [9]. Grover algorithm for two qubit case
does not require entanglement. For more than two pure state and pseudo-pure state im-
plementations involve entanglement [10]. Shors algorithm also requires entanglement.
Since Shors algorithm is exponentially faster, it must make use of entanglement [11].
For any quantum algorithm operating on pure states one can prove that the presence of
multi-partite entanglement, with a number of parties that increases unboundedly with
input size, is necessary if the quantum algorithm is to offer an exponential speed-up
over classical computation [12]. They have also explicitly identified the occurrence of
increasing multi-partite entanglement in Shors algorithm [12].
The original version of Grovers algorithm on multiple qubits necessarily involves
quantum entanglement, even though the initial and target states are product states. Fur-
ther, by counting each active molecule as contributing to the computational resources for
pseudo-pure state machines (like in NMR), it was shown in a non-asymptotic analysis
that not only is entanglement necessary to achieve a speed-up in quantum searching, but
it must be present throughout the computation [10].
It is now clear that entanglement is obviously necessary if the unitaries are entan-
gling ones. Can we say that entanglement is necessary even if there is no entanglement
generated during quantum computation? The answer is yes, entanglement is neces-
sary even if there is no entanglement generated during quantum computation. The key
observation is the following: Quantum computation is a special type of multi-particle
quantum evolution. I argue that any multi-particle quantum evolution needs entangle-
ment, so also any quantum computation! First, I prove that generic infinitesimal change
in multi-particle state creates entanglement. Then, I go on to prove that entanglement
is necessary for any multi-particle quantum evolution. Since quantum computation is a
particular case of quantum evolution, this in turn suggests that entanglement guides the
process of quantum computation.

QUANTUM SEARCHING WITH PURE STATES AND


ENTANGLEMENTAN EXAMPLE
Let me start by giving an illustrative example of quantum searching where entanglement
does play an important role [10]. In quantum searching, we are given an unknown binary
function f x , which returns 1 for a unique target value x y and 0 otherwise, where
x 0 1 2 N 1, with N 2n . We want to find y such that f y 1. In Grovers
algorithm, the N inputs are mapped onto the states of n quantum bits (qubits) such as
spin- 12 particles. The quantum problem thus becomes one of maximizing the overlap
between the state of these n qubits and target state y . This is equivalent to maximizing
the probability of obtaining the desired state upon measurement. The initial state of these

115
qubits is taken to be an equal superposition of all possible bit strings, i.e.,
1 N 1
0 x
Nx 0
(1)

The Grover operator defined as G I0 H n Iy H n is used repeatedly in the algorithm,


where I0 2 0 0 , Iy 2 y y with y being the target (ideally the final) state
and H is the Hadamard transformation. Thus, the Grover operator corresponds to a small
rotation in the two-dimensional subspace spanned by the initial and target states. Each
such rotation requires a single evaluation of f .
Consider the pure-state version of Grovers algorithm. After k iterations of the Grover
operator the combined n qubit state (1) evolves to
cos k
k
N 1x y
x sin k y (2)

where k 2k 1 0 and 0 satisfies sin 0 1 N. The search is complete when


k 2 which takes O N iterations of the Grover operator and hence this many
evaluations of the function f .
Although the initial and target states are product states the intermediate states through
which system evolves are entangled. Since these states are superpositions of product
states they are expected to be entangled. But how much entanglement is there in these
intermediate states? This is difficult to quantify, as we do not have a proper measure
of entanglement for quantum systems consisting of an arbitrary number of subsystems.
However, we can consider the system as being bipartite, with one subsystem consisting
of a single qubit and the second subsystem all the rest. In this way, we will be able
to quantify the bipartite entanglement by calculating the reduced density matrix of any
single qubit. We then ask how close this reduced state is to a maximally entangled qubit
(using say the Hilbert-Schmidt norm criterion).
The reduced density matrix of the th qubit is
k tr1 2 1 1 n k k
a2k H 0 0H b2k j j
ak bk
2 j j j j j j (3)
N
where ak N N 1 cos k , bk sin k cos k N 1 and the single bit j
1 j, j 0 1 . Without loss of generality we take j 1 and the density matrix k
can be expressed in standard form, i.e.,
1 I
k I sk 1 sk sk P (4)
2 2
where s k tr k , s k s k s k 2 1 and P is a pure state projector. The
components of the Bloch vector s k after k iterations are
N 2 1
sx k cos2 k sin2k
N 1 N 1

