a superconducting quantum simulator for topological order … · a superconducting quantum...

21
A Superconducting Quantum Simulator for Topological Order and the Toric Code Mahdi Sameti, 1 Anton Potoˇ cnik, 2 Dan E. Browne, 3 Andreas Wallraff, 2 and Michael J. Hartmann 1 1 Institute of Photonics and Quantum Sciences, Heriot-Watt University Edinburgh EH14 4AS, United Kingdom 2 Department of Physics, ETH Z¨ urich, CH-8093 Z¨ urich, Switzerland 3 Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, United Kingdom (Dated: March 7, 2019) Topological order is now being established as a central criterion for characterizing and classifying ground states of condensed matter systems and complements categorizations based on symmetries. Fractional quantum Hall systems and quantum spin liquids are receiving substantial interest because of their intriguing quantum correlations, their exotic excitations and prospects for protecting stored quantum information against errors. Here we show that the Hamiltonian of the central model of this class of systems, the Toric Code, can be directly implemented as an analog quantum simulator in lattices of superconducting circuits. The four-body interactions, which lie at its heart, are in our concept realized via Superconducting Quantum Interference Devices (SQUIDs) that are driven by a suitably oscillating flux bias. All physical qubits and coupling SQUIDs can be individually controlled with high precision. Topologically ordered states can be prepared via an adiabatic ramp of the stabilizer interactions. Strings of qubit operators, including the stabilizers and correlations along non-contractible loops, can be read out via a capacitive coupling to read-out resonators. Moreover, the available single qubit operations allow to create and propagate elementary excitations of the Toric Code and to verify their fractional statistics. The architecture we propose allows to implement a large variety of many-body interactions and thus provides a versatile analog quantum simulator for topological order and lattice gauge theories. Topological phases of quantum matter [1, 2] are of sci- entific interest for their intriguing quantum correlation properties, exotic excitations with fractional and non- Abelian quantum statistics, and their prospects for phys- ically protecting quantum information against errors [3]. The model which takes the central role in the discussion of topological order is the Toric Code [4–6], which is an example of a Z 2 lattice gauge theory. It consists of a two- dimensional spin lattice with quasi local four-body inter- actions between the spins and features 4 g degenerate, topologically ordered ground states for a lattice on a sur- face of genus g. The topological order of these 4 g ground states shows up in their correlation properties. The topo- logically ordered states are locally indistinguishable (the reduced density matrices of a single spin are identical for all of them) and only show differences for global prop- erties (correlations along non-contractible loops). More- over local perturbations cannot transform these states into one another and can therefore only excite higher en- ergy states that are separated by a finite energy gap. In equilibrium, the Toric Code ground states are therefore protected against local perturbations that are small com- pared to the gap, which renders them ideal candidates for storing quantum information [4]. Yet for a self-correcting quantum memory, the mobility of excitations needs to be suppressed as well, which so far has only been shown to be achievable in four or more lattice dimensions [7]. Ex- citations above the ground states of the Toric Code are therefore generated by any string of spin rotations with open boundaries. The study of these excitations, called anyons, is of great interest as the neither show bosonic nor fermionic but fractional statistics [4, 5]. In two dimensions, the Toric Code can be visualized as a square lattice where the spin-1/2 degrees of freedom are placed on the edges, see Fig. 1a. Stabilizer operators can be defined for this lattice as products of σ x -operators for all four spins around a star (orange diamond in Fig. 1a), i.e. A s = Q jstar(s) σ x j , and products of σ y -operators for all four spins around a plaquette (green diamond in Fig. 1a), i.e. B p = Q jplaq(p) σ y j . The Hamiltonian for this system reads, H TC = -J s X s A s - J p X p B p , (1) where J s (J p ) is the strength of the star (plaquette) in- teractions and the sums s ( p ) run over all stars (pla- quettes) in the lattice [8]. Despite the intriguing perspectives for quantum com- putation and for exploring strongly correlated condensed matter physics that a realization of the Toric Code Hamiltonian would offer, progress towards its implemen- tation has to date been inhibited by the difficulties in realising clean and strong four-body interactions while at the same time suppressing two-body interactions to a sufficient extent. Realisations of topologically encoded qubits have been reported in superconducting qubits [9, 10], photons [11] and trapped ions [12]. There is an important difference, however, in the use and purpose of an implementation of the Hamiltonian (1) compared to error detection via stabilizer measurements as pur- sued in [11, 13–15]. In particular, two-body interactions are all that is needed for arbitrary stabiliser measure- ments, while lattice gauge theories [16–18] and Hamilto- nians such as (1) require k-body (with k> 2) effective interactions. We note that the realisation of k-body inter- actions is also important in quantum annealing architec- arXiv:1608.04565v1 [quant-ph] 16 Aug 2016

Upload: dotu

Post on 25-Mar-2019

223 views

Category:

