13.07.2015 Views

Robert Raussendorf Statement and Readings - Pacific Institute of ...

Robert Raussendorf Statement and Readings - Pacific Institute of ...

Robert Raussendorf Statement and Readings - Pacific Institute of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Decoherence in quantum computation - foe or friend?<strong>Robert</strong> <strong>Raussendorf</strong>, University <strong>of</strong> British ColumbiaAbstract: Decoherence is detrimental to quantum computation because it makes the computation“noisy”. Or is it? Upon closer inspection, it turns out that decoherence can both compromize <strong>and</strong>help realize quantum computation. Which <strong>of</strong> the two applies does very much depend on thedecoherence model considered.I will start out by proving the expected, namely that decoherence, for a certain (justifiable) class <strong>of</strong>decoherence models, does indeed compromise quantum computation. In this regard, I will reviewa result <strong>of</strong> Bravyi <strong>and</strong> Kitaev [1b]/ van Dam <strong>and</strong> Howard [1a] demonstrating an upper bound tothe error threshold for fault-tolerant quantum computation. The significance <strong>of</strong> this upper boundis that no method <strong>of</strong> error correction, however clever, can put the quantum computation back ontrack if the decoherence level per elementary gate operation is above the threshold value.In the second part <strong>of</strong> my introduction, I will discuss two computational models [2], [3] that usedecoherent dynamics to realize quantum computation. In the case <strong>of</strong> [2], the computation is drivenby local projective measurements on a highly entangled quantum state. Therein, the entanglement<strong>of</strong> the initial quantum state is progressively destroyed as the computation proceeds. Thus, entanglementis a resource for this computational model. In the second case, Ref. [3], universal quantumcomputation is implemented in a dissipative quantum system whose evolution is governed by timeindependent<strong>and</strong> local couplings to the environment. Due to the purely dissipative nature <strong>of</strong> theprocess, this way <strong>of</strong> doing quantum computation exhibits some inherent robustness <strong>and</strong> defies some<strong>of</strong> the DiVincenzo criteria for quantum computation.Suggested Reading:[1a] Sergey Bravyi <strong>and</strong> Alexei Kitaev, Phys. Rev. A 71, 022316 (2005).[1b] Wim van Dam <strong>and</strong> Mark Howard, Phys. Rev. Lett. 103, 170504 (2009).[2] R. <strong>Raussendorf</strong> <strong>and</strong> H.J Briegel, Phys. Rev. Lett. 86, 5188 (2001).[3] F. Verstraete, M. Wolf <strong>and</strong> J.I. Cirac, Nature Physics 5, 633 - 636 (2009).Remark: The results <strong>of</strong> refs. [1a] <strong>and</strong> [1b] are very closely related. For background reading, Irecommend Ref. [1b] over [1a] because it is shorter. In my introduction, I will discuss [1a] (firstpart only), however, because the result therein is better suited for graphical display.1


PHYSICAL REVIEW A 71, 022316 2005Universal quantum computation with ideal Clifford gates <strong>and</strong> noisy ancillasSergey Bravyi* <strong>and</strong> Alexei Kitaev †<strong>Institute</strong> for Quantum Information, California <strong>Institute</strong> <strong>of</strong> Technology, Pasadena, 91125 California, USAReceived 6 May 2004; published 22 February 2005We consider a model <strong>of</strong> quantum computation in which the set <strong>of</strong> elementary operations is limited toClifford unitaries, the creation <strong>of</strong> the state 0, <strong>and</strong> qubit measurement in the computational basis. In addition,we allow the creation <strong>of</strong> a one-qubit ancilla in a mixed state , which should be regarded as a parameter <strong>of</strong> themodel. Our goal is to determine for which universal quantum computation UQC can be efficiently simulated.To answer this question, we construct purification protocols that consume several copies <strong>of</strong> <strong>and</strong>produce a single output qubit with higher polarization. The protocols allow one to increase the polarizationonly along certain “magic” directions. If the polarization <strong>of</strong> along a magic direction exceeds a threshold valueabout 65%, the purification asymptotically yields a pure state, which we call a magic state. We show that theClifford group operations combined with magic states preparation are sufficient for UQC. The connection <strong>of</strong>our results with the Gottesman-Knill theorem is discussed.DOI: 10.1103/PhysRevA.71.022316PACS numbers: 03.67.Lx, 03.67.PpI. INTRODUCTION AND SUMMARYThe theory <strong>of</strong> fault-tolerant quantum computation definesan important number called the error threshold. If the physicalerror rate is less than the threshold value , it is possibleto stabilize computation by transforming the quantum circuitinto a fault-tolerant form where errors can be detected <strong>and</strong>eliminated. However, if the error rate is above the threshold,then errors begin to accumulate, which results in rapid decoherence<strong>and</strong> renders the output <strong>of</strong> the computation useless.The actual value <strong>of</strong> depends on the error correction scheme<strong>and</strong> the error model. Unfortunately, this number seems to berather small for all known schemes. Estimates vary from10 −6 see Ref. 1 to 10 −4 see Refs. 2–4, which is hardlyachievable with the present technology.In principle, one can envision a situation in which qubitsdo not decohere, <strong>and</strong> a subset <strong>of</strong> the elementary gates isrealized exactly due to special properties <strong>of</strong> the physical system.This scenario could be realized experimentally usingspin, electron, or other many-body systems with topologicallyordered ground states. Excitations in two-dimensionaltopologically ordered systems are anyons—quasiparticleswith unusual statistics described by nontrivial representations<strong>of</strong> the braid group. If we have sufficient control <strong>of</strong>anyons, i.e., are able to move them around each other, fusethem, <strong>and</strong> distinguish between different particle types, thenwe can realize some set <strong>of</strong> unitary operators <strong>and</strong> measurementsexactly. This set may or may not be computationallyuniversal. While the universality can be achieved with sufficientlynontrivial types <strong>of</strong> anyons 5–8, more realistic systems<strong>of</strong>fer only decoherence protection <strong>and</strong> an incomplete set<strong>of</strong> topological gates. See Refs. 9,10 about non-Abeliananyons in quantum Hall systems <strong>and</strong> Refs. 11,12 abouttopological orders in Josephson junction arrays. Nevertheless,universal computation is possible if we introduce some*Email address: serg@cs.caltech.edu† Email address: kitaev@iqi.caltech.eduadditional operations e.g., measurements by Aharonov-Bohm interference 13 or some gates that are not related totopology at all. Of course, these nontopological operationscannot be implemented exactly <strong>and</strong> thus are prone to errors.In this situation, the threshold error rate may becomesignificantly larger than the values given above because weneed to correct only errors <strong>of</strong> certain special type <strong>and</strong> weintroduce a smaller amount <strong>of</strong> error in the correction stage.The main purpose <strong>of</strong> the present paper is to illustrate thisstatement by a particular computational model.The model is built upon the Clifford group—the group <strong>of</strong>unitary operators that map the group <strong>of</strong> Pauli operators toitself under conjugation. The set <strong>of</strong> elementary operations isdivided into two parts: O=O ideal O faulty . Operations fromO ideal are assumed to be perfect. We list these operationsbelow:i prepare a qubit in the state 0;ii apply unitary operators from the Clifford group;iii measure an eigenvalue <strong>of</strong> a Pauli operator x , y ,or z on any qubit.Here we mean nondestructive projective measurement.We also assume that no errors occur between the operations.It is well known that these operations are not sufficient foruniversal quantum computation UQC unless a quantumcomputer can be efficiently simulated on a classical computer.More specifically, the Gottesman-Knill theorem statesthat by operations from O ideal one can only obtain quantumstates <strong>of</strong> a very special form called stabilizer states. Such astate can be specified as an intersection <strong>of</strong> eigenspaces <strong>of</strong>pairwise commuting Pauli operators, which are referred to asstabilizers. Using the stabilizer formalism, one can easilysimulate the evolution <strong>of</strong> the state <strong>and</strong> the statistics <strong>of</strong> measurementson a classical probabilistic computer see Ref.14 or a textbook 15 for more details.The set O faulty describes faulty operations. In our model, itconsists <strong>of</strong> just one operation: prepare an ancillary qubit in amixed state . The state should be regarded as a parameter<strong>of</strong> the model. From the physical point <strong>of</strong> view, is mixeddue to imperfections <strong>of</strong> the preparation procedure entanglement<strong>of</strong> the ancilla with the environment, thermal fluctua-1050-2947/2005/712/02231614/$23.00022316-1©2005 The American Physical Society


S. BRAVYI AND A. KITAEV PHYSICAL REVIEW A 71, 022316 2005tions, etc.. An essential requirement is that by preparing nqubits we obtain the state n , i.e., all ancillary qubits areindependent. The independence assumption is similar to theuncorrelated errors model in the st<strong>and</strong>ard fault-tolerant computationtheory.Our motivation for including all Clifford group gates intoO ideal relies mostly on the recent progress in the fault-tolerantimplementation <strong>of</strong> such gates. For instance, using a concatenatedstabilizer code with good error correcting propertiesto encode each qubit <strong>and</strong> applying gates transversally so thaterrors do not propagate inside code blocks one can implementClifford gates with an arbitrary high precision, see Ref.16. However, these nearly perfect gates act on encodedqubits. To establish a correspondence with our model, oneneeds to prepare an encoded ancilla in the state . It can bedone using the schemes for fault-tolerant encoding <strong>of</strong> an arbitraryknown one-qubit state described by Knill in Ref. 17.In the more recent paper 18 Knill constructed a scheme <strong>of</strong>fault-tolerant quantum computation which combines i theteleported computing <strong>and</strong> error correction technique by Gottesman<strong>and</strong> Chuang 19; ii the method <strong>of</strong> purification <strong>of</strong>CSS states by Dür <strong>and</strong> Briegel 20; <strong>and</strong> iii the magic statesdistillation algorithms described in the present paper. As wasargued in Ref. 18, this scheme is likely to yield a muchhigher value for the threshold it may be up to 1%.Unfortunately, ideal implementation <strong>of</strong> the Clifford groupcannot be currently achieved in any realistic physical systemwith a topological order. What universality classes <strong>of</strong> anyonsallow one to implement all Clifford group gates but do notallow one to simulate UQC is an interesting open problem.To fully utilize the potential <strong>of</strong> our model, we allow adaptivecomputation. It means that a description <strong>of</strong> an operationto be performed at step t may be a function <strong>of</strong> all measurementoutcomes at steps 1,…,t−1. For even greater generality,the dependence may be probabilistic. This assumptiondoes not actually strengthen the model since tossing a faircoin can be simulated using O ideal At this point, we need tobe careful because the proper choice <strong>of</strong> operations should notonly be defined mathematically—it should be computed bysome efficient algorithm. In all protocols described below,the algorithms will actually be very simple. Let us point outthat dropping the computational complexity restriction stillleaves a nontrivial problem: can we prepare an arbitrary multiqubitpure state with any given fidelity using only operationsfrom the basis O?The main question that we address in this paper is asfollows: For which density matrices can one efficientlysimulate universal quantum computation by adaptive computationin the basis O?It will be convenient to use the Bloch sphere representation<strong>of</strong> one-qubit states: = 1 2 I + x x + y y + z z .The vector x , y , z will be referred to as the polarizationvector <strong>of</strong> . Let us first consider the subset <strong>of</strong> states satisfying x + y + z 1.This inequality says that the vector x , y , z lies inside theoctahedron O with vertices ±1, 0, 0, 0, ±1, 0, 0, 0, ±1,FIG. 1. Left: the Bloch sphere <strong>and</strong> the octahedron O. Right: theoctahedron O projected on the x−y plane. The magic states correspondto the intersections <strong>of</strong> the symmetry axes <strong>of</strong> O with the Blochsphere. The empty <strong>and</strong> filled circles represent T-type <strong>and</strong> H-typemagic states, respectively.see Fig. 1. The six vertices <strong>of</strong> O represent the six eigenstates<strong>of</strong> the Pauli operators x , y , <strong>and</strong> z . We can prepare thesestates by operations from O ideal only. Since is a convexlinear combination probabilistic mixture <strong>of</strong> these states, wecan prepare by operations from O ideal <strong>and</strong> by tossing a coinwith suitable weights. Thus we can rephrase the Gottesman-Knill theorem in the following way.Theorem 1. Suppose the polarization vector x , y , z <strong>of</strong>the state belongs to the convex hull <strong>of</strong> ±1, 0, 0, 0, ±1, 0,0, 0, ±1. Then any adaptive computation in the basis O canbe efficiently simulated on a classical probabilistic computer.This observation leads naturally to the following question:is it true that UQC can be efficiently simulated whenever lies in the exterior <strong>of</strong> the octahedron O? In an attempt toprovide at least a partial answer, we prove the universalityfor a large set <strong>of</strong> states. Specifically, we construct two particularschemes <strong>of</strong> UQC simulation based on a method whichwe call magic states distillation. Let us start by defining themagic states.Definition 1. Consider pure states H,TC 2 such that<strong>and</strong>TT = 1 2I + 1 3 x + y + z ,HH = 1 2I + 1 2 x + z .The images <strong>of</strong> T <strong>and</strong> H under the action <strong>of</strong> one-qubitClifford operators are called magic states <strong>of</strong> T type <strong>and</strong> Htype, respectively.This notation is chosen since H <strong>and</strong> T are eigenvectors<strong>of</strong> certain Clifford group operators: the Hadamard gate H <strong>and</strong>the operator usually denoted T, see Eq. 7. Denote the onequbitClifford group by C 1 . Overall, there are 8 magic states<strong>of</strong> T type, UT,UC 1 up to a phase <strong>and</strong> 12 states <strong>of</strong> Htype, UH,UC 1 , see Fig. 1. Clearly, the polarization vectors<strong>of</strong> magic states are in one-to-one correspondence withrotational symmetry axes <strong>of</strong> the octahedron O H-type statescorrespond to 180° rotations <strong>and</strong> T-type states correspond to120° rotations. The role <strong>of</strong> magic states in our constructionis tw<strong>of</strong>old. First, adaptive computation in the basis O idealtogether with the preparation <strong>of</strong> magic states <strong>of</strong> either typeallows one to simulate UQC see Sec. III. Second, by adap-022316-2


UNIVERSAL QUANTUM COMPUTATION WITH IDEAL…tive computation in the basis O ideal one can “purify” imperfectmagic states. It is a rather surprising coincidence thatone <strong>and</strong> the same state can comprise both <strong>of</strong> these properties,<strong>and</strong> that is the reason why we call them magic states.More exactly, a magic state distillation procedure yieldsone copy <strong>of</strong> a magic state with any desired fidelity fromseveral copies <strong>of</strong> the state , provided that the initial fidelitybetween <strong>and</strong> the magic state to be distilled is large enough.In the course <strong>of</strong> distillation, we use only operations from theset O ideal . By constructing two particular distillationschemes, for T-type <strong>and</strong> H-type magic states, respectively,we prove the following theorems.Theorem 2. Let F T be the maximum fidelity between <strong>and</strong> a T-type magic state, i.e.,F T = maxUC 1 TU † UT.Adaptive computation in the basis O=O ideal allows oneto simulate universal quantum computation wheneverF T F T = +21 1 3 0.910.71/2Theorem 3. Let F H be the maximum fidelity between <strong>and</strong> an H-type magic state,F H = maxUC 1 HU † UH.Adaptive computation in the basis O=O ideal allows oneto simulate universal quantum computation wheneverF H F H 0.927.The quantities F T <strong>and</strong> F H have the meaning <strong>of</strong> thresholdfidelity since our distillation schemes increase the polarization<strong>of</strong> , converging to a magic state as long as the inequalitiesF T F T or F H F H are fulfilled. If they are notfulfilled, the process converges to the maximally mixed state.The conditions stated in the theorems can also be understoodin terms <strong>of</strong> the polarization vector x , y , z . Indeed, let usassociate a “magic direction” with each <strong>of</strong> the magic states.Then Theorems 2 <strong>and</strong> 3 say that the distillation is possible ifthere is a T direction such that the projection <strong>of</strong> the vector x , y , z onto that T direction exceeds the threshold value<strong>of</strong> 2F T 2 −10.655, or if the projection on some <strong>of</strong> the Hdirections is greater than 2F H 2 −10.718.Let us remark that, although the proposed distillationschemes are probably not optimal, the threshold fidelities F T<strong>and</strong> F H cannot be improved significantly. Indeed, it is easy tocheck that the octahedron O corresponding to probabilisticmixtures <strong>of</strong> stabilizer states can be defined aswhereF T * =O = :F T F T * ,121 + 1 31/2 0.888.It means that F T * is a lower bound on the threshold fidelity F Tfor any protocol distilling T-type magic states. Thus any potentialimprovement to Theorem 2 may only decrease F Tfrom 0.910 down to F T * =0.888. From a practical perspective,the difference between these two numbers is not important.On the other h<strong>and</strong>, such an improvement would be <strong>of</strong>great theoretical interest. Indeed, if Theorem 2 with F T replacedby F T * is true, it would imply that the Gottesman-Knilltheorem provides necessary <strong>and</strong> sufficient conditions for theclassical simulation, <strong>and</strong> that a transition from classical touniversal quantum behavior occurs at the boundary <strong>of</strong> theoctahedron O. This kind <strong>of</strong> transition has been discussed incontext <strong>of</strong> a general error model 21. Our model is simpler,which gives hope for sharper results.By the same argument, one can show that the quantitydefF * H =maxOPHYSICAL REVIEW A 71, 022316 2005 HH = 1 21 + 1 21/2 0.924is a lower bound on the threshold fidelity F H for any protocoldistilling H-type magic states.A similar approach to UQC simulation was suggested inRef. 22, where Clifford group operations were used to distillthe entangled three-qubit state 000+001+010+100, which is necessary for the realization <strong>of</strong> the T<strong>of</strong>foligate.The rest <strong>of</strong> the paper is organized as follows. Section IIcontains some well-known facts about the Clifford group <strong>and</strong>stabilizer formalism, which will be used throughout the paper.In Sec. III we prove that magic states together withoperations from O ideal are sufficient for UQC. In Sec. IVideal magic are substituted by faulty ones <strong>and</strong> the error ratethat our simulation algorithm can tolerate is estimated. InSec. V we describe a distillation protocol for T-type magicstates. This protocol is based on the well-known five-qubitquantum code. In Sec. VI a distillation protocol for H-typemagic states is constructed. It is based on a certain CSSstabilizer code that encodes one qubit into 15 <strong>and</strong> admits anontrivial automorphism 23. Specifically, the bitwise application<strong>of</strong> a certain non-Clifford unitary operator preserves thecode subspace <strong>and</strong> effects the same operator on the encodedqubit. We conclude with a brief summary <strong>and</strong> a discussion <strong>of</strong>open problems.II. CLIFFORD GROUP, STABILIZERS, AND SYNDROMEMEASUREMENTSLet C n denote the n-qubit Clifford group. Recall that it is afinite subgroup <strong>of</strong> U2 n generated by the Hadamard gate Happlied to any qubit, the phase-shift gate K applied to anyqubit, <strong>and</strong> the controlled-not gate x which may be appliedto any pair qubits,H = 1 2 1 11 − 1, K = 1 00 i, x = I 00 x.The Pauli operators x , y , z belong to C 1 , for instance, z=K 2 <strong>and</strong> x =HK 2 H. The Pauli group PnC n is generatedby the Pauli operators acting on n qubits. It is known 24that the Clifford group C n augmented by scalar unitary operatorse i I coincides with the normalizer <strong>of</strong> Pn in the uni-1022316-3