116
sy k 0
1
sz k cos2 k sin2 k (5)
N 1
The bipartite entanglement in the pure state may be characterized by calculating the
von Neumann entropy of this reduced state. Using the expansion formula
1 s s 1 2s
log sP log P log 1 (6)
2 2 1 s
which holds for any 0 s 1, the von Neumann entropy may be calculated to be given
by
S k tr k log k
1 sk
1 log 1 s k
2
1 sk
log 1 s k (7)
2
The right-hand-side of this expression is independent of the choice of the remaining
qubit . Therefore, (7) holds for any one qubit versus n 1 qubit partitioning. It shows
that the reduced density matrix of the single qubit does not arise from a maximally
entangled state of n qubits, as the von Neumann entropy is not exactly unity. Since the
reduced state of Eq. (3) is not pure the full state must be entangled. To see how impure
the state in Eq. (3) is one may calculate the linear entropy L of it which is given by
1 sk 2
L k tr k k 2
(8)
2
If the linear entropy is zero the state is pure and as it approaches 12 the state approaches
a completely random mixture. In the quantum search algorithm the parameter s k can
never be zero because that would mean that cos k and sin k are simultaneously zero,
which cannot be satisfied. So although the reduced density matrix of the qubit may lie
close to the completely mixed state it can never become the identity one.
Calculate the Hilbert-Schimdt norm of the difference of the completely mixed one
and the reduced state. This Hilbert-Schmidt distance for kth iteration during quantum
search algorithm is given by
I I
d k 2
k 2
HS tr k 2
2 2
1 sk 2
L k (9)
2 2
The distance d k provides an idea of how the reduced state of an individual qubit
behaves during the kth iteration. It shows that the reduced density matrix of the qubit
differs from a completely random mixture by an order of O s k . From Eq. (5) and (9)
that for 0 sin 1 1 N and for k 2 the reduced density matrix of any remaining
qubit is pure, implying that the whole state must have been non-entangled. Thus, we see
that although the initial and target states are separable, the intermediate states through
which the system evolves are always entangled.

117
MULTIPARTICLE QUANTUM EVOLUTION NEEDS
ENTANGLEMENT
In this section, I make a simple yet an important observation that could throw some light
on the role of entanglement in quantum evolution and this in turn answers the ques-
tion where from the extra power comes for quantum computation. We show that for any
multi-particle state infinitesimal change in the pure product state gives rise to an entan-
gled state. In the language of differential geometry given any manifold of multiparticle
quantum states (product or entangled), the tangent space is an entangled manifold. This
shows that even if at each instant of time the state is non-entangled, at infinitesimal time
steps the state is entangled. Thus, any multiparticle continuous evolution requires entan-
glement. This I call the hidden power of quantum entanglement. Applied to quantum
computation, this implies that during computation even though at each time instant the
state is a product state, it does pass through entangled states at infinitesimal time steps.
In other words, even non-entangling evolution needs entanglement at infinitesimal time
steps.
Consider a composite quantum system consisting of two or more subsystems. (For
simplicity we consider bi-partite systems in finite dimensional Hilbert space, but our
results hold for any number of particles and in any dimension). Let be a set of
vectors in 1 2 . If these vectors are not normalized we can consider a set
of vectors of norm one in . The set of rays of is called the projective
Hilbert space 1 2 . If dim 1 N 1 and dim 2 N 2 then
, CN1 N2 . The
projective Hilbert space space is CN1 N2 0 U 1 which is a complex manifold
of dimension N1 N2 1 . This can also be considered as a real manifold of dimension
2 N1N2 1 . Any quantum state (product or entangled) at a given instant of time can
be represented as a point in . The evolution of the state vector can be represented
by a curve : t t in whose projection lies in . Here, smooth
mappings : 0 t of an interval into a differentiable manifold are called smooth
curves in the manifold.
Let be a differentiable manifold embedded in R, and . A vector v is
called a tangent vector to at if there is a smooth curve passing through
d
such that v dt . The tangent space T of at is the set of all tangent
vectors to at . The tangent space to a differentiable manifold at the point
is a linear space having same dimension as that of . What we will prove is that
given any multiparticle pure state , the tangent vector in infinitesimal time step
d T is entangled.

