Monday, January 28, 2013

DQC01 dissipate (or, more precisely, evacuate) the entropy of the system

http://zyk.thss.tsinghua.edu.cn/51/contents/chap13/sub13-6/13-6.htm
dissipate (or, more precisely, evacuate) the entropy of the system

Finally, optical pumping of the ancilla ion back to its initial state
j0i

constitutes the dissipative element in the sequence, which renders the dynamics of the

four system spins irreversible and enables to carry away entropy and thereby \cool" the

system qubits.

http://arxiv.org/pdf/1104.2507.pdf

http://www.univie.ac.at/qfp/publications3/pdffiles/Quantum%20computation%20and%20quantum-state%20engineering%20driven%20by%20dissipation.pdf

LETTERS



PUBLISHED ONLINE: 20 JULY 2009 |


DOI: 10.1038/NPHYS1342

Quantum computation and quantum-state

engineering driven by dissipation


Frank Verstraete
1*, Michael M.Wolf2 and J. Ignacio Cirac3*

The strongest adversary in quantum information science is

decoherence, which arises owing to the coupling of a system

with its environment

1. The induced dissipation tends to destroy

and wash out the interesting quantum effects that give rise

to the power of quantum computation

2, cryptography2 and

simulation

3. Whereas such a statement is true for many

forms of dissipation, we show here that dissipation can also

have exactly the opposite effect: it can be a fully fledged

resource for universal quantum computation without any

coherent dynamics needed to complement it. The coupling to

the environment drives the system to a steady state where

the outcome of the computation is encoded. In a similar

vein, we show that dissipation can be used to engineer a

large variety of strongly correlated states in steady state,

including all stabilizer codes, matrix product states

4, and their

generalization to higher dimensions

5.

The situation we have in mind is shown in Fig. 1. A quantum

system composed of

N particles (such as qubits) is organized in

space according to a particular geometry (in the figure, a one-

dimensional lattice). Neighbouring systems are coupled to some

local environments, which are dissipative in nature and tend to

drive the system to a steady state. Our idea is to engineer those

couplings, so that the environments drive the system to a desired

final state. The coupling to the environment will be static, so that the

desired state is obtained after some time without having to actively

control the system. Note that the role of the environments is to

dissipate (or, more precisely, evacuate) the entropy of the system,

and by choosing the couplings appropriately we can use this effect

to drive our system.

We will show first how to design the interactions with

the environment to implement universal quantum computation.

This new method, which we refer to as dissipative quantum

computation (DQC), defies some of the standard criteria for

quantum computation because it requires neither state preparation,

nor unitary dynamics

6. However, it is nevertheless as powerful as

standard quantum computation. Then we will show that dissipation

can be engineered

7 to prepare ground states of frustration-free

Hamiltonians. Those include matrix product states

4,8,9 (MPSs) and

projected entangled pair states

5,9 (PEPSs), such as graph states10

and Kitaev

11 and Levin Wen12 topological codes. Both DQC and

dissipative state engineering (DSE) are robust in the sense that,

given the dissipative nature of the process, the system is driven

towards its steady state independent of the initial state and hence

of eventual perturbations along the way.

Here, we will concentrate first on DQC, showing how given

any quantum circuit one can construct a locally acting master

equation for which the steady state is unique, encodes the outcome

of the circuit and is reached in polynomial time (with respect to

the one corresponding to the circuit). Then we will show how


1

Fakultät für Physik, UniversitätWien, 1090Wien, Austria, 2Niels Bohr Institute, 2100 Copenhagen, Denmark, 3Max-Planck-Institut für Quantenoptik,

85748 Garching, Germany. *e-mail: fverstraete@gmail.com; ignacio.cirac@mpq.mpg.de.


to construct dissipative processes that drive the system to the

ground state of any frustration-free Hamiltonian. In the Methods

section, we will prove that MPS (ref. 9) and certain kinds of

PEPS (ref. 9) can be efficiently prepared using this method, and

in Supplementary Information we will give details of the proofs.