S. BRAVYI AND A. KITAEV PHYSICAL REVIEW A 71, 022316 2005tary group U2 n . Hermitian elements <strong>of</strong> the Pauli group are<strong>of</strong> particular importance for quantum error correction theory;they are referred to as stabilizers. These are operators <strong>of</strong> theform± 1 ¯ n, j 0,x,y,z,where 0 =I. Let us denote by Sn the set <strong>of</strong> all n-qubitstabilizers:Sn = S Pn : S † = S.For any two stabilizers S 1 ,S 2 we have S 1 S 2 = ±S 2 S 1 <strong>and</strong> S 12=S 2 2 =I. It is known that for any set <strong>of</strong> pairwise commutingstabilizers S 1 ,…,S k Sn there exists a unitary operator VC n such thatVS j V † = z j,j = 1,…,k,where z j denotes the operator z applied to the jth qubit,e.g., z 1= z I ¯ I.These properties <strong>of</strong> the Clifford group allow us to introducea very useful computational procedure which can berealized by operations from O ideal . Specifically, we can performa joint nondestructive eigenvalue measurement for anyset <strong>of</strong> pairwise commuting stabilizers S 1 ,…,S k Sn. Theoutcome <strong>of</strong> such a measurement is a sequence <strong>of</strong> eigenvalues= 1 ,…, k , j = ±1, which is usually called a syndrome.For any given outcome, the quantum state is acted upon bythe projectork1 = j=1 2 I + jS j .Now, let us consider a computation that begins with anarbitrary state <strong>and</strong> consists <strong>of</strong> operations from O ideal . It isclear that we can defer all Clifford operations until the veryend if we replace the Pauli measurements by general syndromemeasurements. Thus the most general transformationthat can be realized by O ideal is an adaptive syndrome measurement,meaning that the choice <strong>of</strong> the stabilizer S j to bemeasured next depends on the previously measured values <strong>of</strong> 1 ,…, j−1 . In general, this dependence may involve cointossing. Without loss <strong>of</strong> generality one can assume that S jcommutes with all previously measured stabilizersS 1 ,…,S j−1 for all possible values <strong>of</strong> 1 ,…, j−1 <strong>and</strong> cointossing outcomes. Adaptive syndrome measurement hasbeen used in Ref. 25 to distill entangled states <strong>of</strong> a bipartitesystem by local operations.III. UNIVERSAL QUANTUM COMPUTATION WITHMAGIC STATESIn this section, we show that operations from O ideal aresufficient for universal quantum computation if a supply <strong>of</strong>ideal magic states is also available. First, consider a onequbitstateA = 2 −1/2 0 + e i 1<strong>and</strong> suppose that is not a multiple <strong>of</strong> /2. We now describea procedure that implements the phase shift gate2022316-4e i = 1 00 e i by consuming several copies <strong>of</strong> A <strong>and</strong> using only operationsfrom O ideal .Let =a0+b1 be the unknown initial state whichshould be acted on by e i . Prepare the state 0 = A <strong>and</strong> measure the stabilizer S 1 = z z . Note that bothoutcomes <strong>of</strong> this measurement appear with probability 1/2.If the outcome is “+1”, we are left with the state 1 + = a0,0 + be i 1,1.In the case <strong>of</strong> “−1” outcome, the resulting state is 1 − = ae i 0,1 + b1,0.Let us apply the gate x 1,2 the first qubit is the controlone. The above two states are mapped to 2 + = x 1,2 1 + = a0 + be i 1 0, 2 − = x 1,2 1 − = ae i 0 + b1 1.Now the second qubit can be discarded, <strong>and</strong> we are left withthe state a0+be ±i 1, depending upon the measured eigenvalue.Thus the net effect <strong>of</strong> this circuit is the application <strong>of</strong>a unitary operator that is chosen r<strong>and</strong>omly between e i <strong>and</strong> e −i <strong>and</strong> we know which <strong>of</strong> the two possibilities hasoccurred.Applying the circuit repeatedly, we effect the transformationse ip 1 , e ip 2 ,… for some integers p 1 , p 2 ,… whichobey the r<strong>and</strong>om-walk statistics. It is well known that such ar<strong>and</strong>om walk visits each integer with the probability 1. Itmeans that sooner or later we will get p k =1 <strong>and</strong> thus realizethe desired operator e i . The probability that we will needmore than N steps to succeed can be estimated as cN −1/2 forsome constant c0. Note also that if is a rational multiple<strong>of</strong> 2, we actually have a r<strong>and</strong>om walk on a cyclic group Z q .In this case, the probability that we will need more than Nsteps decreases exponentially with N.The magic state H can be explicitly written in the st<strong>and</strong>ardbasis asH = cos 80 + sin 81.Note that HKH=e i/8 A −/4 . So if we are able to preparethe state H, we can realize the operator e −i/4 . It doesnot belong to the Clifford group. Moreover, the subgroup <strong>of</strong>U2 generated by e −i/4 <strong>and</strong> C 1 is dense in U2. 1 Thusthe operators from C 1 <strong>and</strong> C 2 together with e −i/4 constitutea universal basis for quantum computation.The magic state T can be explicitly written in the st<strong>and</strong>ardbasis:1 Recall that the action <strong>of</strong> the Clifford group C 1 on the set <strong>of</strong>operators ± x , ± y , ± z coincides with the action <strong>of</strong> rotational symmetrygroup <strong>of</strong> a cube on the set <strong>of</strong> unit vectors ±e x , ±e y , ±e z ,respectively.3


UNIVERSAL QUANTUM COMPUTATION WITH IDEAL…T = cos 0 + e i/4 sin 1, cos2 = 1 3.Let us prepare an initial state 0 =T T <strong>and</strong> measure thestabilizer S 1 = z z . The outcome +1 appears with probabilityp + =cos 4 +sin 4 =2/3. If the outcome is −1, we discardthe reduced state <strong>and</strong> try again, using a fresh pair <strong>of</strong>magic states. On average, we need three copies <strong>of</strong> the Tstate to get the outcome +1. The reduced state correspondingto the outcome +1 is 1 = cos 0,0 + i sin 1,1, = 12 .Let us apply the gate x 1,2 <strong>and</strong> discard the secondqubit. We arrive at the state 2 = cos 0 + i sin 1.Next apply the Hadamard gate H: 3 = H 2 = 2 −1/2 e i 0 + e −2i 1 = A −/6 .We can use this state as described above to realize the operatore −i/6 . It is easy to check that Clifford operatorstogether with e −i/6 constitute a universal set <strong>of</strong> unitarygates.Thus we have proved that the sets <strong>of</strong> operationsO ideal H <strong>and</strong> O ideal T are sufficient for universalquantum computation.IV. ERROR ANALYSISTo establish a connection between the simulation algorithmsdescribed in Sec. III <strong>and</strong> the universality theoremsstated in the introduction we have to substitute ideal magicstates by faulty ones. Before doing that let us discuss theideal case in more detail. Suppose that a quantum circuit tobe simulated uses a gate basis in which the only non-Cliffordgate is the phase shift e −i/4 or e −i/6 . One can applythe algorithm <strong>of</strong> Sec. III to simulate each non-Clifford gateindependently. To avoid fluctuations in the number <strong>of</strong> magicstates consumed at each round, let us set a limit <strong>of</strong> K magicstates per round, where K is a parameter to be chosen later.As was pointed out in Sec. III, the probability for some particularsimulation round to “run out <strong>of</strong> budget” scales asexp−K for some constant 0. If at least one simulationround runs out <strong>of</strong> budget, we declare a failure <strong>and</strong> the wholesimulation must be aborted. Denote the total number <strong>of</strong> non-Clifford gates in the circuit by L. The probability p a for thewhole simulation to be aborted can be estimated asp a 1 − 1 − exp− K L L exp− K 1,provided that L exp−K1. We will assumeK −1 ln L,so the abort probability can be neglected.Each time the algorithm requests an ideal magic state, itactually receives a slightly nonideal one. Such nearly perfectmagic states must be prepared using the distillation methods4described in Secs. V <strong>and</strong> VI. Let us estimate an affordableerror rate out for distilled magic states. Since there are Lnon-Clifford gates in the circuit, one can tolerate an errorrate <strong>of</strong> the order 1/L in implementation <strong>of</strong> these gates. 2 Eachnon-Clifford gate requires Kln L magic states. Thus thewhole simulation is reliable enough if one chooses out 1/L ln L.What are the resources needed to distill one copy <strong>of</strong> amagic state with the error rate out ? To be more specific, letus talk about H-type states. It will be shown in Sec. VI thatthe number n <strong>of</strong> raw undistilled ancillas needed to distillone copy <strong>of</strong> the H magic state with an error rate not exceeding out scales asn ln1/ out , = log 3 15 2.5,see Eq. 39. Taking out from Eq. 5, one getsn ln L .Since the whole simulation requires KLL ln L copies <strong>of</strong>the distilled H state, we needN Lln L +1raw ancillas overall.Summarizing, the simulation theorems stated in the introductionfollow from the following results the last one willbe proved later:i the circuits described in Sec. III allow one to simulateUQC with the sets <strong>of</strong> operations O ideal H <strong>and</strong>O ideal T;ii these circuits work reliably enough if the states H<strong>and</strong> T are slightly noisy, provided that the error rate doesnot exceed out 1/L ln L;iii a magic state having an error rate out can be preparedfrom copies <strong>of</strong> the raw ancillary state using the distillationschemes provided that F T F T or F H F H .The distillation requires resources that are polynomial inln L.V. DISTILLATION OF T-TYPE MAGIC STATESSuppose we are given n copies <strong>of</strong> a state , <strong>and</strong> our goalis to distill one copy <strong>of</strong> the magic state T. The polarizationvector <strong>of</strong> can be brought into the positive octant <strong>of</strong> theBloch space by a Clifford group operator, so we can assumethat x , y , z 0.In this case, the fidelity between <strong>and</strong> T is the largest oneamong all T-type magic states, i.e.,A related quantity,PHYSICAL REVIEW A 71, 022316 2005F T = TT.2 This fault tolerance does not require any redundancy in theimplementation <strong>of</strong> the circuit e.g., the use <strong>of</strong> concatenated codes.It is achived automatically because in the worst case the error probabilityaccumulates linearly in the number <strong>of</strong> gates. In our modelonly non-Clifford gates are faulty.5022316-5


S. BRAVYI AND A. KITAEV PHYSICAL REVIEW A 71, 022316 2005 = 1 − TT = 1 21 − 1 3 x + y + z ,will be called the initial error probability. By definition, 01/2.The output <strong>of</strong> the distillation algorithm will be some onequbitmixed state out . To quantify the proximity between out<strong>and</strong> T, let us define a final error probability: out = 1 − T out T.It will be certain function <strong>of</strong> n <strong>and</strong> . The asymptotic behavior<strong>of</strong> this function for n→ reveals the existence <strong>of</strong> athreshold error probability, 0 = 1 21 − 3 7 0.173,TT 0 = e +i/3 T 0 , TT 1 = e −i/3 T 1 ,T 0,1 T 0,1 = 1 2I ± 1 3 x + y + z .defNote that T 0 = T <strong>and</strong> T 1 = y HT 0 are T-type magicstates.Let us apply a dephasing transformation,D = 1 3 + TT† + T † T,to each copy <strong>of</strong> the state . The transformation D can berealized by applying one <strong>of</strong> the operators I,T,T −1 chosenwith probability 1/3 each. Since9such that for 0 the function out n, converges to zero.We will see that for small ,we haveDT 0 T 1 = DT 1 T 0 = 0, out n, 5 n , = 1/log 2 30 0.2. 6D = 1 − T 0 T 0 + T 1 T 1 .10On the other h<strong>and</strong>, if 0 , the output state converges to themaximally mixed state, i.e., lim n→ out n,=1/2.Before coming to a detailed description <strong>of</strong> the distillationalgorithm, let us outline the basic ideas involved in its construction.The algorithm recursively iterates an elementarydistillation subroutine that transforms five copies <strong>of</strong> an imperfectmagic state into one copy having a smaller errorprobability. This elementary subroutine involves a syndromemeasurement for certain commuting stabilizers S 1 ,S 2 ,S 3 ,S 4S5. If the measured syndrome 1 , 2 , 3 , 4 is nontrivial j =−1 for some j, the distillation attempt fails <strong>and</strong>the reduced state is discarded. If the measured syndrome istrivial j =1 for all j, the distillation attempt is successful.Applying a decoding transformation a certain Clifford operatorto the reduced state, we transform it to a single-qubitstate. This qubit is the output <strong>of</strong> the subroutine.Our construction is similar to concatenated codes used inmany fault-tolerant quantum computation techniques, but itdiffers from them in two respects. First, we do not need tocorrect errors—it suffices only to detect them. Once an errorhas been detected, we simply discard the reduced state, sinceit does not contain any valuable information. This allows usto achieve higher threshold error probability. Second, we donot use quantum codes in the way for which they were originallydesigned: in our scheme, the syndrome is measured ona product state.The state T is an eigenstate for the unitary operatorT = e i/4 KH = ei/4 2 1 1i − i C 1 . 7Note that T acts on the Pauli operators as follows: 3T x T † = z , T z T † = y , T y T † = x . 8We will denote its eigenstates by T 0 <strong>and</strong> T 1 , so thatWe will assume that the dephasing transformation is appliedat the very first step <strong>of</strong> the distillation, so has the form 10.Thus the initial state for the elementary distillation subroutineis in = 5 =x0,1 5 x 1 − 5−x T x T x ,11where x=x 1 ,…,x 5 is a binary string, x is the number <strong>of</strong>1’s in x, <strong>and</strong>defT x = T x1 ¯ T x5 .The stabilizers S 1 ,…,S 4 to be measured on the state incorrespond to the famous five-qubit code, see Refs. 26,27.They are defined as follows:S 1 = x z z x I,S 2 = I x z z x ,S 3 = x I x z z ,S 4 = z x I x z .12This code has a cyclic symmetry, which becomes explicit ifwe introduce an auxiliary stabilizer, S 5 =S 1 S 2 S 3 S 4 = z z x I x . Let L be the two-dimensional code subspacespecified by the conditions S j =, j=1,…, 4, <strong>and</strong> bethe orthogonal projector onto L:4 = 1 I + S j .16 j=1It was pointed out in Ref. 16 that the operators133 The operator denoted by T in Ref. 16 does not coincide withour T. They are related by the substitution T→e −i/4 T † though.<strong>and</strong>Xˆ = x 5 , Ŷ = y 5 , Ẑ = z 5 ,022316-6