Infinitesimal change creates entanglement:


Here, first I argue that quantum entanglement is necessary for any continuous dy-
namical evolution of multi-particle system. We know that any continuous, finite time
evolution can be thought of as a limit of infinite number of sequence of infinites-
imal changes. Consider a a multi-particle state (say n-particle) which is not entan-
gled initially. Under unitary evolution the initial state evolves as 0 n
i 1 i

118
T U T 0 , where U T exp iHT and H is the total Hamiltonian of
the system. The same U T can be obtained from infinitesimal changes via U T
limN I iH T N N . Now, if U t 0 t T is capable of producing entanglement,
then the state can be written as t i1 i2 in Ci1 i2 in t 1i2 in . There is clearly
entanglement present at any stage of quantum evolution as well as during infinitesimal
time steps. The tangent vector d is also an entangled one.
The surprising thing is that even if U t does not produce any entanglement, to be
able to have a continuous evolution we need quantum entanglement. Suppose we have a
composite system that consists of two subsystems. The state of the combined system is
then 1 2 1 2 , where 1 1 and 2 2 . Now consider the
infinitesimal change in the state vector (i.e., the tangent vector at ). This is a linear
mapping d : d and can be thought of as a derivation at on a differentiable
manifold . The infinitesimal change in is given by
1 d 2 d 1 2 T (10)

The above state is clearly entangled for generic changes in the subsystems 1 and 2 as
we cannot write d as tensor product of two infinitesimal changes in the respective
Hilbert spaces. Once we choose coordinates for 1 and 2 in , then there are
no coordinates which can express (1) as product states unless d i i , i 1 2 .
But the later corresponds to stationary states, whereby the subsystems do not undergo
generic change. Now, if 0 1 1 1 N1 1 are homogeneous coordinates for 1 and
1
0 2 1 2 N2 1
are homogeneous coordinates for 2 , then the tangent vector can
2
be written as
2 1
d 1 d i 2 d i 1 2 (11)
i 2 2 i 1 1
2 1

where i1 1 2 N1 , i2 1 2 N2 , and summation over repeated indices is


understood. This is also true in any dimension and in multi-particle context. Suppose we
have a n-particle pure product state n n
i 1 i i 1 i . Then the infinitesimal
change in the state is given by
d 1 2 d n d 1 2 n (12)
which is again an entangled state for generic changes in the respective subsystems.
In other words the tangent vector to any pure product states is an entangled state.
This shows that the infinitesimal change is not a local-operation. It has the ability to
create entangled states. This is a simple but an important observation that may have
many ramifications. Here, it is not necessary that all n-particles undergo infinitesimal
change locally. For example, if n 1 -particles out of n undergo infinitesimal change
locally, then the infinitesimal change in the combined state is still entangled. But now
the entanglement is present between n 1 and the last one is left out. If we have three
particles, and the last one does not change, then the infinitesimal change in the combined
state is given by
d 1 d 2 d 1 2 3 (13)

119
Thus, we can say that when the state passes through infinitesimal changes entangle-
ment is necessary. This is because the tangent vector which tells us how the state vector
changes is typically entangled as given in (12). This is potentially one hidden power
of quantum entanglement. In any quantum universe, even if there is no direct or indirect
interaction between constituents, mere infinitesimal changes in two or more implies that
the infinitesimal change in the combined state is entangled! For product states one would
think that the global change can always be described as local changes. However, our ob-
servation shows, somewhat surprisingly, that whether the composite system is entangled
or non-entangled, global change cannot always be described as local changes.
Remarks: Note that the infinitesimal change cannot be applied as a bi-local operation,
i.e., an operation taking d d is an impossible one. This violates the
norm preservation. For bi-partite systems this would mean 1 2 1 d 2
cannot happen. We can prove this by contradiction. Suppose we have the mapping
f : d d . Then we have 1 d 1 2 d 2 . For any
normalized state we must have d as a purely imaginary number. This implies
that on lhs we have which is a purely imaginary number and on rhs we have
product of two purely imaginary numbers which is a real number. Since this cannot
hold, there is no bi-local infinitesimal changes. The proof can be generalized for more
than two subsystems. However, if one of of the subsystem does not undergo infinitesimal
change then it is possible to satisfy the norm preservation or isometric evolution. This
means we can have d I and I d but not d d .