In this letter we will not consider specific physical set-ups where

our ideas can be implemented. Nevertheless, the Methods section

will provide a universal way of engineering the master equations

required for DQC and DSE, which can be easily adapted to current

experiments

13 based on, for example, atoms in optical lattices14

or trapped ions

15. Thus, we expect that our predictions may be

experimentally tested in the near future.

Let us start with DQC by considering

N qubits in a line and a

quantum circuit specified by a sequence of nearest-neighbour qubit

operations

fUt gTt

D

1. We define j t i VDUtUt􀀀1 :::U1j0i1:::j0iN, so

that

j T i is the final state after the computation. Our goal is to find

a master equation

DL( ) with a Liouvillian in Lindblad form16

L

( )D

X


k


L

k Ly

k

􀀀

1

2




L

y

k

Lk ;



C

(1)

where the

Lk acts locally and has a steady state, 0: (1) that is unique;

(2) that can be reached in a time poly(

T); (3) such that T can be

extracted from it in a time poly(

T). As in Feynman's construction

of a quantum simulator

3, we consider another auxiliary register

with states

fjt igTt

D

0, which will represent the time. We choose

the Lindblad operators


L

i Dj0iih1jj0it h0j

L

t DUt jt C1iht jCUy

t

jt iht C1j

where

i D 1;:::;N and t D 0;:::;T. It is clear that the L terms act

locally except for the interaction with the extra register, which can

be made local as well. Furthermore,



0 D

1


T

C1

X


t


j

t ih t jjt iht j

is a steady state, that is,

L( 0)D0. Given such a state, the result of the

actual quantum computation can be read out with probability 1

=T

by measuring the time register. In Supplementary Information, we

show that

0 is the unique steady state and that the Liouvillian has

a spectral gap

1D 2=(2T C3)2. This means indeed that the steady

state will be reached in polynomial time in

T. Note that this gap is

independent of

N as well as of the actual quantum computation that

is carried out (that is, independent of the

Ut ). It is also shown that

the same gap is retained if the clock register is encoded in the unary


NATURE PHYSICS


j VOL 5 j SEPTEMBER 2009 j www.nature.com/naturephysics 633

©


2009 Macmillan Publishers Limited. All rights reserved.

LETTERS


NATURE PHYSICS DOI: 10.1038/NPHYS1342

Figure 1


j Schematic representation of the set-up.We consider a collection of N quantum particles, locally coupled to a set of environments. The

couplings are engineered in such a way that the system reaches the desired state in the long-time limit.


way proposed by Kitaev and co-workers

17, making the Lindblad

operators strictly local. A sketch of the proof is as follows. First, we

do a similarity transformation on

Lthat replaces all gatesUi with the

identity gates, showing that its spectrum is independent of the actual

quantum computation. Second, another similarity transformation

is done that makes

L Hermitian and block-diagonal. Each block can

then be diagonalized exactly leading to the claimed gap.

In some sense, the present formalism can be seen as a robust

way of doing adiabatic quantum computation

18 (errors do not

accumulate and the path does not have to be engineered carefully)

and implementing quantum random walks

19, and it might therefore

be easier to tackle interesting open questions, such as the quantum

probabilistically-checkable-proofs theorem, in this setting

20. In

addition, it seems that the dissipative way of preparing ground states

is more natural than to use adiabatic time evolution, as nature itself

prepares them by cooling.

Let us now turn to DSE and consider again a quantum system

with

N particles on a lattice in any dimension. We are interested in

ground states

, of Hamiltonians

H

D

X




H


that are frustration-free, meaning that

minimizes the energy of

each

H individually, and local in the sense thatH acts non-trivially

only on a small set

f1; ::: ;Ng of sites (for example, nearest

neighbours). We can assume the terms

H to be projectors and

we will denote the orthogonal projectors by

P D1􀀀H . States

of the considered form are, for example, all PEPS (including MPS

and stabilizer states

21).