Documents


0 download

TRANSCRIPT

A Superconducting Quantum Simulator for Topological Order and the Toric Code

Mahdi Sameti,1 Anton Potocnik,2 Dan E. Browne,3 Andreas Wallraff,2 and Michael J. Hartmann1

1Institute of Photonics and Quantum Sciences, Heriot-Watt University Edinburgh EH14 4AS, United Kingdom2Department of Physics, ETH Zurich, CH-8093 Zurich, Switzerland3Department of Physics and Astronomy, University College London,

Gower Street, London WC1E 6BT, United Kingdom(Dated: March 7, 2019)

Topological order is now being established as a central criterion for characterizing and classifyingground states of condensed matter systems and complements categorizations based on symmetries.Fractional quantum Hall systems and quantum spin liquids are receiving substantial interest becauseof their intriguing quantum correlations, their exotic excitations and prospects for protecting storedquantum information against errors. Here we show that the Hamiltonian of the central model ofthis class of systems, the Toric Code, can be directly implemented as an analog quantum simulatorin lattices of superconducting circuits. The four-body interactions, which lie at its heart, are inour concept realized via Superconducting Quantum Interference Devices (SQUIDs) that are drivenby a suitably oscillating flux bias. All physical qubits and coupling SQUIDs can be individuallycontrolled with high precision. Topologically ordered states can be prepared via an adiabatic ramp ofthe stabilizer interactions. Strings of qubit operators, including the stabilizers and correlations alongnon-contractible loops, can be read out via a capacitive coupling to read-out resonators. Moreover,the available single qubit operations allow to create and propagate elementary excitations of theToric Code and to verify their fractional statistics. The architecture we propose allows to implementa large variety of many-body interactions and thus provides a versatile analog quantum simulatorfor topological order and lattice gauge theories.

Topological phases of quantum matter [1, 2] are of sci-entific interest for their intriguing quantum correlationproperties, exotic excitations with fractional and non-Abelian quantum statistics, and their prospects for phys-ically protecting quantum information against errors [3].The model which takes the central role in the discussionof topological order is the Toric Code [4–6], which is anexample of a Z2 lattice gauge theory. It consists of a two-dimensional spin lattice with quasi local four-body inter-actions between the spins and features 4g degenerate,topologically ordered ground states for a lattice on a sur-face of genus g. The topological order of these 4g groundstates shows up in their correlation properties. The topo-logically ordered states are locally indistinguishable (thereduced density matrices of a single spin are identical forall of them) and only show differences for global prop-erties (correlations along non-contractible loops). More-over local perturbations cannot transform these statesinto one another and can therefore only excite higher en-ergy states that are separated by a finite energy gap. Inequilibrium, the Toric Code ground states are thereforeprotected against local perturbations that are small com-pared to the gap, which renders them ideal candidates forstoring quantum information [4]. Yet for a self-correctingquantum memory, the mobility of excitations needs to besuppressed as well, which so far has only been shown tobe achievable in four or more lattice dimensions [7]. Ex-citations above the ground states of the Toric Code aretherefore generated by any string of spin rotations withopen boundaries. The study of these excitations, calledanyons, is of great interest as the neither show bosonicnor fermionic but fractional statistics [4, 5].

In two dimensions, the Toric Code can be visualized as

a square lattice where the spin-1/2 degrees of freedom areplaced on the edges, see Fig. 1a. Stabilizer operators canbe defined for this lattice as products of σx-operators forall four spins around a star (orange diamond in Fig. 1a),i.e. As =

∏j∈star(s) σ

xj , and products of σy-operators

for all four spins around a plaquette (green diamond inFig. 1a), i.e. Bp =

∏j∈plaq(p) σ

yj . The Hamiltonian for

this system reads,

HTC = −Js∑s

As − Jp∑p

Bp, (1)

where Js(Jp) is the strength of the star (plaquette) in-teractions and the sums

∑s (∑p) run over all stars (pla-

quettes) in the lattice [8].Despite the intriguing perspectives for quantum com-

putation and for exploring strongly correlated condensedmatter physics that a realization of the Toric CodeHamiltonian would offer, progress towards its implemen-tation has to date been inhibited by the difficulties inrealising clean and strong four-body interactions whileat the same time suppressing two-body interactions to asufficient extent. Realisations of topologically encodedqubits have been reported in superconducting qubits[9, 10], photons [11] and trapped ions [12]. There is animportant difference, however, in the use and purposeof an implementation of the Hamiltonian (1) comparedto error detection via stabilizer measurements as pur-sued in [11, 13–15]. In particular, two-body interactionsare all that is needed for arbitrary stabiliser measure-ments, while lattice gauge theories [16–18] and Hamilto-nians such as (1) require k-body (with k > 2) effectiveinteractions. We note that the realisation of k-body inter-actions is also important in quantum annealing architec-

arX

iv:1

608.