UNIVERSAL QUANTUM COMPUTATION WITH IDEAL…PHYSICAL REVIEW A 71, 022316 2005Tˆ = T 514commute with , thus preserving the code subspace. Moreover,Xˆ ,Ŷ ,Ẑ obey the same algebraic relations as one-qubitPauli operators, e.g., Xˆ Ŷ =iẐ. Let us choose a basis in L suchthat Xˆ ,Ŷ, <strong>and</strong> Ẑ become logical Pauli operators x , y , <strong>and</strong> z , respectively. How does the operator Tˆ act in this basis?From Eq. 8 we immediately getTˆ Xˆ Tˆ † = Ẑ, Tˆ ẐTˆ † = Ŷ, Tˆ ŶTˆ † = Xˆ .Therefore Tˆ coincides with the logical operator T up to anoverall phase factor. This factor is fixed by the condition thatthe logical T has eigenvalues e ±i/3 .Let us find the eigenvectors <strong>of</strong> Tˆ that belong to L. Considertwo particular states from L, namelyT 1 L = 6T00000 , <strong>and</strong> T 0 L = 6T11111 .In the Appendix we show thatT 00000 T 00000 = T 11111 T 11111 = 1 6 ,15so that the states T 0 L <strong>and</strong> T 1 L are normalized. Taking intoaccount that Tˆ ,=0 <strong>and</strong> thatwe getTˆ T x = e i/35−2x T x for all x 0,1 5 ,Tˆ T 1 L = 6Tˆ T00000 = 6Tˆ T00000 = e −i/3 T 1 L .Analogously, one can check thatTˆ T 0 L = e +i/3 T 0 L .16It follows that Tˆ is exactly the logical operator T, includingthe overall phase, <strong>and</strong> T 0 L <strong>and</strong> T 1 L are the logical states T 0 <strong>and</strong> T 1 up to some phase factors, which are not importantfor us. Therefore we haveT L 0,1 T L 0,1 = 2I 1 ± 1 Xˆ + Ŷ + Ẑ. 317Now we are in a position to describe the syndrome measurementperformed on the state in . The unnormalized reducedstate corresponding to the trivial syndrome is as follows: s = in =x0,1 5 x 1 − 5−x T x T x ,18see Eq. 11. The probability for the trivial syndrome to beobserved isp s = Tr s .Note that the state T x is an eigenvector <strong>of</strong> Tˆ for any x0,1 5 . But we know that the restriction <strong>of</strong> Tˆ on L haseigenvalues e ±i/3 . At the same time, Eq. 16 implies thatTˆ T x = − T x whenever x=1 or x=4. This eigenvalue equation is not acontradiction only ifT x = 0 for x = 1,4.This equality can be interpreted as an error correction property.Indeed, the initial state in is a mixture <strong>of</strong> the desiredstate T 00000 <strong>and</strong> unwanted states T x with x0. We caninterpret the number <strong>of</strong> “1” components in x as a number <strong>of</strong>errors. Once the trivial syndrome has been measured, we canbe sure that either no errors or at least two errors have occurred.Such error correction, however, is not directly relatedto the minimal distance <strong>of</strong> the code.It follows from Eq. 16 that for x=2, 3 one hasTˆ T x =e ±i/3 T x , so that T x must be proportional toone <strong>of</strong> the states T 0 L , T 1 L . Our observations can be summarizedas follows:T x =6−1/2 T L 1 , if x = 0,0, if x = 1,a x T L 0 , if x = 2,b x T L 191 , if x = 3,0, if x = 4,6 −1/2 T L 0 , if x = 5.Here the coefficients a x ,b x depend upon x in some way. Theoutput state 18 can now be written as s =2 1 6 5 + 2 1 − 3 a x T 0 L T L 0 x:x=2+ 1 6 1 − 5 + 3 1 − 2 x:x=3b x 2 T 1 L T 1 L . 20To exclude the unknown coefficients a x <strong>and</strong> b x , we can usethe identityT 0 L T 0 L + T 1 L T 1 L = =x0,1 5 T x T x .Substituting Eq. 19 into this identity, we get a x 2 = b x 2 = 5x:x=2 x:x=36 .So the final expression for the output state s is as follows: s = 5 + 5 2 1 − 3T L6 0 T L 0 + 1 − 5 + 5 3 1 − 26T 1 L T 1 L .21Accordingly, the probability to observe the trivial syndromeisp s = 5 + 5 2 1 − 3 + 5 3 1 − 2 + 1 − 5. 226A decoding transformaion for the five-qubit code is a unitaryoperator VC 5 such that022316-7


S. BRAVYI AND A. KITAEV PHYSICAL REVIEW A 71, 022316 2005FIG. 2. The final error probability out <strong>and</strong> the probability p s tomeasure the trivial syndrome as functions <strong>of</strong> the initial error probability for the T-type states distillation.VL = C 2 0,0,0,0.In other words, V maps the stabilizers S j , j=2, 3, 4, 5 to z j. The logical operators Xˆ ,Ŷ ,Ẑ are mapped to the Paulioperators x , y , z acting on the first qubit. From Eq. 17we infer thatVT L0,1 = T 0,1 0,0,0,0maybe up to some phase. The decoding should be followedby an additional operator A= y HC 1 , which swaps thestates T 0 <strong>and</strong> T 1 note that for small the state s is closeto T 1 L , while our goal is to distill T 0 . After that we get anormalized output statewhere out = 1 − out T 0 T 0 + out T 1 T 1 , out =t 5 + 5t 21 + 5t 2 + 5t 3 + t 5, t = 1 − . 23The plot <strong>of</strong> the function out is shown on Fig. 2. Itindicates that the equation out = has only one nontrivialsolution, = 0 0.173. The exact value is 0 = 1 21 − 3 7.If 0 , we can recursively iterate the elementary distillationsubroutine to produce as good an approximation to thestate T 0 as we wish. On the other h<strong>and</strong>, if 0 , the distillationsubroutine increases the error probability <strong>and</strong> iterationsconverge to the maximally mixed state. Thus 0 is athreshold error probability for our scheme. The correspondingthreshold polarization is 1−2 0 = 3/70.655. For a sufficientlysmall , one can use the approximation out 5 2 .The probability p s = p s to measure the trivial syndromedecreases monotonically from 1/6 for =0 to 1/16 for =1/2, see Fig. 2. In the asymptotic regime where is small,we can use the approximation p s p s 0=1/6.Now the construction <strong>of</strong> the whole distillation scheme isstraightforward. We start from n1 copies <strong>of</strong> the state =1−T 0 T 0 +T 1 T 1 . Let us split these states intogroups containing five states each <strong>and</strong> apply the elementarydistillation subroutine described above to each group independently.In some <strong>of</strong> these groups the distillation attemptfails, <strong>and</strong> the outputs <strong>of</strong> such groups must be discarded. Theaverage number <strong>of</strong> “successful” groups is obviously p s n/5n/30 if is small. Neglecting the fluctuations <strong>of</strong>this quantity, we can say that our scheme provides a constantyield r=1/30 <strong>of</strong> output states that are characterized by theerror probability out 5 2 . Therefore we can obtain r 2 nstates with out 5 3 4 , r 3 n states with out 5 7 8 , <strong>and</strong> so on.We have created a hierarchy <strong>of</strong> states with n states on thefirst level <strong>and</strong> four or fewer states on the last level. Let k bethe number <strong>of</strong> levels in this hierarchy <strong>and</strong> out the error probabilitycharacterizing the states on the last level. Up to smallfluctuations, the numbers n,k, out , <strong>and</strong> are related by thefollowing obvious equations:Their solution yields Eq. 6. out 1 5 52k , r k n 1. 24VI. DISTILLATION OF H-TYPE MAGIC STATESA distillation scheme for H-type magic states also worksby recursive iteration <strong>of</strong> a certain elementary distillation subroutinebased on a syndrome measurement for a suitable stabilizercode. Let us start with introducing some relevant codingtheory constructions, which reveal an unusual symmetry<strong>of</strong> this code <strong>and</strong> explain why it is particularly useful forH-type magic states distillation.Let F 2 n be the n-dimensional binary linear space <strong>and</strong> A bea one-qubit operator such that A 2 =I. With any binary vectoru=u 1 ,…,u n F 2 n we associate the n-qubit operatorAu = A u 1 A u 2 ¯ A u n.Let u,v= ni=1 u i v i mod 2 denote the st<strong>and</strong>ard binary innerproduct. If LF n 2 is a linear subspace, we denote by L theset <strong>of</strong> vectors which are orthogonal to L. The Hammingweight <strong>of</strong> a binary vector u is denoted by u. Finally, u·vF n 2 designates the bitwise product <strong>of</strong> u <strong>and</strong> v, i.e., u·v i=u i v i .A systematic way <strong>of</strong> constructing stabilizer codes wassuggested by Calderbank, Shor, <strong>and</strong> Steane, see Refs.28,29. Codes that can be described in this way will bereferred to as st<strong>and</strong>ard CSS codes. In addition, we consider022316-8


UNIVERSAL QUANTUM COMPUTATION WITH IDEAL…their images under an arbitrary unitary transformation VU2 applied to every qubit. Such “rotated” codes will becalled CSS codes.Definition 2. Consider a pair <strong>of</strong> one-qubit Hermitian operatorsA,B such thatA 2 = B 2 = I,AB = − BA,<strong>and</strong> a pair <strong>of</strong> binary vector spaces L A ,L B F 2 n , such thatu,v = 0 for all u L A ,v L B .A quantum code CSSA,L A ;B,L B is a decompositionC 2 n = H,,* *L A L Bwhere the subspace H, is defined by the conditionsAu = − 1 u ,Bv = − 1 v 25for all uL A <strong>and</strong> vL B . The linear functionals <strong>and</strong> arereferred to as A syndrome <strong>and</strong> B syndrome, respectively. Thesubspace H0,0 corresponding to the trivial syndromes ==0 is called the code subspace.The subspaces H, are well defined since the operatorsAu <strong>and</strong> Bv commute for any uL A <strong>and</strong> vL B :AuBv = − 1 u,v BvAu = BvAu.The number <strong>of</strong> logical qubits in a CSS code isk = log 2 dim H0,0 = n − dim L A − dim L B .Logical operators preserving the subspaces H, can bechosen asAu : u L B /L A <strong>and</strong> Bv : v L A /L B .By definition, L A L B <strong>and</strong> L B L A , so the factor spacesare well defined. In the case where A <strong>and</strong> B are Pauli operators,we get a st<strong>and</strong>ard CSS code. Generally, A=V z V †<strong>and</strong> B=V x V † for some unitary operator VSU2, so anarbitrary CSS code can be mapped to a st<strong>and</strong>ard one by asuitable bitwise rotation. By a syndrome measurement for aCSS code we mean a projective measurement associatedwith the decomposition 25.Consider a CSS code such that some <strong>of</strong> the operatorsAu, Bv do not belong to the Pauli group Pn. Let us posethis question: can one perform a syndrome measurement forthis code by operations from O ideal only? It may seem thatthe answer is no, because by definition <strong>of</strong> O ideal one cannotmeasure an eigenvalue <strong>of</strong> an operator unless it belongs to thePauli group. Surprisingly, this naive answer is wrong. Indeed,imagine that we have measured part <strong>of</strong> the operatorsAu, Bv namely, those that belong to the Pauli group.Now we may restrict the remaining operators to the subspacecorresponding to the obtained measurement outcomes. Itmay happen that the restriction <strong>of</strong> some unmeasured operatorAu, which does not belong to the Pauli group, coincideswith the restriction <strong>of</strong> some other operator Ãũ Pn. Ifthis is the case, we can safely measure Ãũ instead <strong>of</strong> Au.The 15-qubit code that we use for the distillation is actuallythe simplest to our knowledge CSS code exhibiting thisPHYSICAL REVIEW A 71, 022316 2005strange behavior. We now come to an explicit description <strong>of</strong>this code.Consider a function f <strong>of</strong> four Boolean variables. Denoteby fF 2 15 the table <strong>of</strong> all values <strong>of</strong> f except f0000. Thetable is considered as a binary vector, i.e.,f = „f0001, f0010, f0011,…, f1111….Let L 1 be the set <strong>of</strong> all vectors f, where f is a linear functionsatisfying f0=0. In other words, L 1 is the linear subspacespanned by the four vectors x j , j=1, 2, 3, 4 where x jis an indicator function for the jth input bit:L 1 = linear spanx 1 ,x 2 ,x 3 ,x 4 .Let also L 2 be the set <strong>of</strong> all vectors f, where f is a polynomial<strong>of</strong> degree at most 2 satisfying f0=0. In other words,L 2 is the linear subspace spanned by the four vectors x j <strong>and</strong>the six vectors x i x j :L 2 = linear spanx 1 ,x 2 ,x 3 ,x 4 ,x 1 x 2 ,x 1 x 3 ,x 1 x 4 ,x 2 x 3 ,x 2 x 4 ,x 3 x 4 .26The definition <strong>of</strong> L 1 <strong>and</strong> L 2 resembles the definition <strong>of</strong> puncturedReed-Muller codes <strong>of</strong> order 1 <strong>and</strong> 2, respectively, seeRef. 30. Note also that L 1 is the dual space for the 15-bitHamming code. The relevant properties <strong>of</strong> the subspaces L jare stated in the following lemma.Lemma 1.1 For any uL 1 one has u0mod 8.2 For any vL 2 one has v0mod 2.3 Let 1 be the unit vector 1, 1,…, 1, 1. Then L 1=L 2 1 <strong>and</strong> L 2 =L 1 1.4 For any vectors u,vL 1 one has u·v0mod 4.5 For any vectors uL 1 <strong>and</strong> vL 2 one has u·v0mod 4.Pro<strong>of</strong>.1 Any linear function f on F 2 4 satisfying f0=0 takesvalue 1 exactly eight times if f 0 or zero times if f =0.2 All basis vectors <strong>of</strong> L 2 have weight equal to 8 thevectors x i or 4 the vectors x i x j . By linearity, all elements<strong>of</strong> L 2 have even weight.3 One can easily check that all basis vectors <strong>of</strong> L 1 areorthogonal to all basis vectors <strong>of</strong> L 2 , therefore L 1 L 2 ,L 2 L 1 . Besides, we have already proved that 1L 1 <strong>and</strong>1L 2 . Now the statement follows from dimension counting,since dim L 1 =4 <strong>and</strong> dim L 2 =10.4 Without loss <strong>of</strong> generality we may assume that u0<strong>and</strong> v0. If u=v, the statement has been already proved, seeproperty 1. If uv, then u=f, v=g for some linearlyindependent linear functions f <strong>and</strong> g. We can introduce newcoordinates y 1 ,y 2 ,y 3 ,y 4 on F 2 4 such that y 1 = fx <strong>and</strong> y 2=gx. Now u·v=y 1 y 2 =4.5 Let uL 1 <strong>and</strong> vL 2 . Since L 2 =L 1 1, there aretwo possibilities: vL 1 <strong>and</strong> v=1+w for some wL 1 . Thefirst case has been already considered. In the second case wehave022316-9


S. BRAVYI AND A. KITAEV PHYSICAL REVIEW A 71, 022316 200515u · v = u j 1 − w j = u − u · w.j=1It follows from properties 1 <strong>and</strong> 4 that u·v0mod 4. Now consider the one-qubit Hermitian operatorA = 1 2 x + y =0e −i/4e +i/4 0 = e −1/4 K x ,where K is the phase shift gate, see Eq. 1. By definition, Abelongs to the Clifford group C 1 . One can easily check thatA 2 =I <strong>and</strong> A z =− z A, so the code CSS z ,L 2 ;A,L 1 is welldefined. We claim that its code subspace coincides with thecode subspace <strong>of</strong> a certain stabilizer code.Lemma 2. Consider the decompositionC 2 15 = L 2* H,,*L 1associated with the code CSS z ,L 2 ;A,L 1 <strong>and</strong> the decompositionC 2 15 = L 2* G,,*L 1associated with the stabilizer code CSS z ,L 2 ; x ,L 1 . Forany syndrome L 1 * one hasH0, = G0,.Moreover, for any L 2 * there exists some wF 2 15 such thatfor any L 1*H, = AwG0,.27This Lemma provides a strategy to measure a syndrome<strong>of</strong> the code CSS z ,L 2 ;A,L 1 by operations from O ideal .Specifically, we measure i.e., the z part <strong>of</strong> the syndromefirst, compute w=w, apply Aw † , measure using thestabilizers x x j , <strong>and</strong> apply Aw.Pro<strong>of</strong> <strong>of</strong> the lemma. Consider an auxiliary subspace,H = L 1*H0, = G0,,*L 1corresponding to the trivial z syndrome for both CSS codes.Each state H0 can be represented as = c v v,vL 2where c v are some complex amplitudes <strong>and</strong> v=v 1 ,…,v 15 are vectors <strong>of</strong> the st<strong>and</strong>ard basis. Let us showthatAu = x u for any H, u L 1 .To this end, we represent A as x e i/4 K † . For any uL 1 <strong>and</strong>vL 2 we haveAuv = x ue i/4u−i/2u·v v = x uv,because u0mod 8 <strong>and</strong> u·v0mod 4 see Lemma 1,parts 1 <strong>and</strong> 5.Since for any uL 1 the operators Au <strong>and</strong> x u act onH in the same way, their eigenspaces must coincide, i.e.,H0,=G0, for any L 1 * .Let us now consider the subspace H, for arbitraryL 2 * , L 1 * . By definition, is a linear functional onL 2 F 2 15 ; we can extend it to a linear functional on F 2 15 , i.e.,represent it in the form v=w,v for some wF 2 15 . Thenfor any H,, vL 2 , <strong>and</strong> uL 1 we have z vAw † = − 1 w,v Aw † z v = Aw † ,AuAw † = Aw † Au = − 1 v Aw † as z <strong>and</strong> A anticommute, hence Aw † H0,. ThusH, = AwH0, = AwG0,.Lemma 2 is closely related to an interesting property <strong>of</strong>the stabilizer code CSS z ,L 2 ; x ,L 1 , namely the existence<strong>of</strong> a non-Clifford automorphism 23. Consider a one-qubitunitary operator W such thatW z W † = z <strong>and</strong> W x W † = A.It is defined up to an overall phase <strong>and</strong> obviously does notbelong to the Clifford group C 1 . However, the bitwise application<strong>of</strong> W, i.e., the operator W 15 , preserves the code subspaceG0,0. Indeed, W 15 G0,0 corresponds to the trivialsyndrome <strong>of</strong> the codeCSSW z W † ,L 2 ;W x W † ,L 1 = CSS z ,L 2 ;A,L 1 .Thus W 15 G0,0=H0,0. But H0,0=G0,0 due to thelemma.Now we are in a position to describe the distillationscheme <strong>and</strong> to estimate its threshold <strong>and</strong> yield. Suppose weare given 15 copies <strong>of</strong> the state , <strong>and</strong> our goal is to distillone copy <strong>of</strong> an H-type magic state. We will actually distillthe state,A 0 = 1 20 + e i 4 1 = e i 8 HK † H.Note that A 0 is an eigenstate <strong>of</strong> the operator A; specifically,AA 0 =A 0 . Let us also introduce the stateA 1 = z A 0 ,which satisfies AA 1 =−A 1 . Since the Clifford group C 1 actstransitively on the set <strong>of</strong> H-type magic states, we can assumethat the fidelity between <strong>and</strong> A 0 is the maximum oneamong all H-type magic states, so that022316-10