We will consider discrete time evolution generated by a trace-

preserving completely positive map instead of a master equation.

These two approaches are basically equivalent

22 as every local

completely positive map

T can be associated with a local Liouvillian

through

L( )DNTT ( )􀀀 U, which leads to the same fixed points

and spectrum.Wechoose completely positive maps of the form


T

( )D

X




p


"


P

P C

1


m


X

m

i

D1

U

;iH H Uy

;

i

#


(2)

where the

p terms are probabilities and U ;1; :::;U ;m is a set of

unitaries acting non-trivially only within region

. They effectively

rotate part of the high-energy space (with support of

H ) to the

zero-energy space, so that tr

TT ( ) U trT U increases. As for

Liouvillians (1), we could similarly take

L ;i D UiH , or the ones

associated with the completely positive map.

We show now that for every frustration-free Hamiltonian,

the completely positive map in equation (2) converges to the

ground-state space if we choose the unitaries

U ;i to be completely

depolarizing, that is,

T ( ) /

P



P P C 1 tr TH U=trT1 U.

For ease of notation, we will explain the proof for the case of a

one-dimensional ring with nearest-neighbour interactions labelled

by the first site

D1;:::;N. Assume is such that its expectation

value with respect to the projector

onto the ground-state space

of

H is non-increasing under applications of T , that is, in particular

tr

T UDtrTT N ( ) U. Expressing this in the Heisenberg picture in

which

T ( )D C

P



H tr ( )=(d2N), we get

tr

T U trT UC

1

(

d2N)N tr

"




X

N


D1

Y

N


D1

􀀀


H

C tr C



(

)

#



trT UC


N

(

d2N)N trT HU

where the first inequality comes from discarding (positive) terms in

the sum and the second one is due to bounding all partial traces

of

H from below by the respective smallest eigenvalue . Note

that the latter is strictly positive unless

H has a product state as

the ground state (in which case the statement becomes trivial).

Hence, we must have tr

T HUD0; that is, is a ground state of H.

It is easily seen that the same argument applies for more general

interactions on arbitrary lattices.

Once we have shown that the steady state after the application

of the completely positive map lies within the desired subspace

(the ground-state space of the frustration-free Hamiltonian), the

next question to be addressed is how efficient the process is. This

depends on the spectral gap,

, of the completely positive map (or,

equivalently, of the corresponding Liouvillian), as the time to reach

the steady state,

D O(1= ). Thus, the above procedure will be

efficient as long as the gap vanishes only polynomially with the

number of systems,

N. Similarly to what occurs with many-body

Hamiltonians, the determination of such a gap is, in general, very

complicated. For a wide range of interesting models, however, it

can be proved that this gap scales favourably. This is the case for

all MPS as well as for a rich subfamily of PEPS that includes all

stabilizer states (such as Kitaev's toric code

11 and the Levin Wen

states

12). In the Methods section, we characterize such a subfamily

of states, and in Supplementary Information we give the technical

proofs of our statements. Here, we will qualitatively explain how

our method works efficiently for some families of states. For that

we note that the action of the completely positive map (2) can

be interpreted as randomly choosing a region

(according to p ,

which we may set equal to 1

=N), then measuring P and applying

a correction according to the unitaries if the outcome was negative.