0456

5v1

[qu

ant-

ph]

16

Aug

201

6

2

tures, where implementations have been proposed utilis-ing local interactions and many-body constraints [19–21].We anticipate that the approach we describe here mayalso be relevant for quantum annealing, however our fo-cus in this paper is on the realisation of Hamiltonianssupporting topological order.

Superconducting circuits have recently made tremen-dous advances in realizing engineered quantum dynamicsfor quantum simulation [22–26] and quantum informa-tion processing [13–15, 27–31]. Their quantum featuresoriginate from integrated Josephson junctions that be-have like ideal nonlinear inductors. They are typicallybuilt as SQUIDs that can be tuned and driven via mag-netic fluxes. Currently superconducting circuits are as-sembled in ever larger connected networks [13, 14, 28–30, 32]. Here we make use of these exquisite propertiesof superconducting circuits for a scheme implementingthe Hamiltonian of the two-dimensional Toric Code ina direct way by creating four-body interactions of theform As =

∏j∈star(s) σ

xj and Bp =

∏j∈plaq(p) σ

yj between

the respective qubits and fully suppressing all two-bodyinteractions.

Superconducting circuits are an ideal platform for ex-perimentally exploring topological order since high preci-sion control and measurements of both, individual qubitsas well as multi-qubit clusters have been demonstrated[13, 24, 33–35] . Importantly, this platform is also suit-able for realizing periodic boundary conditions, since con-ductors can cross each other via air-bridges [36, 37]. Thisproperty, which is essential for exploring topological or-der, is so far not accessible in other platforms unless oneresorts to digitized approaches.

Our approach is designed to keep the necessary cir-cuitry as simple as possible and only employs standardsuperconducting qubits that are coupled via dc-SQUIDs[38, 39]. To generate the desired interactions we make useof the nonlinear nature of the coupling SQUIDs and thepossibilities to bias them with constant and oscillatingmagnetic fluxes. In particular the four-body stabilizer in-teractions of the Toric Code are generated by driving theSQUIDs with suitably oscillating magnetic fluxes. Theycan therefore be switched on and off at desired times oreven be smoothly tuned in strength. In this way the ToricCode Hamiltonian is realized in a frame where the qubitsrotate at their transition frequencies. Importantly, thetransformation between the rotating and the laboratoryframe is composed of single qubit unitaries only, so thatthe entanglement properties of the states are identical inboth frames. In contrast to earlier approaches to multi-body interactions based on gate sequences [40–43], thefour-body interactions in our approach are directly imple-mented as an analog quantum simulator and are thereforetime-independent, which avoids any errors routed in theTrotter sequencing. Moreover, our approach suppressestwo-body interactions to an extent that they are negligi-ble compared to the four-body interactions.

A minimal two-dimensional lattice covering the surfaceof a torus requires eight qubits connected via supercon-

ducting wires such that the lattice has periodic boundaryconditions in both directions and is thus realizable withavailable technology [13, 14, 28]. We show that protectedstates of the Toric Code can be prepared via an adiabaticsweep from a product state for experimentally realizableparameters and that measurements of individual qubitsand qubit-qubit correlations can be performed to demon-strate the topological order of these states.

I. THE PHYSICAL LATTICE

For realizing the Toric Code lattice, star and plaquetteinteractions, indicated by orange and green diamonds inFig. 1a, are both realized via the circuit shown in Fig. 1bbut with different external fluxes Φext applied to the cou-pling SQUID. The small black squares in Figs. 1a and1b indicate spins formed by the two lowest energy eigen-states of superconducting qubits. For the transition fre-quencies between these two lowest states, we assume aperiodic pattern in both lattice directions such that nonearest neighbors and no next nearest neighbors of qubitshave the same transition frequency and hence all qubitsin each four-spin cell have different transition frequencies,see Sec. III for details.

For each cell the four qubits on its corners are con-nected by a dc-SQUID formed by two identical Joseph-son junctions with Josephson energy EJ , see Fig 1b (anasymmetry between the junctions causes a negligible per-turbation only). The dc-SQUID is threaded by a time de-pendent external flux, Φext and its ports are connectedto the four qubits via four inductors, each with induc-tance L. There are thus auxiliary nodes at both ports ofthe SQUID. As we show in Sec. II, these auxiliary nodes(a : n,m) and (b;n,m) can be eliminated from the de-scription while, together with the flux bias Φext on theSQUID, they mediate the desired four-body interactionsof the stabilizers.

A. Hamiltonian of the physical lattice

We describe the lattice in terms of the phases ϕj andtheir conjugate momenta πj for each node, which obeythe commutation relations [ϕj , πl] = i~δj,l. The nodesare labelled by composite indices j = (n,m), where n la-bels the diagonals extending from top left to bottom rightand m labels the qubits along each diagonal in Fig. 1a.