UNIVERSAL QUANTUM COMPUTATION WITH IDEAL…F H = A0 A 0 .As in Sec. V we define the initial error probability = 1 − F H 2 = A 1 A 1 .Applying the dephasing transformationD = 1 2 + AA† to each copy <strong>of</strong> , we can guarantee that is diagonal in theA 0 ,A 1 basis, i.e., = D = 1 − A 0 A 0 + A 1 A 1 .Since AC 1 , the dephasing transformation can be realizedby operations from O ideal . Thus our initial state is in = 15 = u 1 − 15−u A u A u ,15uF 228where A u =A u0 ¯ A u15 .According to the remark following the formulation <strong>of</strong>Lemma 2, we can measure the syndrome , <strong>of</strong> the codeCSS z ,L 2 ;A,L 1 by operations from O ideal only. Let us followthis scheme, omitting the very last step. So, we beginwith the state in , measure , compute w=w, applyAw † , <strong>and</strong> measure . We consider the distillation attemptsuccessful if =0. The measured value <strong>of</strong> is not importantat this stage. In fact, for any L 2 * the unnormalized postmeasurementstate is s = Aw † in Aw = in .In this equation is the projector onto the code subspaceH0,0=G0,0, i.e., = z A for z = 1L 2 vL 2 z v, A = 1L 1 uL 1Au.Let us compute the state s = in . SinceAuA w = − 1 u,w A w , z vA w = A w+v ,29one can easily see that A A w =A w if wL 1 , otherwise A A w =0. On the other h<strong>and</strong>, z A w does not vanish <strong>and</strong>depends only on the coset <strong>of</strong> L 2 that contains w. There areonly two such cosets in L 1 because L 1 =L 2 1, seeLemma 1, <strong>and</strong> the corresponding projected states areA 0 L = L2 z A 0¯0 =A 1 L = L2 z A 1¯1 =1 L2 vL 2A v ,1 L2 vL 2A v+1 .30The states A L0,1 form an orthonormal basis <strong>of</strong> the code subspace.The projections <strong>of</strong> A w for wL 1 onto the code subspaceare given by these formulas:A w =1 L2 A 0 L if w L 2 ,A w =PHYSICAL REVIEW A 71, 022316 20051 L2 A 1 L if w L 2 + 1.Now the unnormalized final state s = in can be exp<strong>and</strong>edas s1L 2 vL 21 − 15−v v A 0 L A 0 L + 1L 2 vL 2 15−v 1 − v A 1 L A 1 L .The distillation succeeds with probabilityp s = L 2 Tr s = 15−v 1 − v .vL 1The factor L 2 reflects the number <strong>of</strong> possible values <strong>of</strong> ,which all give rise to the same state s .To complete the distillation procedure, we need to apply adecoding transformation that would map the twodimensionalsubspace H0,0C 2 15 onto the Hilbertspace <strong>of</strong> one qubit. Recall that H0,0=G0,0 is the codesubspace <strong>of</strong> the stabilizer code CSS z ,L 2 ; x ,L 1 . Its logicalPauli operators can be chosen asXˆ = x 15 , Ŷ = y 15 , Ẑ = − z 15 .It is easy to see that Xˆ ,Ŷ ,Ẑ obey the correct algebraic relations<strong>and</strong> preserve the code subspace. The decoding can berealized as a Clifford operator VC 15 that maps Xˆ ,Ŷ ,Ẑ tothe Pauli operators x , y , z acting on the first qubit. Theremaining 14 qubits become unentangled with the first one,so we can safely disregard them. Let us show that the logicalstate A 0 L is transformed into A 0 up to some phase. Forthis, it suffices to check that A 0 L Xˆ A 0 L =A 0 x A 0 ,A 0 L ŶA 0 L =A 0 y A 0 , <strong>and</strong> A 0 L ẐA 0 L =A 0 z A 0 . Verifyingthese identities becomes a straightforward task if we representA 0 L in the st<strong>and</strong>ard basis:A L 0 = L 2 1/2 2 −15/2 e i/4u uuL 2= 2 −5/2 uL 1u + e −i/4 u + 1.To summarize, the distillation subroutine consists <strong>of</strong> thefollowing steps.1 Measure eigenvalues <strong>of</strong> the Pauli operators z x j , z x j x k for j,k=1,2,3,4. The outcomes determine the zsyndrome, L 2 * .2 Find w=wF 2 15 such that w,v=v for any vL 2 .3 Apply the correcting operator Aw † .4 Measure eigenvalues <strong>of</strong> the operators x x j . Theoutcomes determine the A syndrome, L 1 * .5 Declare failure if 0, otherwise proceed to the nextstep.6 Apply the decoding transformation, which takes the022316-11


S. BRAVYI AND A. KITAEV PHYSICAL REVIEW A 71, 022316 2005code subspace to the Hilbert space <strong>of</strong> one qubit.The subroutine succeeds with probabilityp s = 15−v 1 − v .vL 131In the case <strong>of</strong> success, it produces the normalized outputstate out = 1 − out A 0 A 0 + out A 1 A 1 32characterized by the error probability out = p s−1vL 2 15−v 1 − v .33The sums in Eqs. 31 <strong>and</strong> 33 are special forms <strong>of</strong> socalledweight enumerators. The weight enumerator <strong>of</strong> a subspaceLF 2 n is a homogeneous polynomial <strong>of</strong> degree n intwo variables, namelyIn this notation,W L x,y = x n−u y u .uLp s = W L1,1 − , out = W L 2,1 − W L1,1 − .The MacWilliams identity 30 relates the weight enumerator<strong>of</strong> L to that <strong>of</strong> L :W L x,y = 1L W Lx + y,x − y.Applying this identity <strong>and</strong> taking into account that L 2 =L 1 1 <strong>and</strong> that u0mod 2 for any uL 1 see Lemma 1,we getp s = 1 16 W L 11,1 − 2, out = 1 21 − W L 11 − 2,1W L1 1,1 − 2.34The weight enumerator <strong>of</strong> the subspace L 1 is particularlysimple:W L1 x,y = x 15 + 15x 7 y 8 .Substituting this expression into Eq. 34, we arrive at thefollowing formulas:1 + 151 − 28p s = , 3516 out = 1 − 151 − 27 + 151 − 2 8 − 1 − 2 1521 + 151 − 2 8 . 36The function out is plotted in Fig. 3. Solving the equation out = numerically, we find the threshold error probability: 0 0.141.37FIG. 3. The final error probability out for the H-type statesdistillation.Let us examine the asymptotic properties <strong>of</strong> this scheme.For small the distillation subroutine succeeds with probabilityclose to 1, therefore the yield is close to 1/15. Theoutput error probability is out 35 3 .38Now suppose that the subroutine is applied recursively. Fromn copies <strong>of</strong> the state with a given , we distill one copy <strong>of</strong>the magic state A 0 with the final error probability out n, 1 35 353 k , 15 k n,where k is the number <strong>of</strong> recursion levels here we neglectthe fluctuations in the number <strong>of</strong> successful distillation attempts.Solving these equation, we obtain the relation out n, 35n , = 1/log 3 15 0.4. 39It characterizes the efficiency <strong>of</strong> the distillation scheme.VII. CONCLUSION AND SOME OPEN PROBLEMSWe have studied a simplified model <strong>of</strong> fault-tolerant quantumcomputation in which operations from the Cliffordgroup are realized exactly, whereas decoherence occurs onlyduring the preparation <strong>of</strong> nontrivial ancillary states. Themodel is fully characterized by a one-qubit density matrix describing these states. It is shown that a good strategy forsimulating universal quantum computation in this model is“magic states distillation.” By constructing two particulardistillation schemes we find a threshold polarization <strong>of</strong> above which the simulation is possible.The most exciting open problem is to underst<strong>and</strong> the computationalpower <strong>of</strong> the model in the region <strong>of</strong> parameters1 x + y + z 3/ 7 which corresponds to FT * F T F T , see Sec. I. In this region, the distillation scheme basedon the five-quit code does not work, while the Gottesman-Knill theorem does not yet allow the classical simulation.One possibility is that a transition from classical to universalquantum behavior occurs on the octahedron boundary, x + y + z =1.022316-12


UNIVERSAL QUANTUM COMPUTATION WITH IDEAL…To prove the existence <strong>of</strong> such a transition, one it sufficesto construct a T-type states distillation scheme having thethreshold fidelity F T * . A systematic way <strong>of</strong> constructing suchschemes is to replace the five-qubit by a GF4-linear stabilizercode. A nice property <strong>of</strong> these codes is that the bitwiseapplication <strong>of</strong> the operator T preserves the code subspace <strong>and</strong>acts on the encoded qubit as T, see Ref. 31 for more details.One can check that the error-correcting effect described inSec. V takes place for an arbitrary GF4-linear stabilizercode, provided that the number <strong>of</strong> qubits is n=6k−1 for anyinteger k. Unfortunately, numerical simulations we performedfor some codes with n=11 <strong>and</strong> n=17 indicate thatthe threshold fidelity increases as the number <strong>of</strong> qubits increases.So it may well be the case that the five-qubit code isthe best GF4-linear code as far as the distillation is concerned.From the experimental point <strong>of</strong> view, an exciting openproblem is to design a physical system in which reliablestorage <strong>of</strong> quantum information <strong>and</strong> its processing by Cliffordgroup operations is possible. Since our simulationscheme tolerates strong decoherence on the ancilla preparationstage, such a system would be a good c<strong>and</strong>idate for apractical quantum computer.ACKNOWLEDGMENTSWe thank Mikhail Vyalyi for bringing to our attentionmany useful facts about the Clifford group. This work hasbeen supported in part by the National Science Foundationunder Grant No. EIA-0086038.APPENDIXThe purpose <strong>of</strong> this section is to prove Eq. 15. Let usintroduce this notation:Tˆ 0 = T 00000 <strong>and</strong> Tˆ 1 = T 11111 .Consider the set S + 5S5 consisting <strong>of</strong> all possible tensorproducts <strong>of</strong> the Pauli operators x , y , z on five qubitsclearly, S + 5=4 5 =S5/2 since elements <strong>of</strong> S5 mayhave a plus or minus sign. For each gS + 5 let g0,5 be the number <strong>of</strong> qubits on which g acts nontriviallye.g., x x y I I=3. We haveTˆ 0Tˆ 0 = 1 2 5gS + 5 1g 3Now let us exp<strong>and</strong> the formula 13 for the projector .Denote by G P5 the Abelian group generated by the stabilizersS 1 ,S 2 ,S 3 ,S 4 . It consists <strong>of</strong> 16 elements. Repeatedlyconjugating the stabilizer S 1 by the operator Tˆ =T 5 , we getthree elements <strong>of</strong> G:S 1 = x z z x I,g.S 1 S 3 S 4 = z y y z I,S 3 S 4 = y x x y I.Due to the cyclic symmetry mentioned in Sec. V, the 15cyclic permutations <strong>of</strong> these elements also belong to G; togetherwith the identity operator they exhaust the group G.Thus GS + 5, <strong>and</strong> we have = 1 h.16 hGTaking into account that Trgh=2 5 g,h for any g,hS + 5,we getTˆ 0Tˆ 0 = 1 2 9 PHYSICAL REVIEW A 71, 022316 2005hG gS + 53 −g/2 Trgh = 1 3 −g/2 = 1 16 gG 6 .Similar calculations show that Tˆ 1Tˆ 1= 1 6 .1 E. Knill, R. Laflamme, <strong>and</strong> W. Zurek, Science 279, 3421998.2 C. Zalka, e-print quant-ph/9612028.3 A. Steane, Phys. Rev. Lett. 78, 2252 1997.4 E. Dennis, A. Kitaev, A. L<strong>and</strong>ahl, <strong>and</strong> J. Preskill, J. Math.Phys. 43, 4452 2002.5 A. Kitaev, Ann. Phys. N.Y. 303, 2 2003.6 M. Freedman, M. Larsen, <strong>and</strong> Z. Wang, e-print quant-ph/0001108.7 M. Freedman, A. Kitaev, M. Larsen, <strong>and</strong> Z. Wang, Bull., NewSer., Am. Math. Soc. 40, 31 2002.8 C. Mochon, Phys. Rev. A 69, 032306 2004.9 G. Moore <strong>and</strong> N. Read, Nucl. Phys. B 360, 362 1991.10 C. Nayak <strong>and</strong> F. Wilczek, Nucl. Phys. B 479, 529 1996.11 B. Doucot <strong>and</strong> J. Vidal, Phys. Rev. Lett. 88, 227005 2001.12 M. Feigel’man <strong>and</strong> L. I<strong>of</strong>fe, Phys. Rev. B 66, 224503 2002.13 J. Preskill, e-print quant-ph/9712048.14 D. Gottesman, Ph.D. thesis, Caltech, Pasadena, 1997, URLhttp://arxiv.org/abs/quant-ph/9705052.15 M. Nielsen <strong>and</strong> I. Chuang, Quantum Computation <strong>and</strong> QuantumInformation Cambridge University Press, Cambridge, Engl<strong>and</strong>,2000.16 D. Gottesman, Phys. Rev. A 57, 127 1998.17 E. Knill, e-print quant-ph/0402171.18 E. Knill, e-print quant-ph/0404104.19 D. Gottesman <strong>and</strong> I. Chuang, Nature London 402, 3901999.20 W. Dur <strong>and</strong> H. Briegel, Phys. Rev. Lett. 90, 067901 2003.21 D. Aharonov, e-print quant-ph/9602019.22 E. Dennis, Phys. Rev. A 63, 052314 2001.23 E. Knill, R. Laflamme, <strong>and</strong> W. Zurek, e-print quant-ph/9610011.022316-13


S. BRAVYI AND A. KITAEV PHYSICAL REVIEW A 71, 022316 200524 A. Calderbank, E. Rains, P. Shor, <strong>and</strong> N. Sloane, Phys. Rev.Lett. 78, 405 1997.25 A. Ambainis <strong>and</strong> D. Gottesman, e-print quant-ph/0310097.26 C. Bennett, D. DiVincenzo, J. Smolin, <strong>and</strong> W. Wootters, Phys.Rev. A 54, 3824 1996.27 R. Laflamme, C. Miquel, J. Paz, <strong>and</strong> W. Zurek, Phys. Rev.Lett. 77, 198 1996.28 A. Calderbank <strong>and</strong> P. Shor, Phys. Rev. A 54, 1098 1996.29 A. Steane, Proc. R. Soc. London, Ser. A 452, 2551 1996.30 F. MacWilliams <strong>and</strong> N. Sloane, The Theory <strong>of</strong> Error-Correcting Codes North-Holl<strong>and</strong>, Amsterdam, 1981.31 A. Calderbank, E. Rains, P. Shor, <strong>and</strong> N. Sloane, e-print quantph/9608006.022316-14


PRL 103, 170504 (2009) PHYSICAL REVIEW LETTERSweek ending23 OCTOBER 2009Tight Noise Thresholds for Quantum Computation with Perfect Stabilizer OperationsWim van Dam*Department <strong>of</strong> Computer Science, University <strong>of</strong> California, Santa Barbara, California 93106, USA<strong>and</strong> Department <strong>of</strong> Physics, University <strong>of</strong> California, Santa Barbara, California 93106, USAMark Howard †Department <strong>of</strong> Physics, University <strong>of</strong> California, Santa Barbara, California 93106, USA(Received 21 July 2009; published 23 October 2009)We study how much noise can be tolerated by a universal gate set before it loses its quantumcomputationalpower. Specifically we look at circuits with perfect stabilizer operations in addition toimperfect nonstabilizer gates. We prove that for all unitary single-qubit gates there exists a tightdepolarizing noise threshold that determines whether the gate enables universal quantum computationor if the gate can be simulated by a mixture <strong>of</strong> Clifford gates. This exact threshold is determined by theClifford polytope spanned by the 24 single-qubit Clifford gates. The result is in contrast to the situationwherein nonstabilizer qubit states are used; the thresholds in that case are not currently known to be tight.DOI: 10.1103/PhysRevLett.103.170504Introduction.—A way to study the resources needed foruniversal quantum computation (UQC) is to analyze thetransition from a system that can provide UQC to one thatis classically efficiently simulable. A particularly usefulexample <strong>of</strong> a classically simulable system is given by thestabilizer operations, which are made by a combination <strong>of</strong>preparation <strong>of</strong> j0i states, unitary Clifford gates, measurementsin the fj0i; j1ig basis, <strong>and</strong> classical control determinedby the measurement outcomes. The Gottesman-Knill theorem tells us that stabilizer operations can beefficiently simulated classically (see, for example, [1],Theorem 10.7), while it also known that the addition <strong>of</strong>any other one-qubit gate outside the Clifford group willenable the system to perform UQC. This fact provides uswith a framework for testing tolerance to noise—one canexamine how noisy this additional non-Clifford gate can bebefore it becomes classically simulable itself. If the non-Clifford operation has become a probabilistic combination<strong>of</strong> Clifford gates due to the noise, then we know that we areunequivocally in the classical computational regime. Thenoise rate where the extra gate becomes simulable (whereit enters the ‘‘Clifford polytope’’ [2]) is thus an upperbound for fault tolerance. If the converse is true—i.e., ifany operation outside the Clifford polytope enables UQC,then the threshold is tight. In this Letter we show that forsingle-qubit gates undergoing depolarizing such a tightnoise threshold does indeed apply. We will do so byproving that any depolarized gate that lies outside theClifford polytope <strong>of</strong> single-qubit operations, in combinationwith noiseless stabilizing operations, allows for UQC.This result should be contrasted to the situation for nonstabilizerqubit states where the thresholds in that case arenot currently known to be tight. In fact, a recent result byCampbell <strong>and</strong> Browne [3] states that achieving tightthresholds for all nonstabilizer qubit states is impossibleif the number <strong>of</strong> copies <strong>of</strong> the resource state must be finite.PACS numbers: 03.67.Lx, 03.67.PpWe will consistently assume that Clifford gates can beimplemented perfectly, motivated by the fact that thesegates can be implemented fault tolerantly by applyingthem transversally <strong>and</strong> to encoded states [4–7]. The faulttolerantimplementation <strong>of</strong> Clifford gates naturally carrieswith it a threshold <strong>of</strong> its own, independent <strong>of</strong> the kind wediscuss in this Letter. The current model is particularlyrelevant to the so-called Pfaffian quantum Hall state intopological quantum computation [8,9], the two-qubitClifford group (but only the Clifford group) can be implementedusing braiding making these operations naturallyfault tolerant. The additional resource required to performUQC will likely be highly noisy, <strong>and</strong> so we can see theparallels with our model.We will begin by listing a couple <strong>of</strong> previously knownresults in this area. Next, we will discuss the connectionbetween the geometry <strong>of</strong> the Clifford polytope <strong>and</strong> stabilizermeasurements, <strong>and</strong> show that tightness <strong>of</strong> a magicstatedistillation procedure for single-qubit states automaticallyensures tight thresholds for non-Clifford gates undergoingany kind <strong>of</strong> unital noise. Finally, we show thatcurrently known magic-state distillation techniques aresufficient to prove tight thresholds for a non-Clifford gateundergoing depolarizing noise.Previously known results.—The idea <strong>of</strong> using perfectstabilizer operations in conjunction with imperfect nonstabilizerstates to perform UQC originates with Knill [7].Shortly after, Bravyi <strong>and</strong> Kitaev [10] showed that mostnonstabilizer qubit states (when sufficiently many copiesare available) can be purified (‘‘distilled’’), using onlystabilizer operations, towards a pure nonstabilizer state (a‘‘magic state’’). Since a universal gate set can be createdfrom perfect stabilizer operations <strong>and</strong> a supply <strong>of</strong> magicstates [10], we see that allowing access to a supply <strong>of</strong>appropriate (possibly impure) nonstabilizer qubit ancillaspromotes the power <strong>of</strong> stabilizer operations from classi-0031-9007=09=103(17)=170504(4) 170504-1 Ó 2009 The American Physical Society