We denote by

Rn the set of regions where ' satisfies the condition

H

j'i D 0. If we measure now in one of those regions, we will

obviously obtain a positive result, and thus

Rn will remain the

same. If we measure in another region, we may have a positive or

negative result, something that may change the set

Rn. By imposing

certain conditions on the operators

H and U ;i, we can make sure

that in each step

Rn cannot be reduced and that the probability of

634


NATURE PHYSICS j VOL 5 j SEPTEMBER 2009 j www.nature.com/naturephysics

©


2009 Macmillan Publishers Limited. All rights reserved.

NATURE PHYSICS


DOI: 10.1038/NPHYS1342 LETTERS

being enlarged is non-vanishing. This automatically ensures that

the

scales only polynomially with the number of systems. In

one dimension, however, one can get rid of all those restrictions

and show that any MPS can be prepared in a time that also scales

favourably with

N. The fact that all MPS states can be prepared

with our method, together with the results reported in refs 23, 24,

automatically implies the existence of phase transitions driven by

dissipation in the following sense. By changing the parameters of

the operators

H appearing in the completely positive map (2), we

change the steady state of that map. It is possible to choose models

for which that state changes abruptly at some particular value of

that parameter in such a way that the correlation length diverges

and an order parameter appears (an example can be found in the

Supplementary Information).

We have investigated the computational power of purely

dissipative processes, and proved that it is equivalent to that of

the quantum circuit model of quantum computation. We have

also shown that dissipative dynamics can be used to create ground

states (such as MPS or PEPS) of frustration-free Hamiltonians of

strongly correlated quantum spin systems. We believe that these

new methods can be experimentally tested using atoms or ions with

current set-ups (see the Methods section).

Let us stress that we have been concerned here with a proof-

of-principle demonstration that dissipation provides us with an

alternative way of carrying out quantum computations or state

engineering. We believe, however, that much more efficient and

practical schemes can be developed and adapted to specific

implementations. We also think that these results open up

some interesting questions that deserve further investigation: for

example, how the use of fault-tolerant computations can make

our scheme more robust, or how one can design translationally

invariant completely positive maps that prepare MPS more

efficiently, or the importance and generality of the set of commuting

Hamiltonians (see the Methods section), which is intimately

connected to the fixed points of the renormalization group

transformations on PEPS (as it happens with MPS; ref. 25).

Furthermore, the model ofDQCmight well lead to the construction

of new quantum algorithms, as, for example, quantum random

walks can more easily be formulated within this context. Finally,

other ideas related to this work can be easily addressed using the

methods introduced; for example, thermal states of commuting

Hamiltonians can be engineered using DSE because the Metropolis

way of sampling over classical spin configurations can be adopted

to the case of commuting operators. Similar techniques could be

applied to free fermionic and bosonic systems, and, more generally,

it should be possible to devise DSE schemes converging to the

ground or thermal states of frustrated Hamiltonians by combining

unitary and dissipative dynamics.


Note added.

Concurrently with the submission of this paper,

refs 26 and 27 appeared in which a similar quantum-reservoir

engineering was used to prepare many-body states and non-

equilibrium quantum phases.


Methods


Engineering dissipation.

Here we show how to engineer the local dissipation that

gives rise to the master equations (1) and completely positive maps (2). They are

composed of local terms, involving few particles (typically two), so that we just have

to show how to implement those. To simplify the exposition, we will treat those

particles as a single one and assume that one has full control over its dynamics (for

example, one can apply arbitrary gates).

Let us start with the completely positive maps. It is clear that by applying a

quantum gate to the particle and a `fresh' ancilla and then tracing the ancilla one

can generate any physical action (that is, completely positive map) on the system.

Furthermore, by repeating the same process with short time intervals one can

subject the system to an arbitrary time-independent master equation. This last

process may not be efficient. An alternative way works as follows. Let us assume

that the ancilla is a qubit interacting with a reservoir such that it fulfils a master

equation with Liouville operator

La D

p


􀀀

􀀀, where 􀀀 Dj0ih1j. Now, we couple

the ancilla to the system with a Hamiltonian

H D( 􀀀Ly C

y


􀀀

L). In the limit

􀀀

 , one can adiabatically eliminate the level j1i of the ancilla28 by applying

second-order perturbation theory to the Liouvillian (albeit for non-Hermitian

operators). In this way we obtain an effective master equation for

describing

the system alone, with Liouville operator

=

p


􀀀

L. By using several ancillas

with Hamiltonians

H D( 􀀀Li C

y


􀀀

Lyi

) and following the same procedure we

obtain the desired master equation. Although we have not specified here a physical

system, one could use atoms. In that case, the ancilla could be an atom itself with


j

0i and j1i an electronic ground and excited level, respectively, so that spontaneous

emission gives rise to the dissipation. The coupling to the system (other atoms)

could be achieved using standard ideas used in the implementation of quantum

computation using those systems

13.