ENTANGLEMENT COST OF CHANGE


We can also quantify how much entanglement is required for a given change in the state
vector. Geometrically, the change in the state is represented by a curve whose length
is measured in terms of Fubini-Study metric on the projective Hilbert space of the
quantum system [13, 14]. If we have two quantum states that differ infinitesimally, i.e.,
t and t dt , then the distance between them is given by
dS2 41 t t dt 2
d d i d 2
(14)
This shows that when the state changes by d the system travels a distance dS. Since
entanglement is necessary for this change, we can say that for a distance dS to be traveled
by a composite system, we need E d d amount of entanglement, where E will
be a suitable measure of entanglement for the composite system. For bi-partite system it
would be the entropy of the any one of the reduced subsystem.
To see the entangling power of infinitesimal change, let us consider two identical
copies of a (say real) qubit cos 2 0 sin 2 1 . One may be surprised to see that
the infinitesimal change in the two-qubit product state is a maximally entangled state
(unnormalized).
d d d
2d cos sin (15)
where 1
2
01 10 and 1
2
00 11 are two orthogonal Bell-
states. To be precise, the normalized form of infinitesimal changed state is a maximally

120
d
entangled state given by d cos sin T . This implies that
two identical, non-entangled qubits however far separated, when we look at the change
in the combined state through infinitesimal time steps, then the infinitesimal change in
the state is a highly non-local state. For example, the quantum mechanical correlation in
the state d (up to local unitaries) is given by

d a b d a a E a b d 2 (16)
where E a b a b is the standard quantum mechanical correlation is a maximally
entangled state. This means if one looks at change in infinitesimal steps, one may
observe violation of Bells inequality even for product states.

INFINITESIMAL CHANGE AND REDUCED DYNAMICS


In quantum information theory entanglement plays a dual role: sometimes it acts as a
perfect quantum channel and sometimes also acts as a noisy channel. When we describe
a quantum operation acting on a system, we can always imagine as a unitary
evolution on a combined system (system + ancilla) and then tracing over the ancilla. If
is the state (pure or mixed) of the system and , then tr 2 U U .
This unitary version of quantum operation always produces an entangled state which
in effect amounts to passing the system through a noisy channel. Now one may ask
since the infinitesimal change in the combined product state creates an entangled state
what kind of noise does that introduce for reduced dynamics. First, we note that the
infinitesimal change is not a unitary evolution. If we look at the unitary evolution of a
bi-partite system 1 t 2 t still it is local. The infinitesimal version of the unitary
evolution is
U1 U2 I1 d1 I2 d2
I1 I2 I1 d2 d1 I2 d1 d2 (17)
where I1 , I2 are identity matrices and d1 and d2 are infinitesimal changes acting on
1 and 2 , respectively. The infinitesimal change of the combined state is an operator
which can be represented as I1 d2 d1 I2 , which by itself is not unitary. But still we
can ask what is the reduced dynamics of any one of the subsystem. It can be verified that
when d 1 d 2 d 1 2 , then the first subsystem undergoes an
evolution as
1 1 1 1 tr2 d d d 1 d 1
1 d 1 d 1 1 2 d 2
1 1 d 2 d 2 (18)
Here, may be thought of as the quantum channel arising from global infinitesimal
changes. This clearly shows that when d , then does not simply go to d ,
rather there are additional contributions coming due to entangled nature of infinitesimal
change of the combined state. Similarly, we can see that the second subsystem undergoes

121
an evolution given by
2 2 2 2 tr1 d d d 2 d 2
2 d 2 d 2 2 1 d 1
2 2 d 1 1 (19)
This shows that when d , then 2 transforms to d 2 along with noise
terms. Eqs(18) and (19) clearly show the entangling nature of infinitesimal change. If
it has no ability to create entanglement, then 1 would have gone to d 1 and 2
would have gone to d 2 under global infinitesimal changes.