PRL 103, 170504 (2009) PHYSICAL REVIEW LETTERSweek ending23 OCTOBER 2009cally simulable to UQC. The conditions on the ancillaryqubits to enable UQC is that they are sufficiently close tobeing one <strong>of</strong> the 20 so-called magic states that lie on thesurface <strong>of</strong> the Bloch sphere. The two classes <strong>of</strong> magic state(see Fig. 1) are the jHi type <strong>and</strong> jTi type, where all jHitype states can be derived by applying a Clifford operationpt<strong>of</strong>fiffiffisome canonical representative jHi ¼ðj0iþe i=4 j1iÞ=2 , <strong>and</strong> similarly for jTi-type states likep jTi¼cosð#Þj0iþe i=4 sinð#Þj1i, where cosð2#Þ ¼1= ffiffiffi3 . The routines usedin [10] were unable to distill qubit states just outside theedges <strong>and</strong> faces <strong>of</strong> the octahedron <strong>of</strong> Fig. 1. Reichardt [11]subsequently proposed an improved routine that closed thegap in the jHi direction (along the edges <strong>of</strong> the octahedron).Virmani et al. [12] suggested using the convex hull<strong>of</strong> Clifford operations in order to find gates’ robustness tovarious types <strong>of</strong> noise. In particular, they considered gatesthat are diagonal in the computational basis. Plenio <strong>and</strong>Virmani [13] subsequently extended this idea by analyzingcases where noise was allowed to affect the stabilizer operationstoo. Buhrman et al. [2] used a similar idea (thatnoise causes non-Clifford gates to eventually become ableto be implemented via Clifford gates only) to find the non-Clifford gate that is most resistant to depolarizing noise—a=8 rotation about the Z axis (or the same gate modulosome Clifford operation). Reichardt [14] showed that thisparticular gate enabled UQC right up to its threshold noiserate (about 45%), as well as considering in detail the process<strong>of</strong> reducing multiqubit states to single-qubit states usingpostselected stabilizer operations. Our current resulthere generalizes this tightness result to all possible singlequbitgates.Preliminaries <strong>and</strong> notation.—Let us parameterize anarbitrary single-qubit SU(2) gate as followsUð; ; Þ ¼ ei cosðÞ e i sinðÞe i sinðÞ e i cosðÞ: (1)The representation <strong>of</strong> this rotation in SO(3) is denoted byFIG. 1 (color online). Magic states <strong>and</strong> the octahedron: Some<strong>of</strong> the single-qubit magic states: jHi type states are designatedwith black arrows, jTi type states with white arrows. Theoctahedron defined by jxjþjyjþjzj 1 depicts the singlequbitstates that can be created by stabilizer operations.Reichardt [11] showed that distillation techniques work rightup to the edges <strong>of</strong> the octahedron (i.e., tight in the jHi direction).Current distillation techniques are unable to distill states justoutside the faces <strong>of</strong> the octahedron (i.e., not tight in the jTidirection).170504-2Rð; ; Þ. Implementing a rotation R while suffering depolarizingnoise (with noise rate p), means that this noisyoperation is represented by the rescaling M ¼ð1 pÞR, afact that we will need later.Often we will apply the unitary Uð; ; Þ to one half <strong>of</strong>an entangled Bell pair, ji ¼p 1 ffiffi2ðj00iþj11iÞ, yielding ¼ðI UÞjihjðI UÞ y : (2)If we use the two-qubit Pauli operators as a basis for thedensity matrix then we can find the 16 real coefficientsc ij ¼ Trðð i j ÞÞ so that ¼ 1 4Xcij ð i j Þ; i;j 2fI; X; Y; Zg: (3)Since we have applied a local unitary to a maximallyentangled state, the coefficients (c IX , c IY , c IZ , c XI , c YI ,c ZI ) are always zero. Comparing the 9 coefficientsfc XX ;c XY ; ...;c ZZ g one can see that these are the same asthe entries <strong>of</strong> the SO(3) matrix Rð; ; Þ. More precisely,01c XX c YX c ZXRð; ; Þ ¼@c XY c YY c ZYA; (4)c XZ c YZ c ZZwhere the c ij are obviously also functions <strong>of</strong> (, , ).If we represent the 24 single-qubit Clifford operations asSO(3) matrices, then they are simply signed permutationmatrices with unit determinant (they are a matrix representation<strong>of</strong> the elements <strong>of</strong> the chiral octahedral symmetrygroup or, equivalently, the symmetry group S 4 ). We labelthese operations C i <strong>and</strong> so the convex hull <strong>of</strong> the C i (the socalledClifford polytope) is given by X 24P ¼ p i C i with p i 0 <strong>and</strong> X24p i ¼ 1 : (5)i¼1Geometrically, the Clifford polytope is a closed polyhedronin R 9 that has 24 vertices (each vertex representingone <strong>of</strong> the C i ). This polytope can also be defined by thebounding inequalities <strong>of</strong> its 120 facets. The concise description<strong>of</strong> these facets used by Buhrman et al. [2]isgivenby the setF ¼fC i FC j ji; j 2f1; ...; 24g;F 2fA; A T ;Bgg; (6)where0 10 11 0 00 1 0A ¼ @ 1 0 0A <strong>and</strong> B ¼ @ 1 0 1 A: (7)1 0 01 0 1At times we will have reason to refer to different subsets <strong>of</strong>the set <strong>of</strong> facets F so we use the obvious notation F ¼F A [ F AT [ F B . It is useful to note that all the facetsderived from A comprise a single column with 1 entries<strong>and</strong> zeros elsewhere, <strong>and</strong> similarly for the row facetsderived from A T ; hence, jF A j¼jF A Tj¼3 2 3 ¼ 24.There are jF B j¼72 ‘‘B-type’’ facets, which can be constructedas follows: (i) Pick one position in a 3 3 matrixF, e.g., row i <strong>and</strong> column j <strong>and</strong> put 1 there (9 2 ¼ 18i¼1


PRL 103, 170504 (2009) PHYSICAL REVIEW LETTERSweek ending23 OCTOBER 2009choices), (ii) Fill the remaining entries not in row i orcolumn j with 1 such that detðFÞ ¼ 2 (4 choices).To determine whether or not an operation M is inside theClifford polytope P we take the elementwise inner product(or Frobenius inner product) between M <strong>and</strong> the facets F 2F <strong>of</strong> the polytopeM F ¼ X3i;j¼1M i;j F i;j ¼ TrðM T FÞ: (8)Using the above notation, a 3 3 matrix M is inside thepolytope P if <strong>and</strong> only if for all F 2 F we have M F 1.Interpreting the facets <strong>of</strong> the Clifford polytope.—Ourpro<strong>of</strong> will involve applying some non-Clifford gate toone half <strong>of</strong> a Bell Pair [as in Eq. (2)] <strong>and</strong> then postselectingon the outcomes <strong>of</strong> various stabilizer measurements. Aftersome further stabilizer operations, this measurement ultimatelyhas the effect <strong>of</strong> taking our two-qubit state to asingle-qubit state 0 (times some stabilizer state that we donot care about), which we then distill using magic-statedistillation (see [14] for a more general discussion <strong>of</strong> thesekinds <strong>of</strong> techniques). For example, performing a YX measurementon <strong>and</strong> postselecting on a ‘‘þ1’’ outcome (i.e.,projecting with ¼ 1 2ðII þ YXÞ) leads to a single-qubitstate 0 with a Bloch vector given by~rð 0 Þ¼ 0; c XZ c ZY;c II þ c YXc XY þ c ZZ: (9)c II þ c YXThe form <strong>of</strong> this vector means that it lies in the YZ plane(see Fig. 1), where we know that distillation techniqueswork right up to the edge jyjþjzj ¼1 <strong>of</strong> the octahedron.We can check if ~r is outside the octahedron by simplycomparing the L 1 norm <strong>of</strong> ~r with 1. Rearranged, thecondition k~rk 1 > 1 for being outside the octahedron isjc XZ c ZY jþj ðc XY þ c ZZ Þj > jc II þ c YX j: (10)Given the correspondence between the coefficients c ij <strong>and</strong>the elements <strong>of</strong> R [see Eq. (4)] we can rewrite the abovecondition (dropping the absolute value operators) as a facetinequality010 1 0R F>1 where F ¼ @ 1 0 1 A: (11)1 0 1This facet is a legitimate ‘‘B-type’’ facet <strong>and</strong> a littlethought shows that, had we applied the single-qubit Paulioperations [as SO(3) rotations] X, Y or Z to ~rð 0 Þ above, wewould arrive at three other ‘‘B-type’’ facets0 10 1 0@ 1 0 1A;1 0 10 10 1 0@ 1 0 1A;1 0 10@10 1 01 0 1 A; (12)1 0 1respectively. Note that all four facet inequalities combinedcould be simplified to the form Eq. (10) above. We omit thedetails, but it is straightforward to show that all 72‘‘B-type’’ facets correspond to postselecting on some(weight two) Pauli operator, <strong>and</strong> possibly performing asingle-qubit Pauli rotation on the resulting 0 .170504-3It is somewhat more straightforward to see the geometricalinterpretation <strong>of</strong> the ‘‘A ðTÞ -type’’ facets. For example,the canonical A given in Eq. (7), merely returns the sum <strong>of</strong>the elements <strong>of</strong> the Bloch vector ~r, arising from a rotationapplied to the X ‘‘þ1’’ eigenstate.0 1R A ¼ X31r i where ~r ¼ R@0 A: (13)i¼10In general, an operation M having an inner product greaterthan one with some ‘‘A-type’’ facet simply means that M,applied to some initial vector corresponding to a stabilizerstate, brings that vector to a final position outside theoctahedron.The preceding discussion shows us that if magic-statedistillation were possible everywhere outside the octahedron,then every unital operation outside the Clifford polytopewould be distillable—either straightforwardly or byusing postselection, depending on which facet it violated.Using current (not tight) distillation routines however, wewould be unable to deal with some operations violating an‘‘A-type’’ facet by a fairly small amount. In the nextsection we show that, for depolarizing noise, any noisyrotation violating an ‘‘A-type’’ facet also violates a‘‘B-type’’ facet. Since ‘‘B-type’’ facets correspond to jHistate distillation, the results we obtain are tight.Tight threshold for depolarizing noise.—The claim weshall prove is that, anytime a matrix M ¼ð1 pÞR, representinga depolarized rotation, is outside some ‘‘A-type’’facet then there exists a ‘‘B-type’’ facet that M also liesoutside. In fact, we will prove the slightly stronger statementthat for all R 2 SOð3Þ8A 2 F A [ F A T; 9B 2 F B such that R ðB AÞ0:(14)To simplify the pro<strong>of</strong> we will repeatedly make use <strong>of</strong> thesymmetries <strong>of</strong> the problem [see Eq. (6)]. We will pick acanonical ‘‘A-type’’ facet <strong>and</strong> assume that this gives thelargest inner product with R <strong>of</strong> all the F 2 F A . If there wasa larger inner product with some F 2 F A T, then we couldjust relabel R T as R. We can assume that the facet with onesin the first column [the A in Eq. (7)] gives the biggest innerproduct since Tr½R T ðC i AC j ÞŠ ¼ TrðC j R T C i AÞ¼Tr½ðC k RC l Þ T AŠ, which shows that the inner product <strong>of</strong> Rwith a different F 2 F A is the same as the inner productbetween a different (but related via Clifford operations)rotation <strong>and</strong> the canonical ‘‘A-type’’ facet.The pro<strong>of</strong> will hinge on an entry <strong>of</strong> R, outside <strong>of</strong> the firstcolumn, being larger than the rest <strong>of</strong> the elements outsidethe first column. As such, let us define 12 matrices closelyrelated to A <strong>and</strong> call them A 0 i0 1 0 10 11 10 110A 0 1 ¼ @ 1 0 0A A 0 2 ¼ @ 100A A 0 12 ¼ @ 100A1 0 0 100101(15)


PRL 103, 170504 (2009) PHYSICAL REVIEW LETTERSweek ending23 OCTOBER 2009such that the index i <strong>of</strong> the largest inner product R A 0 i tellsus the sign <strong>and</strong> location <strong>of</strong> the largest magnitude elementoutside the first column. Once again, symmetry allows usto assume that A 0 1 yields the largest inner product becausethe rest <strong>of</strong> the A 0 i can be derived from A 0 1 via Cliffordrotations801 < 001fA 0 i g¼ @ 100A:010jA 0 10 191 0 0@ 0 0 1Aj 2f1;2;3g=kk 2f1;2;3;4g; :0 10(16)801 0 1 0< þ þ þþR 2 @ þ þ A; @ þ þ A; @ þ:þ þ þ þ þ þ þFor a matrix R to be an SO(3) rotation there are constraintson the signs <strong>of</strong> the elements R i;j ; i.e., there are eightchoices for the first column, 6 choices for the secondcolumn <strong>and</strong> 2 for the third. Given that A is the maximumfacet for R, we have fixed the signs positively in the firstcolumn, reducing the number <strong>of</strong> types <strong>of</strong> rotation to 6 2 ¼ 12. Since A 0 1 gives the maximum inner product with R<strong>of</strong> all A 0 i we have that R 1;2 < 0, which reduces the number<strong>of</strong> rotation types further to 3 2 ¼ 6. It can be shown thatR 1;2 having larger magnitude than the rest <strong>of</strong> the elementsR i;j ði 2f1; 2; 3g;j2f2; 3gÞ restricts the type <strong>of</strong> rotationfurther to one the following four typesþ1 0 19þ þ =A; @ þ A; : (17)þ þ þ þThis should not be surprising if one considers that R 1;2 ¼ðR 2;1 R 3;3 R 2;3 R 3;1 Þ because <strong>of</strong> the structure <strong>of</strong> SO(3)matrices, <strong>and</strong> the sign patterns listed above ensure jR 1;2 j isas large as possible.We claim that the B 2 F B <strong>of</strong> Eq. (9) will suffice toprove the desired inequality R ðB AÞ 0, which readsin matrix form0 101þ 1 1 0@ þ A@0 0 1 A 0: (18)þ þ 0 0 1Using the relevant entries <strong>of</strong> R we define a pair <strong>of</strong> 2 vectors~u <strong>and</strong> ~v as ~u ¼ðR 1;1 ;R 1;2 Þ, ~v ¼ðR 2;3 ;R 3;3 Þ so that we canrewrite the above inequality Eq. (17) ask ~vk 1 k~uk 1 0: (19)The L 2 normalization <strong>of</strong> all the rows <strong>and</strong> columns <strong>of</strong> therotation matrix R means that ~u <strong>and</strong> ~v have the same L 2norm. With reference to Fig. 2, it should be clear thatbecause ~u has an L 1 norm at least as big as that <strong>of</strong> ~v(because jR 1;2 jjR 2;3 j; jR 3;3 j), it holds that the L 1 norm <strong>of</strong>~v is automatically at least as large as the L 1 norm <strong>of</strong> ~u, asdesired.Summary.—We showed that for any unitary one-qubitgate undergoing depolarizing noise with rate p it holds thatFIG. 2. Pro<strong>of</strong> <strong>of</strong> Eq. (19): For any pair <strong>of</strong> two vectors ~u <strong>and</strong> ~vwith the same L 2 norm, the vector with greater L 1 norm hassmaller L 1 norm. A vector pointing towards the black dot hassimultaneously minimal L 1 norm <strong>and</strong> maximal L 1 norm.if its SO(3) representation M ¼ð1 pÞR lies outside theClifford polytope P , then it must be the case that there is afacet B 2 F B such that M B>1. In turn, this means thatif this noisy gate is applied to a Bell pair ji ¼p 1 ffiffi2ðj00iþj11iÞ <strong>and</strong> an appropriate stabilizer measurement is performed,then, conditionally on the outcome <strong>of</strong> the measurement,one obtains a state that can be transformed usingClifford gates into a single-qubit state with jyjþjzj > 1 inthe Bloch ball representation. By the result <strong>of</strong> Reichardt[11] such states enable stabilizing operations to performuniversal quantum computation.This material is based upon work supported by theNational Science Foundation under Grant No. 0917244.*v<strong>and</strong>am@cs.ucsb.edu† mhoward@physics.ucsb.edu[1] Michasel A. Nielsen <strong>and</strong> Isaac L. Chuang, QuantumComputation <strong>and</strong> Quantum Information (CambridgeUniversity Press, Cambridge, Engl<strong>and</strong>, 2000).[2] H. Buhrman, R. Cleve, M. Laurent, N. Linden, A.Schrijver, <strong>and</strong> F. Unger, Annual IEEE Symposium onFoundations <strong>of</strong> Computer Science (2006), p 411.[3] E. T. Campbell <strong>and</strong> D. E. Browne, arXiv:0908.0836v2.[4] A. R. Calderbank <strong>and</strong> P. W. Shor, Phys. Rev. A 54, 1098(1996).[5] A. Steane, Proc. R. Soc. A 452, 2551 (1996).[6] D. Gottesman, Phys. Rev. A 57, 127 (1998).[7] E. H. Knill, arXiv:quant-ph/0402171v1.[8] M. Freedman, C. Nayak, <strong>and</strong> K. Walker, Phys. Rev. B 73,245307 (2006).[9] L. S. Georgiev, Phys. Rev. B 74, 235112 (2006).[10] S. Bravyi <strong>and</strong> A. Kitaev, Phys. Rev. A 71, 022316 (2005).[11] Ben W. Reichardt, Quant. Info. Proc. 4, 251 (2005).[12] S. Virmani, S. F. Huelga, <strong>and</strong> M. B. Plenio, Phys. Rev. A71, 042328 (2005).[13] M. B. Plenio <strong>and</strong> S. Virmani, arXiv:0810.4340.[14] B. W. Reichardt, arXiv:quant-ph/0608085v1.170504-4