Efficient state preparation.

We have shown that it is possible to engineer

dissipative processes that prepare ground states of frustration-free Hamiltonians in

steady state. In the proof, the time for this preparation scales as

NN , which may be

an issue for experiments with large number of particles. Here we give much more

efficient methods for certain classes of frustration-free Hamiltonians.

We consider first frustration-free Hamiltonians for which

TH ;H UD0 and

show that, under certain conditions, the corresponding ground states can be

prepared in a time that scales only polynomially with the number of particles. The

corresponding set of ground states contains important families, such as stabilizer

states (for example, cluster states and topological codes), or certain kinds of PEPS,

namely, those that have (commuting) parent Hamiltonians with the injectivity

condition (as defined in refs 8, 29). Note that there was no known way of efficient

preparation for the latter.

Loosely speaking, we will consider two classes of Hamiltonians.

(1) Hamiltonians for which all excitations can be locally annihilated. In this case the

time of convergence scales as

DO(logN). (2) Interactions where excitations have

to be moved along the lattice before they can annihilate and

DO(NlogN).

To see how the first case can occur notice that, when iterating

T , the

correction on

does not change the outcome of previous measurements on

neighbouring regions because


8

6D 0V TU ;i;H 0 UD0 (3)

In fact, this can always be achieved by regrouping the regions into larger ones

having an interior

I ( ) on which only H acts non-trivially and letting the

U

;i solely act on I ( ). Denote by q the largest probability for obtaining twice a

negative measurement outcome on the same region

. The energy trTHT M( )U

after

M applications of T decreases then as N(1􀀀(1􀀀q)=N)M such that it takes

O

((NlogN)=(1􀀀q)) steps to converge to a ground state. The relaxation time of the

corresponding Liouvillian is thus

DO(logN1=1􀀀q). Clearly, this is a reasonable

bound only if

q<1, a condition possibly incompatible with equation (3).

Note that for all stabilizer states we can achieve

qD0, because there exists

always a local unitary (acting on a single qubit) so that

H U H D0. A class of

stabilizer states where this is compatible with equation (3) are the so-called graph

states

10. In this case, labels (with some abuse of notation) a vertex of a graph and

H

D(1􀀀 ( )

x


Q


(

; )2E ( )

z

)=2, where ( ) is a Pauli operator acting on site and

E

is the set of edges of the graph. Obviously, U D ( )

z

does the job. In this special

case, we can get even faster convergence when using the Liouvillian


L

( )D

X




U

H H Uy





􀀀


1

2


n


H

;

o


C


The corresponding relaxation time can be determined exactly by realizing

that the spectrum of

L equals that of 􀀀(H 1C1H)=2 so that D1 (see

Supplementary Information).

For the second type of commuting Hamiltonians, equation (3) and

q<1 are

incompatible. However, we can still prove fast convergence by relaxing equation (3)

such that within each region

the U acts on a site closest to a predetermined

site (say the origin) on the lattice and thus commutes with all terms

H that are

further away (see Supplementary Information for details). In this way excitations

are moved over the lattice before they can annihilate. As this requires extra time

proportional to the system's size, we get

DO(NlogN).