Each qubit is characterized by a Josephson energyEJq;j and a charging energy ECq;j = e2/(2Cq;j), wheree is the elementary charge and Cq;j the qubit’s capaci-tance. Including a contribution 4ELϕ

2j that arises from

the coupling to its four adjacent auxiliary nodes via theinductances L, the Hamiltonian of the qubit at node jreads,

Hq;j = 4ECq;j~−2π2j − EJq;j cos(ϕj) + 4ELϕ

2j , (2)

3

(a)

'n,m

'n,m+1

'n+1,m+1

'n+1,m

'a;n,m

'b;n,m

2CJ

L

�ext

LL

L

EJ

(b)

correlations are built upcorrelations are built up Tuesday, June 16, 15

EJq;n,m+1

'n,m+1

Cq;n,m+1

FIG. 1. Toric Code lattice and its implementation. (a) Lat-tice for the Toric Code. Small rectangles on the lattice edgesindicate spin-1/2 systems (qubits). Star interactions As aremarked with orange diamonds and plaquette interactions Bp

with green diamonds. The transition frequencies of the qubitsform a periodic pattern in both lattice directions such that nonearest neighbors and no next nearest neighbors of spins havethe same transition frequency and therefore all four spins atthe corners of a physical cell have different transition frequen-cies. (b) Circuit of a physical cell. The rectangles at the cor-ners represent qubits with different transition frequencies, seeinset for the circuit of each qubit. Qubits (n,m) and (n,m+1)[(n + 1,m) and (n + 1,m + 1)] are connected via the induc-tances L to the node (a;n,m) [(b;n,m)]. The nodes (a;n,m)and (b;n,m) in turn are connected via a dc-SQUID with anexternal magnetic flux Φext threaded through its loop. TheJosephson junctions in the dc-SQUID have Josephson energiesEJ and capacitances CJ , which include shunt capacitances.

where EL = φ20/(2L) and φ0 = ~/(2e) is the rescaled

quantum of flux. We consider transmon qubits [13, 14,23, 24, 28, 29, 44] because of their favorable coherenceproperties and moderate charging energies, c.f. appendixD 1.

For the description of the dc-SQUIDs it is useful to

consider the modes ϕ±;j = ϕa;j ± ϕb;j , where ϕa/b;j isthe phase at the auxiliary node (a/b;n,m). The modesϕ+;j behave as harmonic oscillators described by theHamiltonians H+;j = 4EC+~−2π2

+;j + ELϕ2+;j , where

π±;j are the canonically conjugate momenta to ϕ±;j andEC+ = e2/Cg, with Cg the capacitance between each ofthe nodes (a; j) or (b; j) and ground. The modes ϕ−;j inturn are influenced by the dc-SQUID. Their Hamiltoni-ans read

H−;j = 4EC−~2

π2−;j +ELϕ

2−;j−2EJ cos

(ϕext2

)cos(ϕ−;j),

(3)where ϕext = Φext/φ0 is the phase associated to the ex-ternal flux through the SQUID loop, EJ the Josephsonenergy of each junction in the SQUID and EC− = e2/2CJwith CJ the capacitance of one junction in the SQUID,including a shunt capacitance parallel to the junction.

We choose the indices j = (n,m) for the auxiliarynodes such that node (a;n,m) couples to the qubits atnodes (n,m) and (n,m+ 1) whereas node (b;n,m) cou-ples to the qubits at nodes (n+ 1,m) and (n+ 1,m+ 1),see Fig. 1b. The interaction terms between the qubitvariables ϕj and the SQUID modes ϕ±;j due to the in-ductances L thus read,

HL;j = −ELϕq+;jϕ+;j − ELϕq−;jϕ−;j (4)

where ϕq±;j = ϕn,m+ϕn,m+1±ϕn+1,m±ϕn+1,m+1. Here,inductive couplings via mutual inductances are an alter-native option [45]. The full Hamiltonian of the lattice istherefore given by

H =

(N,M)∑j=(1,1)

[Hq;j +H+;j +H−,j +HL;j ] (5)

A detailed derivation of the HamiltonianH, starting froma Lagrangian for the circuit, is provided in appendix A.We now turn to explain how the four-body interactionsof the stabilizers emerge in our architecture.

II. STABILIZER OPERATORS

The guiding idea for our approach comes from the fol-lowing approximate consideration (We present a quan-titatively precise derivation in Sec. II A). In each singlecell, as sketched in Fig. 1b, the auxiliary nodes (a;n,m)and (b;n,m) have very small capacitances with respectto ground and therefore cannot accumulate charge. As aconsequence the node phases ϕa;n,m and ϕb;n,m are notdynamical degrees of freedom but instantly follow thephases at nodes (n,m) and (n,m+ 1) or (n+ 1,m) and(n + 1,m + 1) as given by Kirchoff’s current law say-ing that the sum of all currents into and out of such anode must vanish. Therefore the Josephson energy of thecoupling SQUID, 2EJ cos(ϕext/2) cos(ϕa;n,m − ϕb;n,m),becomes a nonlinear function of the node phases ϕn,m,

4

ϕn,m+1, ϕn+1,m and ϕn+1,m+1, which contains four-bodyterms.