VOLUME 86, NUMBER 22 P H Y S I C A L R E V I E W L E T T E R S 28 MAY 2001A One-Way Quantum Computer<strong>Robert</strong> <strong>Raussendorf</strong> <strong>and</strong> Hans J. BriegelTheoretische Physik, Ludwig-Maximilians-Universität München, Germany(Received 25 October 2000)We present a scheme <strong>of</strong> quantum computation that consists entirely <strong>of</strong> one-qubit measurements on aparticular class <strong>of</strong> entangled states, the cluster states. The measurements are used to imprint a quantumlogic circuit on the state, thereby destroying its entanglement at the same time. Cluster states are thusone-way quantum computers <strong>and</strong> the measurements form the program.DOI: 10.1103/PhysRevLett.86.5188A quantum computer promises efficient processing <strong>of</strong>certain computational tasks that are intractable with classicalcomputer technology [1]. While basic principles <strong>of</strong> aquantum computer have been demonstrated in the laboratory[2], scalability <strong>of</strong> these systems to a large number <strong>of</strong>qubits [3], essential for practical applications such as theShor algorithm, represents a formidable challenge. Most<strong>of</strong> the current experiments are designed to implement sequences<strong>of</strong> highly controlled interactions between selectedparticles (qubits), thereby following models <strong>of</strong> a quantumcomputer as a (sequential) network <strong>of</strong> quantum logicgates [4,5].Here we propose a different model <strong>of</strong> a scalable quantumcomputer. In our model, the entire resource for thequantum computation is provided initially in the form <strong>of</strong>a specific entangled state (a so-called cluster state [6]) <strong>of</strong>a large number <strong>of</strong> qubits. Information is then written ontothe cluster, processed, <strong>and</strong> read out from the cluster byone-particle measurements only. The entangled state <strong>of</strong>the cluster thereby serves as a universal “substrate” for anyquantum computation. Cluster states can be created efficientlyin any system with a quantum Ising-type interaction(at very low temperatures) between two-state particles ina lattice configuration.We consider two- <strong>and</strong> three-dimensional arrays <strong>of</strong>qubits that interact via an Ising-type next-neighbor interaction[6] described by a Hamiltonian H int gt 3P 11sza12s a0 za,a 0 2 2 2 1 4 gt P a,a 0 sza s a0 z [7] whosestrength gt can be controlled externally. A possiblerealization <strong>of</strong> such a system is discussed below. A qubit atsite a can be in two states j0 a j0 z,a or j1 a j1 z,a ,the eigenstates <strong>of</strong> the Pauli phase flip operator szasza ji a 21 i ji a . These two states form the computationalbasis. Each qubit can equally be in an arbitrarysuperposition state aj0 1 bj1, jaj 2 1 jbj 2 1. Forour purpose, we initially prepare all qubits in the superpositionj1 j0 1 j1 p 2, an eigenstate <strong>of</strong> thePauli spin flip operator s x s x j6 6j6. H int isthen switched on for an appropriately chosen finite timeinterval T, where R T0 dt gt p, by which a unitarytransformation S is realized. Since H int acts uniformly onthe lattice, entire clusters <strong>of</strong> neighboring particles becomeentangled in one single step. The quantum state jF C ,PACS numbers: 03.67.Lx, 03.65.Udthe state <strong>of</strong> a cluster C <strong>of</strong> neighboring qubits, which isthereby created provides in advance all entanglement thatis involved in the subsequent quantum computation. It hasbeen shown [6] that the cluster state jF C is characterizedby a set <strong>of</strong> eigenvalue equationss axOa 0 [ngbhas a0 z jF C 6jF C , (1)where ngbha specifies the sites <strong>of</strong> all qubits that interactwith the qubit at site a [ C . The eigenvalues are determinedby the distribution <strong>of</strong> the qubits on the lattice.The equations (1) are central for the proposed computationscheme. As an example, a measurement on an individualqubit <strong>of</strong> a cluster has a r<strong>and</strong>om outcome. On the otherh<strong>and</strong>, Eqs. (1) imply that any two qubits at sites a, a 0 [ Ccan be projected into a Bell state by measuring a subset <strong>of</strong>the other qubits in the cluster. This property will be used todefine quantum channels that allow us to propagate quantuminformation through a cluster.We show that a cluster state jF C can be used as a substrateon which any quantum circuit can be imprinted byone-qubit measurements. In Fig. 1 this scheme is illustrated.For simplicity, we assume that in a certain region<strong>of</strong> the lattice each site is occupied by a qubit. This requirementis not essential as will be explained below [see(d)]. In the first step <strong>of</strong> the computation, a subset <strong>of</strong>qubits is measured in the basis <strong>of</strong> s z which effectivelyremoves them. In Fig. 1 these qubits are denoted by “ Ø.”information flowquantum gateFIG. 1. Quantum computation by measuring two-state particleson a lattice. Before the measurements the qubits are in thecluster state jF C <strong>of</strong> (1). Circles Ø symbolize measurements <strong>of</strong>s z , vertical arrows are measurements <strong>of</strong> s x , while tilted arrowsrefer to measurements in the x-y plane.5188 0031-90070186(22)5188(4)$15.00 © 2001 The American Physical Society


VOLUME 86, NUMBER 22 P H Y S I C A L R E V I E W L E T T E R S 28 MAY 2001The state jF C is thereby projected into a tensor productjm C nN ≠ j ˜F N consisting <strong>of</strong> the state jm C nN <strong>of</strong>all measured particles (subset C nN ) on one side <strong>and</strong> anentangled state j ˜F N <strong>of</strong> yet unmeasured particles (subsetN , C ), on the other side. These unmeasured particlesdefine a “network” N corresponding to the shaded structurein Fig. 1. The state j ˜F N <strong>of</strong> the network is relatedto a cluster state jF N on N by a local unitary transformationwhich depends on the set <strong>of</strong> measurement resultsm. More specifically, j ˜F N satisfies Eqs. (1)—with Creplaced by the subcluster N —except for a possible differencein the sign factors, which are determined by themeasurement results m.To process quantum information with this network, itsuffices to measure its particles in a certain order <strong>and</strong> in acertain basis. Quantum information is thereby propagatedhorizontally through the cluster by measuring the qubitson the wire while qubits on vertical connections are usedto realize two-bit quantum gates. The basis in which acertain qubit is measured depends in general on the results<strong>of</strong> preceding measurements. The processing is finishedonce all qubits except the last one on each wire have beenmeasured. At this point, the results <strong>of</strong> previous measurementsdetermine in which basis these “output” qubits needto be measured for the final readout. We note that, in theentire process, only one-qubit measurements are required.The amount <strong>of</strong> entanglement therefore decreases with everymeasurement [8] <strong>and</strong> all entanglement involved in theprocess is provided by the initial resource, the cluster state.This is different from the scheme <strong>of</strong> Ref. [11], which usesBell measurements (capable <strong>of</strong> producing entanglement)to realize quantum gates.In the following, we show that any quantum logic circuitcan be implemented on a cluster state. The purpose <strong>of</strong> thisis tw<strong>of</strong>old. First, it serves as an illustration <strong>of</strong> how to implementa particular quantum circuit in practice. Second,in showing that any quantum circuit can be implementedon a sufficiently large cluster we demonstrate the universality<strong>of</strong> the proposed scheme. For pedagogical reasons wefirst explain a scheme with one essential modification withrespect to the proposed scheme: before the entanglementoperation S, certain qubits are selected as input qubits <strong>and</strong>the input information is written onto them, while the remainingqubits are prepared in j1. This step weakens thescheme since it affects the character <strong>of</strong> the cluster state asa genuine resource. It can, however, be avoided [see (e)].Points (a) to (c) are concerned with the basic elements <strong>of</strong> aquantum circuit, quantum gates, <strong>and</strong> wires, point (d) withthe composition <strong>of</strong> gates to circuits.(a) Information propagation in a wire for qubits. A qubitcan be teleported from one site <strong>of</strong> a cluster to any othersite. In particular, consider a chain <strong>of</strong> an odd number <strong>of</strong>qubits 1 to n prepared in the state jc in 1 ≠ j1 2 ≠ · · · ≠j1 n <strong>and</strong> subsequently entangled by S. The state that wasoriginally encoded in qubit 1, jc in , is now delocalized<strong>and</strong> can be transferred to site n by performing s x measurements(basis j1 j j0 x,j , j2 j j1 x,j ) at qubitsites j 1, . . . , n 2 1 with measurement outcomes s j [0, 1. The resulting state is js 1 x,1 ≠ · · · ≠ js n21 x,n21 ≠jc out n . The output state jc out is related to the input statejc in by a unitary transformation U S [ 1, s x , s z , s x s z which depends on the outcomes <strong>of</strong> the s x measurementsat sites 1 to n 2 1. A similar argument can be given for aneven number <strong>of</strong> qubits. The effect <strong>of</strong> U S can be accountedfor at the end <strong>of</strong> a computation as shown below [see (d)].It is noteworthy that not all classical information gainedby the s x measurements needs to be stored to identify thetransformation U S . Instead, U S is determined by the values<strong>of</strong> only two classical bits which are updated with everymeasurement.(b) An arbitrary rotation U R [ SU2 can be achievedin a chain <strong>of</strong> five qubits. Consider a rotation in itsEuler representation U R j, h, z U x zU z hU x j,where U x a exp2ia s x2 ,U za exp2ia s z2 . Initially,the first qubit is in some state jc in , whichis to be rotated, <strong>and</strong> the other qubits are in j1;i.e., their common state reads jC 1,...,5 jc in 1 ≠j1 2 ≠ j1 3 ≠ j1 4 ≠ j1 5 . After the five qubits areentangled by S they are in the state SjC 1,...,5 12jc in 1 j0 2 j2 3 j0 4 j2 5 2 12jc in 1 j0 2 j1 3 j1 4 j1 5 212jcin 1 j1 2 j1 3 j0 4 j2 5 1 12jcin 1 j1 2 j2 3 j1 4 j1 5 ,where jcin s z jc in . Now, the state jc in can be rotatedby measuring qubits 1 to 4, while it is teleported to site 5 atthe same time. The qubits 1, . . . , 4 are measured in appropriatelychosen bases B j a j j0 j1e ia jj1 p2 j, j0 j2e ia jj1 p2 jwhereby the measurement outcomes s j [ 0, 1 for j 1, . . . , 4 are obtained. Here, s j 0 means that qubit j isprojected into the first state <strong>of</strong> B j a j . The resultingstate is js 1 a1 ,1 ≠ js 2 a2 ,2 ≠ js 3 a3 ,3 ≠ js 4 a4 ,4 ≠ jc out 5with jc out Ujc in . For the choice a 1 0 (measurings x <strong>of</strong> qubit 1) the rotation U has the form U s s 21s 4x s s 11s 3z U R 21 s111 a 2 , 21 s 2 a 3, 21 s 11s 3a 4 . Insummary, the procedure to implement an arbitraryrotation U R j, h, z , specified by its Euler anglesj, h, z is (i) measure qubit 1 in B 1 0; (ii) measurequbit 2 in B 2 21 s111 j; (iii) measure qubit 3 inB 3 21 s 2 h; (iv) measure qubit 4 in B 4 21 s 11s 3z .In this way the rotation UR 0 is realized: URj, 0 h, z s s 21s 4x s s 11s 3s s 21s 4x s s 11s 3zz U R j, h, z . The extra rotation U S can be accounted for at the end <strong>of</strong> the computation,as is described below in (d).(c) To perform the gate CNOTc, t in ! t out j0 cc 0j ≠ 1 t in!t out 1 j1 cc 1j ≠ s t in!t out x between a controlqubit c <strong>and</strong> a target qubit t, four qubits, arrangedas depicted Fig. 2a, are required. During the action<strong>of</strong> the gate, the target qubit t is transferred from t into t out . The following procedure has to be implemented.Let qubit 4 be the control qubit. First, the stateji 1 z,1 ≠ ji 4 z,4 ≠ j1 2 ≠ j1 3 is prepared <strong>and</strong> then theentanglement operation S is performed. Second, s x <strong>of</strong>qubits 1 <strong>and</strong> 2 is measured. The measurement results5189


VOLUME 86, NUMBER 22 P H Y S I C A L R E V I E W L E T T E R S 28 MAY 2001(a)target in1 2 34controltarget out(b)target incontrol intarget outcontrol outFIG. 2. Realization <strong>of</strong> a CNOT gate by one-particle measurements.See text.s j [ 0, 1 correspond to projections <strong>of</strong> the qubits j intojs j x,j , j 1, 2. The quantum state created by this procedureis js 1 x,1 ≠ js 2 x,2 ≠ U 34S ji 4 z,4 ≠ ji 1 1 i 4 mod2 z,3 ,where U 34S s 3s 111. The input state is thusz s 3s 2xs 4s 1zacted upon by the CNOT <strong>and</strong> successive s x <strong>and</strong> s z rotationsU 34S , depending on the measurement results s 1 , s 2 .These unwanted extra rotations can again be accountedfor as described in (d). For practical purposes it is moreconvenient if the control qubit is, as the target qubit,transferred to another site during the action <strong>of</strong> the gate.When a CNOT is combined with other gates to form aquantum circuit it will be used in the form shown inFig. 2b.To explain the working principle <strong>of</strong> the CNOT gate we,for simplicity, refer to the minimal implementation withfour qubits. The minimal CNOT can be viewed as a wirefrom qubit 1 to qubit 3 with an additional qubit, No. 4,attached. From the eigenvalue equations (1) it can now bederived that, if qubit 4 is in an eigenstate ji 4 z,4 <strong>of</strong> s z , thenthe value <strong>of</strong> i 4 [ 0, 1 determines whether a unit wire ora spin flip s x (modulo the same correction U 3S for bothvalues <strong>of</strong> i 4 ) is being implemented. In other words, onces x <strong>of</strong> qubits 1 <strong>and</strong> 2 have been measured, the value i 4 <strong>of</strong>qubit 4 controls whether the target qubit is flipped or not.(d) Quantum circuits. The gates described —the CNOT<strong>and</strong> arbitrary one-qubit rotations — form a universal set[5]. In the implementation <strong>of</strong> a quantum circuit on a clusterstate the site <strong>of</strong> every output qubit <strong>of</strong> a gate overlapswith the site <strong>of</strong> an input qubit <strong>of</strong> a subsequent gate. Because<strong>of</strong> this, the entire entanglement operation can beperformed at the beginning. To see this, compare thefollowing two strategies. Given a quantum circuit implementedon a network N <strong>of</strong> qubits which is dividedinto two consecutive circuits, circuit 1 is implemented onnetwork N 1 <strong>and</strong> circuit 2 is implemented on networkN 2 , <strong>and</strong> N N 1 < N 2 . There is an overlap O N 1 > N 2 which contains the sites <strong>of</strong> the output qubits<strong>of</strong> circuit 1 (these are identical to the sites <strong>of</strong> the inputqubits <strong>of</strong> circuit 2). The sites <strong>of</strong> the readout qubits form aset R , N 2 . Strategy (i) consists <strong>of</strong> the following steps:(1) write input <strong>and</strong> entangle all qubits on N ; (2) measurequbits [ N nR to implement the circuit. Strategy(ii) consists <strong>of</strong> (1) write input <strong>and</strong> entangle the qubits onN 1 , (2) measure the qubits in N 1 nO . This implementsthe circuit on N 1 <strong>and</strong> writes the intermediate output toO ; (3) entangle the qubits on N 2 ; (4) measure all qubitsin N 2 nR. Steps 3 <strong>and</strong> 4 implement the circuit 2 on N 2 .The measurements on N 1 nO commute with the entanglementoperation restricted to N 2 , since they act on differentsubsets <strong>of</strong> particles. Therefore the two strategies aremathematically equivalent <strong>and</strong> yield the same results. Itis therefore consistent to entangle in a single step at thebeginning <strong>and</strong> perform all measurements afterwards.Two further points should be addressed in connectionwith circuits. First, the r<strong>and</strong>omness <strong>of</strong> the measurementresults does not jeopardize the function <strong>of</strong> the circuit.Depending on the measurement results, extra rotationss x <strong>and</strong> s z act on the output qubit <strong>of</strong> every implementedgate. By use <strong>of</strong> the relations U R j, h, z szs s xs0 szs s x s0 U R 21 s j, 21 s0 h, 21 s z , <strong>and</strong> CNOTc, ts ts tzs cs cz s ts0 txs cs0 cx s ts tz s cs c1s t c1sz s 0 ts0 txs cs0 cx CNOTc, t,these extra rotations can be pulled through the network toact upon the output state. There they can be accountedfor by adjusting the measurement basis for the finalreadout. The above relations imply that for a rotationU R j, h, z—different from the CNOT gate—the accumulatedextra rotations U S at the input side <strong>of</strong> U R need tobe determined before the measurement bases that realizeU R can be specified. This introduces a partial temporalordering <strong>of</strong> the measurements on the whole cluster.Second, quantum circuits can also be implemented onirregular clusters. In that case, qubits may be missingwhich are required for the st<strong>and</strong>ard implementation <strong>of</strong> thecircuit. This can be compensated by a large flexibility inshape <strong>of</strong> the gates <strong>and</strong> wires. The components can be bent<strong>and</strong> stretched to fit to the cluster structure as long as thetopology <strong>of</strong> the circuit implementation does not change.Irregular clusters are found in lattices with a finite siteoccupation probability 0 , p , 1. In such a situation,the possibility <strong>of</strong> universal quantum computation isclosely linked to the phenomenon <strong>of</strong> percolation. For pabove a certain critical value p c , which depends on thedimension <strong>of</strong> the lattice, an infinitely extended clusterexists that may be used as the carrier <strong>of</strong> the quantumcircuit. In two dimensions, for example, exactly onesuch cluster C exists. Suppose this cluster is dividedinto two subclusters C 1 <strong>and</strong> C 2 by a one-dimensional cutO C 1 > C 2 . It can be shown, e.g., by using Russo’sformula [12] from percolation theory that, for any cut O ,jO j `. Therefore there is no upper bound, in principle,to the “capacity” <strong>of</strong> the cluster, i.e., to the number <strong>of</strong>qubits that can be processed across such a cut.(e) Full scheme. It is important to note that the step<strong>of</strong> writing the input information onto the qubits beforethe cluster is entangled was introduced only for pedagogicalreasons. For illustration <strong>of</strong> this point considera chain <strong>of</strong> five qubits in the state Sj1 1 ≠ j1 2 ≠ · · · ≠j1 5 . Clearly, there is no local information on any <strong>of</strong> thequbits. However, by measuring qubits 1 to 4 along suitabledirections, qubit 5 can be projected into any desired state(modulo U S ). What is used here is the knowledge that the5190