We turn now to another family of ground states of frustration-free

Hamiltonians, namely MPS (ref. 9). For the sake of clearness, we will consider

here translationally invariant Hamiltonians, although the analysis can be

straightforwardly extended to systems without that symmetry. We will specify a

completely positive map to prepare states of the form


j

iD

X

d

i

D1

tr(

Ai1 :::AiN )ji1 :::iNi

where the

A terms are DD matrices. We assume the injectivity property29, which

implies that

is the unique ground state of a nearest-neighbour frustration-free

NATURE PHYSICS


j VOL 5 j SEPTEMBER 2009 j www.nature.com/naturephysics 635

©


2009 Macmillan Publishers Limited. All rights reserved.

LETTERS


NATURE PHYSICS DOI: 10.1038/NPHYS1342

`parent' Hamiltonian that has a gap. Denoting by

the reduced density operator

corresponding to particles

k and k C1, Hk and Pk D 1􀀀Hk will denote the

projectors onto its kernel and range, respectively. Note that tr(

Pk ) D D2. We

take

N D2n for simplicity, but this is clearly not necessary. We construct the

channel

T in several steps. We first define a channel acting on two neighbouring

particles

k;kC1, as follows

R

r;c (X) VD PkXPk C

P

k

D

2 tr(HkX)

Here,

k D2r􀀀1(2c 􀀀1), where r D1;:::;n and c D1;:::;2n􀀀r . The action of these

maps has a tree structure, where the index

r indicates the row in the tree, whereas c

does it for the column. Now we define recursively,


S

r;c VD

(1

􀀀 r )

2

(

Sr􀀀1;2c CSr􀀀1;2cC1)C rRr;c

Here,

r D2;:::;n, c D1;:::;2n􀀀r , S1;c VDR1;c and rC1 D1=Mr , where M DCN2

and

C  1 (see Supplementary Information). Note that Sr;1 acts on the first 2r

particles,

Sr;2 on the next 2r and so on. We finally define

T

VD (1􀀀 nC1)Sn;1C nC1Rn;2 (4)

In the Supplementary Information, we show that this map achieves the fixed point

(up to an exponentially small error in

C) in a time O(Nlog2 (N)). The intuition

behind the completely positive map (4) is that the channels

S1;c , which are the ones

that most often applied, project the state of every second nearest neighbour onto

the right subspace. Then

S2;c do the same with half of the pairs that have not been

projected. Then

S3;c does the same on half of the rest, and so on.

Received 11 March 2008; accepted 18 June 2009; published online

20 July 2009


References


1. Aliferis, P., Gottesman, D. & Preskill, J. Quantum accuracy threshold for

concatenated distance-3 codes.

Quant. Inf. Comput. 6, 97 165 (2006).

2. Nielsen, M. A. & Chuang, I. L.

Quantum Computation and Quantum

Information

(Cambridge Univ. Press, 2000).

3. Feynman, R. P. Simulating physics with computers.

Int. J. Theor. Phys. 21,

467 488 (1982).

4. Fannes, M., Nachtergaele, B. & Werner, R. F. Finitely correlated states on

quantum spin chains.

Commun. Math. Phys. 144, 443 490 (1992).

5. Verstraete, F. & Cirac, J. I. Renormalization algorithms for

quantum-many body systems in two and higher dimensions. Preprint at


<

http://arxiv.org/abs/cond-mat/0407066> (2004).

6. DiVincenzo, D. P. The physical implementation of quantum computation.


Fortschr. Phys.

48, 771 783 (2000).

7. Poyatos, J. F., Cirac, J. I. & Zoller, P. Quantum Reservoir Engineering with

laser cooled trapped ions.

Phys. Rev. Lett. 77, 4728 4731 (1996).

8. Perez-Garcia, D., Verstraete, F., Wolf, M. M. & Cirac, J. I. Matrix product state

representations.

Quant. Inf. Comput. 7, 401 430 (2007).

9. Verstraete, F., Murg, V. & Cirac, J. I. Matrix product states, projected entangled

pair states, and variational renormalization group methods for quantum spin

systems.

Adv. Phys. 57, 143 224 (2008).

10. Briegel, H. J. & Raussendorf, R. Persistent entanglement in arrays of interacting

qubits.

Phys. Rev. Lett. 86, 910 913 (2001).