The external flux applied through the SQUID is thenchosen such that it triggers the desired four-body interac-tions forming the stabilizers. To this end it is composedof a constant part with associated phase ϕdc and an os-cillating part associated with phase amplitude ϕac,

ϕext(t) = ϕdc + ϕacF (t), (6)

where the oscillation frequencies of F (t) are linear com-binations of the qubit transition frequencies such thatthe desired four-body spin interactions of the Hamilto-nian (1) are enabled and all other interaction terms aresuppressed. Moreover we choose ϕdc = π to suppress thecritical current Ic = 2EJφ

−10 cos(ϕdc/2) of the dc-SQUID

[46] and thus unwanted interactions of excitations be-tween qubits at opposite sides of it, c.f. Fig. 1b. Assum-ing furthermore that the oscillating part of the flux biasis small compared to the constant part, we approximatecos(ϕext/2) ≈ −ϕacF (t)/2.

The above consideration uses approximations and onlyserves us as a guide. To obtain a quantitatively precisepicture, we here take the small but finite capacitances ofthe Josephson junctions in the SQUID into account andderive the stabilizer interactions in a fully quantum me-chanical approach. Our derivation makes use of meth-ods for generating effective four-body Hamiltonians inthe low energy sector of a two-body Hamiltonian [47–50]by employing a Schrieffer-Wolff transformation and com-bines these with multi-tone driving in order to single outspecific many-body interactions while fully suppressingall two-body interactions.

Further intuition for our approach can be obtainedfrom the macroscopic analogy explained in Fig. 2. Inshort, the plasma frequency of the coupling SQUIDs dif-fers form all transition frequencies of the qubits they cou-ple to as well as from all frequencies contained in the os-cillating flux that is threaded through their loops. Theythus decouple from the qubits but mediate the desiredfour-body interactions via the drives applied to them.Further details of this analogy are explained in the cap-tion of Fig. 2. We now turn to present the quantitativelyprecise derivation.

A. Elimination of the SQUID degrees of freedom

For the parameters of our architecture, the oscillationfrequencies of the SQUID modes ϕ±;j differ from thetransition frequencies of the qubits and the frequencies ofthe time dependent fluxes threaded through the SQUIDloops by an amount that exceeds their mutual couplingsby more than an order of magnitude. These modes willthus remain in their ground states to very high precision.To eliminate them from the description and obtain aneffective Hamiltonian for the qubits only, it is useful toconsider the Schrieffer-Wolff transformation

H → H = eSHe−S with (7)

cos(!dt)

!�

!1

q1 e�i!1t

!2

q2 e�i!2t

q3 e�i!3t

!3

q4 e�i!4t

!4

qubit

SQUID

anharmonic spring

harmonic spring

FIG. 2. Macroscopic analogy providing an intuitive expla-nation of our approach to generate the stabilizer interactions.The qubits are nonlinear oscillators so that one can think ofthem as masses (corresponding to the qubits’ capacitances)attached to anharmonic springs (corresponding to the qubitsJosephson energy). We here sketch four nonlinear oscillatorsrepresenting four qubits as green spheres connected by black(anharmonic) springs to the support frame. Note that all fourqubits have different transition frequencies so that all fourspheres representing them here would have different massesand their springs different spring constants. For vanishingcritical current the relevant SQUID modes ϕ−;j , behave asharmonic oscillators and in our macroscopic analogy we canrepresent them by a mass (here sketched as a red sphere, cor-responding to the capacitance 2CJ) that is attached via foursprings (sketched in orange, corresponding to the inductors L)to the four masses representing the neighboring qubits. Forvanishing critical current, the only potential energy of themodes ϕ−;j comes from their inductive coupling to the neigh-boring qubits, and hence the mass representing the SQUIDmode (red sphere) has no spring connecting it to a supportframe. We now imagine that we shake the mass representingthe SQUID mode with a driving term that is proportional tothe fourth power of the mass’ position variable, ϕ4

−;j . Thiscorresponds to the oscillating flux bias ∝ ϕacF (t). If we shakethis mass at a frequency ωd that is distinct from its eigenfre-quency ω−;j but equals the sum of the oscillation frequenciesof all four neighboring masses (neighboring qubits), we drive a

term proportional to q†1q†2q†3q†4, where q†j creates an excitation

in qubit j. If, on the other hand we shake the SQUID-massat a frequency ωd = ω1 +ω2−ω3−ω4 we drive a term propor-tional to q†1q

†2q3q4. The selective driving of specific four-body

terms, as we consider it here, of course relies on choosing thetransition frequencies of all four qubits to be different. Ourapproach considers a superposition of various drive frequen-cies such that the sum of the triggered four-body terms addsup to the desired stabilizer interaction, see text.