VOLUME 86, NUMBER 22 P H Y S I C A L R E V I E W L E T T E R S 28 MAY 2001resource has been prepared with qubit 1 in the state j1 1before the entanglement operation. By the four measurements,this qubit is then rotated as described in (b). Inorder to use qubit 5 for further processing, the five-qubitchain considered here should, <strong>of</strong> course, be part <strong>of</strong> a largercluster such that particle 5 is still entangled with the remainingnetwork, after particles 1 to 4 have been measured.The method <strong>of</strong> preparing the input state remains thesame, in this case, as explained in (d). In a similar mannerany desired input state can be prepared if the rotationsare replaced by a circuit preceding the proper circuit forcomputation. In summary, no input information needs tobe written to the qubits before they are entangled. Clusterstates are thus a genuine resource for quantum computationvia measurements only.For a cluster <strong>of</strong> a given finite size, the number <strong>of</strong> computationalsteps may be too large to fit on the cluster. In thiscase, the computation can be split into consecutive parts,for each <strong>of</strong> which there is sufficient space on the cluster.The modified procedure consists then <strong>of</strong> repeatedly (re)entanglingthe cluster <strong>and</strong> imprinting the actual part <strong>of</strong> thecircuit —by measuring all <strong>of</strong> the lattice qubits exceptthe ones carrying the intermediate quantum output — untilthe whole calculation is performed. This procedure hasalso the virtue that qubits involved in the later part <strong>of</strong> acalculation need not be protected from decoherence for along time while the calculation is still being performed ata remote place <strong>of</strong> the cluster. St<strong>and</strong>ard error-correctiontechniques [13,14] may then be used on each part <strong>of</strong> thecircuit to stabilize the computation against decoherence.A possible implementation <strong>of</strong> such a quantum computeruses neutral atoms stored in periodic micropotentials[15–18] where Ising-type interactions can be realizedby controlled collisions between atoms in neighboringpotential wells [16,18]. This system combines smalldecoherence rates with a high scalability. The question<strong>of</strong> scalability is linked to the percolation phenomenon,as mentioned earlier. For a site occupation probabilityabove the percolation threshold, there exists a clusterwhich is bounded in size only by the trap dimensions.For optical lattices in three dimensions, single-atom siteoccupation with a filling factor <strong>of</strong> 0.44 has been reported[19] which is significantly above the percolation threshold<strong>of</strong> 0.31 [20]. As in other proposed implementations forquantum computing, the addressability <strong>of</strong> single qubitsin the lattice is, however, still a problem. (For recentprogress, see Ref. [21]). Recently, it has also been shownthat implementations based on arrays <strong>of</strong> capacitively coupledquantum dots may be used to realize an Ising-typeinteraction [22].In conclusion, we have described a new scheme <strong>of</strong>quantum computation that consists entirely <strong>of</strong> one-qubitmeasurements on a particular class <strong>of</strong> entangled states,the cluster states. The measurements are used to imprint aquantum circuit on the state, thereby destroying its entanglementat the same time. Cluster states are thus one-wayquantum computers <strong>and</strong> the measurements form theprogram.We thank D. E. Browne, D. P. DiVincenzo, A. Schenzle,<strong>and</strong> H. Wagner for helpful discussions. This work wassupported by the Deutsche Forschungsgemeinschaft.[1] C. H. Bennett <strong>and</strong> D. P. DiVincenzo, Nature (London) 404,247 (2000).[2] See Ref. [1] for a recent review.[3] J. I. Cirac <strong>and</strong> P. Zoller, Nature (London) 404, 579 (2000).[4] D. Deutsch, Proc. R. Soc. London 425, 73 (1989).[5] A. Barenco et al., Phys. Rev. A 52, 3457 (1995).[6] H.-J. Briegel <strong>and</strong> R. <strong>Raussendorf</strong>, Phys. Rev. Lett. 86, 910(2001).[7] The second Hamiltonian is <strong>of</strong> the st<strong>and</strong>ard Ising form. Thesymbol “” means that the states generated from a giveninitial state, under the action <strong>of</strong> these Hamiltonians, areidentical up to a local rotation on certain qubits. We usethe first Hamiltonian to make the computational schememore transparent. The conclusions drawn in the paper are,however, the same for both Hamiltonians.[8] By the “amount <strong>of</strong> entanglement” contained in the resource,we mean any measure that satisfies the criteria <strong>of</strong>an entanglement monotone [9]. For cluster states, the entanglementcan be calculated, e.g., in terms <strong>of</strong> the Schmidtmeasure <strong>of</strong> Ref. [10].[9] G. Vidal, J. Mod. Opt. 47, 355 (2000).[10] J. Eisert <strong>and</strong> H.-J. Briegel, quant-ph/0007081.[11] D. Gottesman <strong>and</strong> I. L. Chuang, Nature (London) 402, 390(1999).[12] See, e.g., G. Grimmett, Percolation (Springer-Verlag, NewYork, 1989).[13] A. Calderbank <strong>and</strong> P. W. Shor, Phys. Rev. A 54, 1098(1996).[14] A. Steane, Phys. Rev. Lett. 77, 793 (1996).[15] G. K. Brennen et al., Phys. Rev. Lett. 82, 1060 (1999).[16] D. Jaksch et al., Phys. Rev. Lett. 82, 1975 (1999).[17] T. Calarco et al., Phys. Rev. A 61, 022304 (2000).[18] H.-J. Briegel et al., J. Mod. Opt. 47, 415 (2000).[19] M. T. DePue et al., Phys. Rev. Lett. 82, 2262 (1999).[20] J. M. Ziman, Models <strong>of</strong> Disorder (Cambridge UniversityPress, Cambridge, United Kingdom, 1979).[21] R. Scheunemann et al., Phys. Rev. A 62, 051801(R) (2000).[22] T. Tanamoto, quant-ph/0009030.5191


LETTERSPUBLISHED ONLINE: 20 JULY 2009 | DOI: 10.1038/NPHYS1342Quantum computation <strong>and</strong> quantum-stateengineering driven by dissipationFrank Verstraete 1 *, Michael M. Wolf 2 <strong>and</strong> J. Ignacio Cirac 3 *The strongest adversary in quantum information science isdecoherence, which arises owing to the coupling <strong>of</strong> a systemwith its environment 1 . The induced dissipation tends to destroy<strong>and</strong> wash out the interesting quantum effects that give riseto the power <strong>of</strong> quantum computation 2 , cryptography 2 <strong>and</strong>simulation 3 . Whereas such a statement is true for manyforms <strong>of</strong> dissipation, we show here that dissipation can alsohave exactly the opposite effect: it can be a fully fledgedresource for universal quantum computation without anycoherent dynamics needed to complement it. The coupling tothe environment drives the system to a steady state wherethe outcome <strong>of</strong> the computation is encoded. In a similarvein, we show that dissipation can be used to engineer alarge variety <strong>of</strong> strongly correlated states in steady state,including all stabilizer codes, matrix product states 4 , <strong>and</strong> theirgeneralization to higher dimensions 5 .The situation we have in mind is shown in Fig. 1. A quantumsystem composed <strong>of</strong> N particles (such as qubits) is organized inspace according to a particular geometry (in the figure, a onedimensionallattice). Neighbouring systems are coupled to somelocal environments, which are dissipative in nature <strong>and</strong> tend todrive the system to a steady state. Our idea is to engineer thosecouplings, so that the environments drive the system to a desiredfinal state. The coupling to the environment will be static, so that thedesired state is obtained after some time without having to activelycontrol the system. Note that the role <strong>of</strong> the environments is todissipate (or, more precisely, evacuate) the entropy <strong>of</strong> the system,<strong>and</strong> by choosing the couplings appropriately we can use this effectto drive our system.We will show first how to design the interactions withthe environment to implement universal quantum computation.This new method, which we refer to as dissipative quantumcomputation (DQC), defies some <strong>of</strong> the st<strong>and</strong>ard criteria forquantum computation because it requires neither state preparation,nor unitary dynamics 6 . However, it is nevertheless as powerful asst<strong>and</strong>ard quantum computation. Then we will show that dissipationcan be engineered 7 to prepare ground states <strong>of</strong> frustration-freeHamiltonians. Those include matrix product states 4,8,9 (MPSs) <strong>and</strong>projected entangled pair states 5,9 (PEPSs), such as graph states 10<strong>and</strong> Kitaev 11 <strong>and</strong> Levin–Wen 12 topological codes. Both DQC <strong>and</strong>dissipative state engineering (DSE) are robust in the sense that,given the dissipative nature <strong>of</strong> the process, the system is driventowards its steady state independent <strong>of</strong> the initial state <strong>and</strong> hence<strong>of</strong> eventual perturbations along the way.Here, we will concentrate first on DQC, showing how givenany quantum circuit one can construct a locally acting masterequation for which the steady state is unique, encodes the outcome<strong>of</strong> the circuit <strong>and</strong> is reached in polynomial time (with respect tothe one corresponding to the circuit). Then we will show howto construct dissipative processes that drive the system to theground state <strong>of</strong> any frustration-free Hamiltonian. In the Methodssection, we will prove that MPS (ref. 9) <strong>and</strong> certain kinds <strong>of</strong>PEPS (ref. 9) can be efficiently prepared using this method, <strong>and</strong>in Supplementary Information we will give details <strong>of</strong> the pro<strong>of</strong>s.In this letter we will not consider specific physical set-ups whereour ideas can be implemented. Nevertheless, the Methods sectionwill provide a universal way <strong>of</strong> engineering the master equationsrequired for DQC <strong>and</strong> DSE, which can be easily adapted to currentexperiments 13 based on, for example, atoms in optical lattices 14or trapped ions 15 . Thus, we expect that our predictions may beexperimentally tested in the near future.Let us start with DQC by considering N qubits in a line <strong>and</strong> aquantum circuit specified by a sequence <strong>of</strong> nearest-neighbour qubitoperations {U t } T t=1 . We define |ψ t 〉 := U t U t−1 ...U 1 |0〉 1 ⊗...|0〉 N , sothat |ψ T 〉 is the final state after the computation. Our goal is to finda master equation ˙ρ =L(ρ) with a Liouvillian in Lindblad form 16L(ρ) = ∑ kL k ρL † k − 1 2{L†kL k ,ρ } +(1)where the L k acts locally <strong>and</strong> has a steady state, ρ 0 : (1) that is unique;(2) that can be reached in a time poly(T); (3) such that ψ T can beextracted from it in a time poly(T). As in Feynman’s construction<strong>of</strong> a quantum simulator 3 , we consider another auxiliary registerwith states {|t〉} T t=0, which will represent the time. We choosethe Lindblad operatorsL i = |0〉 i 〈1|⊗|0〉 t 〈0|L t = U t ⊗|t +1〉〈t|+U †t⊗|t〉〈t +1|where i = 1,...,N <strong>and</strong> t = 0,...,T. It is clear that the L terms actlocally except for the interaction with the extra register, which canbe made local as well. Furthermore,ρ 0 = 1 ∑|ψ t 〉〈ψ t |⊗|t〉〈t|T +1tis a steady state, that is, L(ρ 0 )=0. Given such a state, the result <strong>of</strong> theactual quantum computation can be read out with probability 1/Tby measuring the time register. In Supplementary Information, weshow that ρ 0 is the unique steady state <strong>and</strong> that the Liouvillian hasa spectral gap = π 2 /(2T +3) 2 . This means indeed that the steadystate will be reached in polynomial time in T. Note that this gap isindependent <strong>of</strong> N as well as <strong>of</strong> the actual quantum computation thatis carried out (that is, independent <strong>of</strong> the U t ). It is also shown thatthe same gap is retained if the clock register is encoded in the unary1 Fakultät für Physik, Universität Wien, 1090 Wien, Austria, 2 Niels Bohr <strong>Institute</strong>, 2100 Copenhagen, Denmark, 3 Max-Planck-Institut für Quantenoptik,85748 Garching, Germany. *e-mail: fverstraete@gmail.com; ignacio.cirac@mpq.mpg.de.NATURE PHYSICS | VOL 5 | SEPTEMBER 2009 | www.nature.com/naturephysics 633© 2009 Macmillan Publishers Limited. All rights reserved.