IMPLICATION FOR QUANTUM COMPUTING


Usual quantum computation paradigm involves preparation of initial logical states and
application of sequence of unitary evolution operators (prescribed by a particular quan-
tum mechanical algorithm) and then finally reading out the desired answer [15]. In this
context an important question has been whether linear superposition alone is sufficient
to have the required speed-up or we need quantum entanglementthe weirdest feature
of quantum world. Though the existing quantum algorithms such as Deustch-Jozsa [8],
Grover [10] and Shor [11] require quantum entanglement it is not clear whether in gen-
eral entanglement is the key for quantum speed-up [12] . In particular, there has been
debates in NMR implementation of quantum algorithms as to what gives the power to
quantum computers if there is no entanglement generated during computation [7].
Here comes the surprise! I show that even though the initial state of n-qubit registrar is
a product state, even though all the n-qubits during computation are product states, there
is entanglement present during computation. To see this clearly, consider the initial state
of n-qubit state prepared in equal superpositions (which is a product state)
2n 1
1
0
2
n xi (20)
i 0

At any stage of the computation (say kth step) we can write the n-qubit state generically
as
2n 1
1
k UkUk 1 U1
2
n xi (21)
i 0

Now, let us have a simple geometric way of understanding quantum computing.

Geometry of Quantum Computation


Any computation starts with an initial state and reaches a final state via a sequence of
unitary operators. This can be viewed geometrically as sequence of curves in the projec-
tive Hilbert space of n-qubit system. Therefore, geometrically any quantum computation
is a sequence of paths and efficiency of an algorithm will depend on how well we can

122
optimize these paths. If i 0 is the initial state and f is the final state then the
total distance between them is given by [13, 14]
d 2 i f 41 i f 2

If one application of U takes the state to U , then it moves a distance

d 2 i U 41 i U 2

The number of steps NS to complete the computation is then


NS d i f d i U

The success of an algorithm depends on how to minimize the number of steps. If the
system moves along geodesic paths (shortest paths) in the projective Hilbert space then
it can reach the desired state much faster. For example, in Grovers algorithm it can be
shown that the states indeed pass through geodesics and the number of steps calculated
using the above formula is exactly O N [16].

Entanglement guides the process of computation


Suppose that the unitary operators Uk s are such that at any stage of the computation
there is no entanglement generated. Can we say that entanglement has no role to play?
The answer is no. From our earlier observation, we know that even if a multi-particle
state is a product state, infinitesimal change in the multi-particle product state is an
entangled state. Thus, if we look at infinitesimal changes during quantum computation
we will see that the infinitesimal change in the n-qubit registrar is indeed entangled. For
example, the state at kth step can be written as

k u1 k u2 k un k 0 (22)

where each of these ui s are some local unitaries acting on single qubit Hilbert space
2 . But the infinitesimal change in the above state T is given by
k k

dk du1 k u2 k un k

u1 k u2 k dun k 0 (23)

which is an entangled state. This shows that the weirdest feature of quantum world
plays its role in every computation in a very subtle way. This is truly a hidden power
of quantum entanglement in quantum computation. It is hidden because, we do not look
at infinitesimal steps; we always consider finite time steps. Quantum entanglement is
necessary for any evolution (entangling or not) and hence for quantum computation.
We can say that during quantum computation possible directions in which one can pass
through a n-qubit register state is guided by entanglement.

123
One may ask does entanglement also plays any role when mixed state are involved
during quantum computation? This is exactly the case when one deals with NMR
implementations. There one typically encounters pseudo-pure states which comes as
a convex combination of a random mixture and a pure state. For n-qubits this is given
by 1 2In 0 0 , where is the purity parameter. After application of
sequence of unitary operators during certain computation the state changes as
U U 1 2In . It has been claimed that the states produced thus are
still not entangled even though the pure state component U 0 is entangled.
However, now we see how does entanglement play a role during quantum computation?
What we say is that even though the the pure state component is not entangled, the
infinitesimal change in , i.e., d indeed is entangled. This is because d d
d is an entangled one. Thus any continuous evolution of does require quantum
entanglement.
Our observation not only applies to multi-particle pure and pseudo-pure states but to
any mixed states as well. Consider a separable multi-particle (again for simplicity say
bi-partite) mixed state given by

pi i 1 i 2 (24)
i

where i 1 1 and i
2
2 are pure state components of subsystems 1
and 2, respectively with pi 0 i pi 1. We can show that even this separable state
when evolves in time, the infinitesimal change in the state is an entangled one. The
infinitesimal change in is given by