S =i

2~

(N,M)∑j=(1,1)

(ϕq+;jπ+;j + ϕq−;jπ−;j),

5

which eliminates the qubit-SQUID interactions describedin Eq. (4). The effective Hamiltonian for the qubits isthen obtained by projecting the resulting Hamiltonianonto the ground states of the SQUID modes ϕ±;j . Thedetails of this derivation are discussed in appendix B.

For the transformed Hamiltonian H, see Eq. (7), weconsider terms up to fourth order in S to get the leadingcontributions of the terms which are in resonance withthe oscillating part of ϕext(t), c.f. Eq. (6) . On theother hand, a truncation of the expansion is here welljustified since both fields ϕq±;j and π±;j have low enoughamplitudes, see Eq. (B16), so that higher order terms cansafely be ignored. The transformed Hamiltonian can bewritten as

H = H0q +HIq +HS +HqS , (8)

where H0q =∑(N,M)j=(1,1)Hq;j with Hq;j as in Eq. (2) de-

scribes the individual qubits and

HIq = −2EL

N,M∑n,m=1

ϕn,mϕn,m+1 (9)

interactions between neighboring qubits that are ineffec-tive since their transition frequencies differ by an amountthat exceeds the relevant coupling strength by more thanan order of magnitude, see appendix D 1.

The remaining two terms HS and HqS in Eq. (8) de-scribe the SQUIDs and the residual couplings betweenthe qubits and the SQUIDs, see Eqs. (B7) and (B8) in ap-pendix B. Since the oscillation frequencies of the SQUIDmodes are chosen such that they cannot be excited by theapplied driving fields or via their coupling to the qubits,an effective Hamiltonian, describing only the qubits, canbe obtained by projecting HqS onto the ground states ofthe SQUID modes,

HqS → H(0)qS = 〈0SQUIDs|HqS |0SQUIDs〉. (10)

As the SQUID modes remain in their ground states,their Hamiltonian HS may be discarded. The projected

Hamiltonian H(0)qS contains two-body interactions and

four-body interactions between the qubits,

H(0)qS = −χϕacEJ

8

N,M∑n,m=1

F (t)ϕ2q−;n,m

+1

4!

χϕacEJ16

N,M∑n,m=1

F (t)ϕ4q−;n,m

(11)

with χ = 〈0−;n,m| cos(ϕ−;n,m)|0−;n,m〉, where |0−;n,m〉 isthe ground state of the mode ϕ−;n,m.

We emphasize here that our approach requires non-linear coupling circuits, such as the considered SQUIDs.The terms in the second line of Eq. (11), which give rise tothe desired four-body interactions, only emerge becausethe Hamiltonian of the coupling SQUIDs contains termsthat are at least fourth power in ϕ−;n,m, see appendix B.

The oscillation frequencies contained in the externalflux, i.e. in ϕext, can be chosen such that selected termsin the second line of Eq. (11) are enabled as they becomeresonant. For suitable amplitudes of the frequency com-ponents of ϕext, these terms add up to the desired four-body stabilizer interactions whereas all other interactionscontained in Eq. (11) are for our choice of parametersstrongly suppressed in a rotating wave approximation.

Before discussing this point in more detail in the nextsection, let us here emphasize the versatility of our ap-proach: With suitably chosen Fourier components of ϕextalmost any two-, three- or four-body interaction can begenerated. The only limitation here is that terms withthe same rotation frequency can not be distinguished, e.g.one can not choose between generating σzjσ

zk or σzjσ

zl for

j 6= k 6= l which are both non-rotating. On the otherhand, straightforward modifications of the circuit allowto implement higher than four-body (e.g. five- and six-body) interactions.

B. Effective Hamiltonian of one cell

We now focus on the interactions between the fourqubits in one cell only, see Fig. 1b. To simplify thenotation we label these four qubits clockwise with in-dices 1,2,3 and 4, i.e. (n,m) → 1, (n + 1,m) → 2,(n + 1,m + 1) → 3 and (n,m + 1) → 4. We thus di-vide one term, say for (n,m) = (n0,m0), in the sum ofEq. (11) into the following parts,

Hcell = H(0)qS

∣∣∣cell

= H1 +H2 +H3, (12)

where

H1 =χϕacEJ

16F (t)ϕ1ϕ2ϕ3ϕ4,

contains the four-body interactions of the stabilizers,

H2 = −χϕacEJ8

F (t)(ϕ1 + ϕ4 − ϕ2 − ϕ3)2,

contains all terms that are quadratic in the node phasesϕj and

H3 = −χϕacEJF (t)∑

(i,j,k,l)

Cijklϕiϕjϕkϕl

contains all terms that are fourth-order in the nodephases ϕj except for the four-body interactions. Theseare terms where at least two of the indices i, j, k and lare the same as indicated by the notation

∑(i,j,k,l). De-

tails of the derivation of the Hamiltonian (12) are givenin appendix C, where we also state the exact form of thecoefficients Cijkl.