LETTERSNATURE PHYSICS DOI: 10.1038/NPHYS1342Figure 1 | Schematic representation <strong>of</strong> the set-up. We consider a collection <strong>of</strong> N quantum particles, locally coupled to a set <strong>of</strong> environments. Thecouplings are engineered in such a way that the system reaches the desired state in the long-time limit.way proposed by Kitaev <strong>and</strong> co-workers 17 , making the Lindbladoperators strictly local. A sketch <strong>of</strong> the pro<strong>of</strong> is as follows. First, wedo a similarity transformation on L that replaces all gates U i with theidentity gates, showing that its spectrum is independent <strong>of</strong> the actualquantum computation. Second, another similarity transformationis done that makes L Hermitian <strong>and</strong> block-diagonal. Each block canthen be diagonalized exactly leading to the claimed gap.In some sense, the present formalism can be seen as a robustway <strong>of</strong> doing adiabatic quantum computation 18 (errors do notaccumulate <strong>and</strong> the path does not have to be engineered carefully)<strong>and</strong> implementing quantum r<strong>and</strong>om walks 19 , <strong>and</strong> it might thereforebe easier to tackle interesting open questions, such as the quantumprobabilistically-checkable-pro<strong>of</strong>s theorem, in this setting 20 . Inaddition, it seems that the dissipative way <strong>of</strong> preparing ground statesis more natural than to use adiabatic time evolution, as nature itselfprepares them by cooling.Let us now turn to DSE <strong>and</strong> consider again a quantum systemwith N particles on a lattice in any dimension. We are interested inground states Ψ, <strong>of</strong> HamiltoniansH = ∑ λthat are frustration-free, meaning that Ψ minimizes the energy <strong>of</strong>each H λ individually, <strong>and</strong> local in the sense that H λ acts non-triviallyonly on a small set λ ⊂ {1,...,N } <strong>of</strong> sites (for example, nearestneighbours). We can assume the terms H λ to be projectors <strong>and</strong>we will denote the orthogonal projectors by P λ = 1 − H λ . States Ψ<strong>of</strong> the considered form are, for example, all PEPS (including MPS<strong>and</strong> stabilizer states 21 ).We will consider discrete time evolution generated by a tracepreservingcompletely positive map instead <strong>of</strong> a master equation.These two approaches are basically equivalent 22 as every localcompletely positive map T can be associated with a local Liouvillianthrough L(ρ) = N [T (ρ)−ρ], which leads to the same fixed points<strong>and</strong> spectrum. We choose completely positive maps <strong>of</strong> the formT (ρ) = ∑ λH λ[]p λ P λ ρP λ + 1 m∑U λ,i H λ ρH λ U † λ,imwhere the p λ terms are probabilities <strong>and</strong> U λ,1 ,...,U λ,m is a set <strong>of</strong>unitaries acting non-trivially only within region λ. They effectivelyrotate part <strong>of</strong> the high-energy space (with support <strong>of</strong> H λ ) to thezero-energy space, so that tr[T (ρ)Ψ] ≥ tr[ρΨ] increases. As forLiouvillians (1), we could similarly take L λ,i = U i H λ , or the onesassociated with the completely positive map.We show now that for every frustration-free Hamiltonian,the completely positive map in equation (2) converges to theground-state space if we choose the unitaries U λ,i to be completelydepolarizing, that is, T (ρ) ∝ ∑ λ P λρP λ + 1 λ ⊗ tr λ [H λ ρ]/tr[1 λ ].For ease <strong>of</strong> notation, we will explain the pro<strong>of</strong> for the case <strong>of</strong> ai=1(2)one-dimensional ring with nearest-neighbour interactions labelledby the first site λ = 1,...,N . Assume ρ is such that its expectationvalue with respect to the projector Ψ onto the ground-state space<strong>of</strong> H is non-increasing under applications <strong>of</strong> T , that is, in particulartr[ρΨ] = tr[T N (ρ)Ψ]. Expressing this in the Heisenberg picture inwhich T ∗ (Ψ) = Ψ + ∑ λ H λtr λ (Ψ)/(d 2 N ), we get[]1N∑ N∏tr[ρΨ] ≥ tr[ρΨ]+(d 2 N ) tr ( )ρ Hλ+µ tr N λ+µ (Ψ)νN≥ tr[ρΨ]+(d 2 N ) tr[ρH]Nµ=1 λ=1where the first inequality comes from discarding (positive) terms inthe sum <strong>and</strong> the second one is due to bounding all partial traces<strong>of</strong> H λ from below by the respective smallest eigenvalue ν. Notethat the latter is strictly positive unless H has a product state asthe ground state (in which case the statement becomes trivial).Hence, we must have tr[ρH] = 0; that is, ρ is a ground state <strong>of</strong> H.It is easily seen that the same argument applies for more generalinteractions on arbitrary lattices.Once we have shown that the steady state after the application<strong>of</strong> the completely positive map lies within the desired subspace(the ground-state space <strong>of</strong> the frustration-free Hamiltonian), thenext question to be addressed is how efficient the process is. Thisdepends on the spectral gap, δ, <strong>of</strong> the completely positive map (or,equivalently, <strong>of</strong> the corresponding Liouvillian), as the time to reachthe steady state, τ = O(1/δ). Thus, the above procedure will beefficient as long as the gap vanishes only polynomially with thenumber <strong>of</strong> systems, N . Similarly to what occurs with many-bodyHamiltonians, the determination <strong>of</strong> such a gap is, in general, verycomplicated. For a wide range <strong>of</strong> interesting models, however, itcan be proved that this gap scales favourably. This is the case forall MPS as well as for a rich subfamily <strong>of</strong> PEPS that includes allstabilizer states (such as Kitaev’s toric code 11 <strong>and</strong> the Levin–Wenstates 12 ). In the Methods section, we characterize such a subfamily<strong>of</strong> states, <strong>and</strong> in Supplementary Information we give the technicalpro<strong>of</strong>s <strong>of</strong> our statements. Here, we will qualitatively explain howour method works efficiently for some families <strong>of</strong> states. For thatwe note that the action <strong>of</strong> the completely positive map (2) canbe interpreted as r<strong>and</strong>omly choosing a region λ (according to p λ ,which we may set equal to 1/N ), then measuring P λ <strong>and</strong> applyinga correction according to the unitaries if the outcome was negative.We denote by R n the set <strong>of</strong> regions λ where ϕ satisfies the conditionH λ |ϕ〉 = 0. If we measure now in one <strong>of</strong> those regions, we willobviously obtain a positive result, <strong>and</strong> thus R n will remain thesame. If we measure in another region, we may have a positive ornegative result, something that may change the set R n . By imposingcertain conditions on the operators H λ <strong>and</strong> U λ,i , we can make surethat in each step R n cannot be reduced <strong>and</strong> that the probability <strong>of</strong>634 NATURE PHYSICS | VOL 5 | SEPTEMBER 2009 | www.nature.com/naturephysics© 2009 Macmillan Publishers Limited. All rights reserved.


NATURE PHYSICS DOI: 10.1038/NPHYS1342being enlarged is non-vanishing. This automatically ensures thatthe τ scales only polynomially with the number <strong>of</strong> systems. Inone dimension, however, one can get rid <strong>of</strong> all those restrictions<strong>and</strong> show that any MPS can be prepared in a time that also scalesfavourably with N . The fact that all MPS states can be preparedwith our method, together with the results reported in refs 23, 24,automatically implies the existence <strong>of</strong> phase transitions driven bydissipation in the following sense. By changing the parameters <strong>of</strong>the operators H λ appearing in the completely positive map (2), wechange the steady state <strong>of</strong> that map. It is possible to choose modelsfor which that state changes abruptly at some particular value <strong>of</strong>that parameter in such a way that the correlation length diverges<strong>and</strong> an order parameter appears (an example can be found in theSupplementary Information).We have investigated the computational power <strong>of</strong> purelydissipative processes, <strong>and</strong> proved that it is equivalent to that <strong>of</strong>the quantum circuit model <strong>of</strong> quantum computation. We havealso shown that dissipative dynamics can be used to create groundstates (such as MPS or PEPS) <strong>of</strong> frustration-free Hamiltonians <strong>of</strong>strongly correlated quantum spin systems. We believe that thesenew methods can be experimentally tested using atoms or ions withcurrent set-ups (see the Methods section).Let us stress that we have been concerned here with a pro<strong>of</strong><strong>of</strong>-principledemonstration that dissipation provides us with analternative way <strong>of</strong> carrying out quantum computations or stateengineering. We believe, however, that much more efficient <strong>and</strong>practical schemes can be developed <strong>and</strong> adapted to specificimplementations. We also think that these results open upsome interesting questions that deserve further investigation: forexample, how the use <strong>of</strong> fault-tolerant computations can makeour scheme more robust, or how one can design translationallyinvariant completely positive maps that prepare MPS moreefficiently, or the importance <strong>and</strong> generality <strong>of</strong> the set <strong>of</strong> commutingHamiltonians (see the Methods section), which is intimatelyconnected to the fixed points <strong>of</strong> the renormalization grouptransformations on PEPS (as it happens with MPS; ref. 25).Furthermore, the model <strong>of</strong> DQC might well lead to the construction<strong>of</strong> new quantum algorithms, as, for example, quantum r<strong>and</strong>omwalks can more easily be formulated within this context. Finally,other ideas related to this work can be easily addressed using themethods introduced; for example, thermal states <strong>of</strong> commutingHamiltonians can be engineered using DSE because the Metropolisway <strong>of</strong> sampling over classical spin configurations can be adoptedto the case <strong>of</strong> commuting operators. Similar techniques could beapplied to free fermionic <strong>and</strong> bosonic systems, <strong>and</strong>, more generally,it should be possible to devise DSE schemes converging to theground or thermal states <strong>of</strong> frustrated Hamiltonians by combiningunitary <strong>and</strong> dissipative dynamics.Note added. Concurrently with the submission <strong>of</strong> this paper,refs 26 <strong>and</strong> 27 appeared in which a similar quantum-reservoirengineering was used to prepare many-body states <strong>and</strong> nonequilibriumquantum phases.MethodsEngineering dissipation. Here we show how to engineer the local dissipation thatgives rise to the master equations (1) <strong>and</strong> completely positive maps (2). They arecomposed <strong>of</strong> local terms, involving few particles (typically two), so that we just haveto show how to implement those. To simplify the exposition, we will treat thoseparticles as a single one <strong>and</strong> assume that one has full control over its dynamics (forexample, one can apply arbitrary gates).Let us start with the completely positive maps. It is clear that by applying aquantum gate to the particle <strong>and</strong> a ‘fresh’ ancilla <strong>and</strong> then tracing the ancilla onecan generate any physical action (that is, completely positive map) on the system.Furthermore, by repeating the same process with short time intervals one cansubject the system to an arbitrary time-independent master equation. This lastprocess may not be efficient. An alternative way works as follows. Let us assumethat the ancilla is a qubit interacting with a reservoir such that it fulfils a masterLETTERSequation with Liouville operator L a = √ Γ σ − , where σ − = |0〉〈1|. Now, we couplethe ancilla to the system with a Hamiltonian H = (σ − L † + σ † −L). In the limitΓ ≫ , one can adiabatically eliminate the level |1〉 <strong>of</strong> the ancilla 28 by applyingsecond-order perturbation theory to the Liouvillian (albeit for non-Hermitianoperators). In this way we obtain an effective master equation for ρ describingthe system alone, with Liouville operator / √ Γ L. By using several ancillaswith Hamiltonians H = (σ − L i + σ † −L † i ) <strong>and</strong> following the same procedure weobtain the desired master equation. Although we have not specified here a physicalsystem, one could use atoms. In that case, the ancilla could be an atom itself with|0〉 <strong>and</strong> |1〉 an electronic ground <strong>and</strong> excited level, respectively, so that spontaneousemission gives rise to the dissipation. The coupling to the system (other atoms)could be achieved using st<strong>and</strong>ard ideas used in the implementation <strong>of</strong> quantumcomputation using those systems 13 .Efficient state preparation. We have shown that it is possible to engineerdissipative processes that prepare ground states <strong>of</strong> frustration-free Hamiltonians insteady state. In the pro<strong>of</strong>, the time for this preparation scales as N N , which may bean issue for experiments with large number <strong>of</strong> particles. Here we give much moreefficient methods for certain classes <strong>of</strong> frustration-free Hamiltonians.We consider first frustration-free Hamiltonians for which [H λ ,H µ ] = 0 <strong>and</strong>show that, under certain conditions, the corresponding ground states can beprepared in a time that scales only polynomially with the number <strong>of</strong> particles. Thecorresponding set <strong>of</strong> ground states contains important families, such as stabilizerstates (for example, cluster states <strong>and</strong> topological codes), or certain kinds <strong>of</strong> PEPS,namely, those that have (commuting) parent Hamiltonians with the injectivitycondition (as defined in refs 8, 29). Note that there was no known way <strong>of</strong> efficientpreparation for the latter.Loosely speaking, we will consider two classes <strong>of</strong> Hamiltonians.(1) Hamiltonians for which all excitations can be locally annihilated. In this case thetime <strong>of</strong> convergence scales as τ = O(logN ). (2) Interactions where excitations haveto be moved along the lattice before they can annihilate <strong>and</strong> τ =O(N logN ).To see how the first case can occur notice that, when iterating T , thecorrection on λ does not change the outcome <strong>of</strong> previous measurements onneighbouring regions because∀λ ≠ λ ′ : [U λ,i ,H λ ′ ] = 0 (3)In fact, this can always be achieved by regrouping the regions into larger oneshaving an interior I(λ) ⊂ λ on which only H λ acts non-trivially <strong>and</strong> letting theU λ,i solely act on I(λ). Denote by q the largest probability for obtaining twice anegative measurement outcome on the same region λ. The energy tr[HT M (ρ)]after M applications <strong>of</strong> T decreases then as N (1−(1−q)/N ) M such that it takesO((N logN )/(1−q)) steps to converge to a ground state. The relaxation time <strong>of</strong> thecorresponding Liouvillian is thus τ = O(logN 1/1−q ). Clearly, this is a reasonablebound only if q


LETTERS‘parent’ Hamiltonian that has a gap. Denoting by ρ the reduced density operatorcorresponding to particles k <strong>and</strong> k + 1, H k <strong>and</strong> P k = 1 − H k will denote theprojectors onto its kernel <strong>and</strong> range, respectively. Note that tr(P k ) = D 2 . Wetake N = 2 n for simplicity, but this is clearly not necessary. We construct thechannel T in several steps. We first define a channel acting on two neighbouringparticles k,k +1, as followsR r,c (X) := P k XP k + P kD 2 tr(H kX)Here, k = 2 r−1 (2c −1), where r = 1,...,n <strong>and</strong> c = 1,...,2 n−r . The action <strong>of</strong> thesemaps has a tree structure, where the index r indicates the row in the tree, whereas cdoes it for the column. Now we define recursively,S r,c := (1−ɛ r )(S r−1,2c +S r−1,2c+1 )+ɛ r R r,c2Here, r = 2,...,n, c = 1,...,2 n−r , S 1,c := R 1,c <strong>and</strong> ɛ r+1 = 1/M r , where M = CN 2<strong>and</strong> C ≫ 1 (see Supplementary Information). Note that S r,1 acts on the first 2 rparticles, S r,2 on the next 2 r <strong>and</strong> so on. We finally defineT := (1−ɛ n+1 )S n,1 +ɛ n+1 R n,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(N log 2 (N ) ). The intuitionbehind the completely positive map (4) is that the channels S 1,c , which are the onesthat most <strong>of</strong>ten applied, project the state <strong>of</strong> every second nearest neighbour ontothe right subspace. Then S 2,c do the same with half <strong>of</strong> the pairs that have not beenprojected. Then S 3,c does the same on half <strong>of</strong> the rest, <strong>and</strong> so on.Received 11 March 2008; accepted 18 June 2009; published online20 July 2009References1. Aliferis, P., Gottesman, D. & Preskill, J. Quantum accuracy threshold forconcatenated distance-3 codes. Quant. Inf. Comput. 6, 97–165 (2006).2. Nielsen, M. A. & Chuang, I. L. Quantum Computation <strong>and</strong> QuantumInformation (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 onquantum spin chains. Commun. Math. Phys. 144, 443–490 (1992).5. Verstraete, F. & Cirac, J. I. Renormalization algorithms forquantum-many body systems in two <strong>and</strong> higher dimensions. Preprint at (2004).6. DiVincenzo, D. P. The physical implementation <strong>of</strong> quantum computation.Fortschr. Phys. 48, 771–783 (2000).7. Poyatos, J. F., Cirac, J. I. & Zoller, P. Quantum Reservoir Engineering withlaser cooled trapped ions. Phys. Rev. Lett. 77, 4728–4731 (1996).8. Perez-Garcia, D., Verstraete, F., Wolf, M. M. & Cirac, J. I. Matrix product staterepresentations. Quant. Inf. Comput. 7, 401–430 (2007).9. Verstraete, F., Murg, V. & Cirac, J. I. Matrix product states, projected entangledpair states, <strong>and</strong> variational renormalization group methods for quantum spinsystems. Adv. Phys. 57, 143–224 (2008).NATURE PHYSICS DOI: 10.1038/NPHYS134210. Briegel, H. J. & <strong>Raussendorf</strong>, R. Persistent entanglement in arrays <strong>of</strong> interactingqubits. 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 fortopological phases. Phys. Rev. B 71, 045110 (2005).13. Cirac, J. I. & Zoller, P. New frontiers in quantum information with atoms <strong>and</strong>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. & Winel<strong>and</strong>, D. Quantum dynamics <strong>of</strong>single trapped ions. Rev. Mod. Phys. 75, 281–324 (2003).16. Lindblad, G. On the generators <strong>of</strong> quantum dynamical semigroups.Commun. Math. Phys. 48, 119–130 (1976).17. Kempe, J., Kitaev, A. Y. & Regev, O. The complexity <strong>of</strong> the local Hamiltonianproblem. SIAM J. Comput. 35, 1070–1097 (2004).18. Aharonov, D. et al. Adiabatic quantum computation is equivalent to st<strong>and</strong>ardquantum computation. SIAM J. Comput. 37, 166–194 (2007).19. Kempe, J. Quantum r<strong>and</strong>om walks—an introductory overview. Contemp. Phys.44, 307–327 (2003).20. Arora, S. & Safra, S. Probabilistic checking <strong>of</strong> pro<strong>of</strong>s: A new characterization <strong>of</strong>NP. J. ACM 45, 70–122 (1998).21. Gottesman, D. A theory <strong>of</strong> fault-tolerant quantum computation. Phys. Rev. A57, 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 transitionsin matrix product systems. Phys. Rev. Lett. 97, 110403 (2006).24. Verstraete, F., Wolf, M. M., Perez-Garcia, D. & Cirac, J. I. Criticality, the arealaw, <strong>and</strong> the computational power <strong>of</strong> 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 <strong>and</strong> phases in driven open quantum systemswith cold atoms. Nature Phys. 4, 878–883 (2008).27. Kraus, B. et al. Preparation <strong>of</strong> entangled states by quantum Markov processes.Phys. Rev. A 78, 042307 (2008).28. Cohen-Tannoudji, C., Dupont-Roc, J. & Grynberg, G. Atom-PhotonInteractions (Wiley, 1992).29. Perez-Garcia, D., Verstraete, F., Cirac, J. I. & Wolf, M. M. PEPS as uniqueground states <strong>of</strong> local Hamiltonians. Quant. Inf. Comput. 8, 0650–0663 (2008).AcknowledgementsWe thank D. Perez-Garcia for discussions <strong>and</strong> acknowledge financial support bythe EU projects QUEVADIS, SCALA, the FWF, QUANTOP, FNU, SFB FoQuS, theDFG, Forschungsgruppe 635, the Munich Center for Advanced Photonics (MAP)<strong>and</strong> Caixa Manresa.Author contributionsAll authors have contributed equally to this paper.Additional informationSupplementary information accompanies this paper on www.nature.com/naturephysics.Reprints <strong>and</strong> permissions information is available online at http://npg.nature.com/reprints<strong>and</strong>permissions. Correspondence <strong>and</strong> requests for materials should beaddressed to F.V. or J.I.C.636 NATURE PHYSICS | VOL 5 | SEPTEMBER 2009 | www.nature.com/naturephysics© 2009 Macmillan Publishers Limited. All rights reserved.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!