11. Kitaev, A. Y. Fault-tolerant quantum computation by anyons.

Ann. Phys. 303,

2 30 (2003).

12. Levin, M. A. & Wen, X. G. String-net condensation: A physical mechanism for

topological phases.

Phys. Rev. B 71, 045110 (2005).

13. Cirac, J. I. & Zoller, P. New frontiers in quantum information with atoms and

ions.

Phys. Today 57, 38 44 (2004).

14. Bloch, I., Dalibard, J. & Zwerger, W. Many-body physics with ultracold gases.


Rev. Mod. Phys.

80, 885 964 (2008).

15. Leibfried, D., Blatt, R., Monroe, C. & Wineland, D. Quantum dynamics of

single trapped ions.

Rev. Mod. Phys. 75, 281 324 (2003).

16. Lindblad, G. On the generators of quantum dynamical semigroups.


Commun. Math. Phys.

48, 119 130 (1976).

17. Kempe, J., Kitaev, A. Y. & Regev, O. The complexity of the local Hamiltonian

problem.

SIAM J. Comput. 35, 1070 1097 (2004).

18. Aharonov, D.

et al. Adiabatic quantum computation is equivalent to standard

quantum computation.

SIAM J. Comput. 37, 166 194 (2007).

19. Kempe, J. Quantum random walks an introductory overview.

Contemp. Phys.

44,

307 327 (2003).

20. Arora, S. & Safra, S. Probabilistic checking of proofs: A new characterization of

NP.

J. ACM 45, 70 122 (1998).

21. Gottesman, D. A theory of fault-tolerant quantum computation.

Phys. Rev. A

57,

127 137 (1998).

22. Wolf, M. M. & Cirac, J. I. Dividing quantum channels.

Commun. Math. Phys.

279,

147 168 (2008).

23. Wolf, M. M., Ortiz, G., Verstraete, F. & Cirac, J. I. Quantum phase transitions

in matrix product systems.

Phys. Rev. Lett. 97, 110403 (2006).

24. Verstraete, F., Wolf, M. M., Perez-Garcia, D. & Cirac, J. I. Criticality, the area

law, and the computational power of PEPS.

Phys. Rev. Lett. 96, 220601 (2006).

25. Verstraete, F., Cirac, J. I., Latorre, J. I., Rico, E. & Wolf, M. M.

Renormalization-group transformations on quantum states.

Phys. Rev. Lett.

94,

140601 (2005).

26. Diehl, S.

et al. Quantum states and phases in driven open quantum systems

with cold atoms.

Nature Phys. 4, 878 883 (2008).

27. Kraus, B.

et al. Preparation of entangled states by quantum Markov processes.

Phys. Rev. A

78, 042307 (2008).

28. Cohen-Tannoudji, C., Dupont-Roc, J. & Grynberg, G.

Atom-Photon

Interactions

(Wiley, 1992).

29. Perez-Garcia, D., Verstraete, F., Cirac, J. I. & Wolf, M. M. PEPS as unique

ground states of local Hamiltonians.

Quant. Inf. Comput. 8, 0650 0663 (2008).

Acknowledgements


We thank D. Perez-Garcia for discussions and acknowledge financial support by

the EU projects QUEVADIS, SCALA, the FWF, QUANTOP, FNU, SFB FoQuS, the

DFG, Forschungsgruppe 635, the Munich Center for Advanced Photonics (MAP)

and Caixa Manresa.


Author contributions


All authors have contributed equally to this paper.


Additional information


Supplementary information accompanies this paper on www.nature.com/naturephysics.

Reprints and permissions information is available online at http://npg.nature.com/

reprintsandpermissions. Correspondence and requests for materials should be

addressed to F.V. or J.I.C.


636


NATURE PHYSICS j VOL 5 j SEPTEMBER 2009 j www.nature.com/naturephysics

©

2009 Macmillan Publishers Limited. All rights reserved.

No comments:

Post a Comment