The Hamiltonian (12) can be written in second quan-

tized form using ϕj = ϕj(qj + q†j ), where q†j (qj) creates

(annihiliates) a bosonic excitation at node j with energy

6

~ωj =√

8EC;j(EJq;j + 4EL) and zero point fluctuation

amplitude ϕj = [2EC;j/(EJq;j + 4EL)]1/4. Note that theeffective Josephson energy EJq;j+4EL includes here con-tributions of the attached inductances, see appendix B 1for details. Due to the qubit nonlinearities we can re-strict the description to the two lowest energy levels of

each node by replacing qj → σ−j (q†j → σ+j ) and get for

the four-body interactions,

H1 =χϕacEJ

16F (t)

4∏i=1

ϕi∑

a,b,c,d∈{+,−}σa1σ

b2σc3σd4 , (13)

Here, the sum∑a,b,c,d∈{+,−} runs over all 16 possible

strings of σ+ and σ− operators, which are listed in tableII in appendix C.

The explicit form of the Hamiltonians H2 in secondquantized form is given in Eq. (D15). The terms con-tained in H2 only lead to frequency shifts for the qubitsand negligible two-body interactions. The frequencyshifts can be compensated for by modified frequenciesfor the driving fields and the residual two-body interac-tions can be made two orders of magnitude smaller thanthe targeted four-body interactions, see Sec. II C and ap-pendix D. Moreover the terms contained in H3 only leadto corrections that are at least an order of magnitudesmaller than those of H2 and can safely be discarded, seeappendix D 3. Here, we therefore focus on the part H1

which gives rise to the dominant terms in the form offour-body stabilizer interactions.

C. Triggering four-body interactions

In a rotating frame where each qubit rotates at its tran-sition frequency, σ+

j (t) = eiωjtσ+j [σ−j (t) = e−iωjtσ−j ], the

strings σa1σb2σc3σd4 that appear in Eq. (13) rotate at the

frequencies ωa,b,c,d = aω1 + bω2 + cω3 +dω4. Hence, if wethread a magnetic flux, which oscillates at the frequencyωa,b,c,d, through the SQUID loop, the corresponding op-erator σa1σ

b2σc3σd4 acquires a pre-factor with oscillation fre-

quency ωa,b,c,d, which renders the term non-oscillatingand hence enables this particular four-body interaction.Importantly, all qubits of a cell have different transitionfrequencies such that a drive with frequency ωa,b,c,d onlyenables the particular interaction σa1σ

b2σc3σd4 whereas all

other interactions are suppressed by their fast oscillatingprefactors.

As the star and plaquette interactions Jsσx1σ

x2σ

x3σ

x4 and

Jpσy1σ

y2σ

y3σ

y4 can be written as linear combinations of op-

erator products σa1σb2σc3σd4 , their generation requires mag-

netic fluxes that are a superposition of oscillations at therespective frequencies. To enable the specific four-bodyinteractions of the Toric Code we thus consider,

F (t) =

{Fs = FX + FYFp = FX − FY

(14)

where FX = f1,1,1,1 + f1,1,−1,−1 + f1,−1,1,−1 + f1,−1,−1,1

and FY = f1,1,1,−1 + f1,1,−1,1 + f1,−1,1,1 + f−1,1,1,1 with

fa,b,c,d = cos(ωa,b,c,dt). The driving field Fs generates astar interaction As and Fp a plaquette interaction Bp.Hence star and plaquette interactions only differ by therelative sign of the FX and FY contributions correspond-ing to a relative phase of π of the respective Fourier com-ponents. Applying Fs or Fp to a cell, its Hamiltonian (13)thus reads,

H(1)cell(t)

∣∣∣Fs

= −Jsσx1σx2σx3σx4 + r.t. for Fs (15a)

H(1)cell(t)

∣∣∣Fp

= −Jpσy1σy2σ

y3σ

y4 + r.t. for Fp (15b)

where r.t. refers to the remaining fast rotating terms, and

Js = Jp =χEJϕac

16

4∏i=1

ϕi (16)

Note that the signs and magnitudes of the interactionsJs and Jp can for each cell be tuned independently viathe respective choices for the drive amplitudes ϕac.