d pi i 1 d i 2 d i 1 i 2 (25)
i

To prove that d is entangled, let us assume that it is separable and then arrive at a
contradiction. If d is separable then there must be a decomposition such that we can
write this as
d wid i 1 d i 2 (26)
i

for some pure components i 1 1 and i


2
2 and wi 0 i wi 1. This
is the only form consistent with separability because the state is classically correlated
and there are infinitesimal changes for both the subsystems. This then implies that if d
is separable we have
I d d I d d (27)
which cannot be satisfied for arbitrary . Alternately, if we look at the reduced changes
then from (25) we have tr2 d i pi d i 1 and from (26) we have tr2 d 0 which is
a clear contradiction. Here, we have used the fact that tr2 d i 2 0 and tr2 d i 2 0
which is true for any pure state density operators. Thus, the infinitesimal change in any
multi-particle separable density operator is an entangled one. This shows that any multi-
particle quantum evolution be it pure or mixed does require quantum entanglement.

124
One may ask since any classical computer state can be written as a separable state
would that have the same behavior? The answer is no, because to have entangled tangent
vector we need derivative behavior and tensor product structure on the linear space. Also
one may ask if we have ordinary probabilistic description for a system comprising two
subsystems would we say that non-product states are necessary for any probabilistic
evolution as well? The answer will depend on whether we have tensor product structure
on a linear space as a form of description.

CONCLUSION
We have shown that entanglement is necessary for any quantum evolution. In short
any generic change in a quantum universe does require entanglement. Geometrically,
given a manifold of multi-particle quantum states if there are changes in two or more
subsystems then the tangent space is an entangled manifold. Since the tangent space
vectors tell the state-space vectors how to change so our result tells us something deep
about motion or change in general. Also we have shown that infinitesimal change cannot
be a bi-local operation. We have studied the reduced dynamics of the subsystem under
infinitesimal operation. This has immediate implication in quantum computation, where
one can argue that even though there is no entanglement generated during any stage
of computation, the evolution of multi-qubit state is guided by entanglement. This result
applies to pure state, pseudo-pure state and mixed state implementations as well. Though
the result of this paper may appear simple, it is nevertheless non-trivial. We hope that
this observation will unfold many other results in quantum theory and in the fast growing
field of quantum information theory.
In the spirit of the statement Space-time tells matter how to move and matter tells
space-time how to curve, I would like to conclude by saying that entanglement tells the
computation how to move and computation tells the tangent vector how to entangle.

Acknowledgment: I thank S. L. Braunstein for useful comments. Also I would like to


thank R. Jozsa for his remarks. Comments from A. K. Rajagopal, and P Gralewicz are
highly appreciated.

REFERENCES
1. A. Einstein, B. Podolsky, and N. Rosen, Phys. Rev. 47, 777 (1935).
2. R. Jozsa, Geometric Issues in the Foundations of Science, Eds. S. Huggett et al, (Oxford University
Press, Oxford, 1997).
3. R. Fitzgerlad, Phys. Today, 53(1), 20 (2000).
4. D. Deutsch and R. Proc. R. Soc. London 439, 553 (1992).
5. L. K. Grover, Phys. Rev. Lett. 79, 325 (1997).
6. P. W. Shor, Symposium on Fundamentals of Computer Science (FOCS) 56 (1994).
7. S. L. Braunstein, C. M. Caves, R. Jozsa, N. Linden, S. Popescu, and R. Schack, Phys. Rev. Lett. 83,
1054 (1999).
8. D. Collins, K. W. Kim and W. C. Holton, Phys. Rev. A 58, R1633 (1998).

125
9. Arvind, Pramana J. of Phys. 56, 357 (2001).
10. S. L. Braunstein and A. K. Pati, Quantum Information and Computation (QIC), 2(2), 399 (2002).
11. N. Linden and S. Popescu, Phys. Rev. Lett. 87, 047901 (2001).
12. R. Jozsa and N. Linden, Proc. R. Soc. Lond. A 459, 2011 (2003).
13. A. K. Pati, Phys. Lett. A 159, 105 (1991).
14. A. K. Pati, Phys. Rev. A 52, 2576 (1995).
15. M. A. Nielsen and I. Chuang, Quantum Computation and Information, (Cambridge University Press,
Cambridge, 2000)
16. A. K. Pati, Quant-Ph/9807067, (1998).
.

126

Potrebbero piacerti anche