Besides the stabilizer interactions, there are furtherterms in the Hamiltonian (8). As we show in appendixD, all these other interactions are strongly suppressed bytheir fast oscillating prefactors (the oscillation frequen-cies of the prefactors exceed the interaction strengths bymore than an order of magnitude in all cases). Theireffect can be estimated using time dependent perturba-tion theory. We find for our parameters that they lead tofrequency shifts for the individual qubits, which can becompensated for by a modification of the frequencies ofthe driving field ϕext. For our choice of qubit transitionfrequencies, where no nearest neighbors or next nearestneighbors of qubits have the same transition frequencies,the leading corrections to the Toric Code Hamiltonianof Eq. (1), besides local frequency shifts, are two-bodyinteractions between qubits that are further apart (i.e.neither nearest neighbors nor next neighrest neighbors).For suitable parameters as discussed in Sec. III, theseinteractions are about two orders of magnitude weakerthan the stabilizers and thus even weaker than dissipa-tion processes in the device, see appendix D.

In this context, we also note that the leading contribu-tion to the perturbations of the Hamiltonian (1) is pro-portional to the charging energy ECq;j of the employedqubits, c.f. appendix D. Hence the considered transmonqubits are a well suited choice for our aims.

D. Including dissipative processes

As we have shown, the dynamics of the architecturewe envision can be described by the Hamiltonian (1) in arotating frame with respect to H0 =

∑j ωjσ

+j σ−j , where

we assume that all frequency shifts discussed in the pre-vious section (see also appendix D) have been absorbedinto a redefinition of the ωj . In an experimentally re-alistic scenario, the qubits will however be affected by

7

FIG. 3. Lattice for minimal realization on a torus with8 physical qubits including 4 star interactions (orange dia-monds) and 4 plaquette interactions (green diamonds). Thequbits are indicated by small colored squares, where the col-oring encodes their transition frequencies and the numberingindicates the periodic boundary conditions. Here, 8 differenttransition frequencies are needed.

dissipation. The full dynamics of the circuit can there-fore be described by the Markovian master equation,

ρ = −i~−1[H, ρ] +Dr[ρ] +Dd[ρ] (17)

where ρ is the state of the qubit excitations, H is as inEq. (1), Dr[ρ] = κ

2

∑j(2σ

−j ρσ

+j − σ

+j σ−j ρ − ρσ

+j σ−j ) de-

scribes relaxation at a rate κ and Dd[ρ] = κ′∑j(σ

zj ρσ

zj −

ρ) pure dephasing at a rate κ′. Note that both, the relax-ation and dephasing terms Dr[ρ] and Dd[ρ] are invariantunder the transformation into the rotating frame.

In the following we turn to discuss realistic experi-mental parameters that lead to an effective Toric CodeHamiltonian, the preparation and verification of topolog-ical order in experiments and protocols for probing theanyonic statistics of excitations in a realization of theToric Code with our approach. In all these discussions,we fully take the relaxation and dephasing processes de-scribed by Eq. (17) into account.

III. TOWARDS AN EXPERIMENTALREALISATION

An experimental implementation of our approach isfeasible with existing technology as it is based on trans-mon qubits which are the current state of the art in manylaboratories. Moreover, periodic driving of a coupling el-ement at frequencies comparable to the qubit transitionfrequencies is a well established technique [35, 51].

The minimal lattice of a Toric Code with periodicboundary conditions in both lattice directions is theeight-qubit lattice sketched in Fig. 3. A working exam-ple for this eight-qubit realization with experimentallyrealistic parameters is provided by the qubit transitionfrequencies given in table I, which are realized in qubitswith Josephson energies in the range 4.9 GHz ≤ EJq/h ≤

qubit trans. frequency qubit trans. frequencyωj/2π ωj/2π

1 13.8 GHz 5 13.1 GHz2 13.0 GHz 6 13.7 GHz3 3.5 GHz 7 4.2 GHz4 4.1 GHz 8 3.4 GHz

TABLE I. Possible transition frequencies for an eight-qubitToric Code in natural frequency units. The qubit numberscorrespond to the numbering chosen in Fig. 3. These fre-quencies are assumed to already include the frequency shiftscalculated in appendix D.

94 GHz (h is Planck’s constant) and capacitances in therange 52 fF ≤ Cq;j ≤ 77 fF. Note that the frequenciesare chosen such that neighboring qubits are detuned fromeach other by several GHz whereas the transition frequen-cies of next nearest neighbor qubits differ by ∼ 100 MHz.

For the coupling SQUIDs, the Josephson energy of eachjunction should be EJ/h ≈ 10 GHz and the associatedcapacitance 3.47 fF. Since high transparency Josephsonjunctions typically have capacitances of 1 fF per 50 GHzof EJ , each SQUID is shunted by a capacitance of afew fF. For these parameters the modes ϕ−;j oscillateat ω−/2π = 8.45 GHz, whereas the oscillation frequencyof the modes ϕ+;j is much higher since Cg � CJ .

Furthermore the inductances coupling the qubits tothe SQUIDs need to be chosen large enough so that theresulting coupling term remains weak compared to thequbit nonlinearities. This requires inductances of about100 nH which can be built as superinductors in the formof Josephson junction arrays [52, 53] or with high ki-netic inductance superconducting nanowires [54]. Wenote that our concept is compatible with nonlinearitiesJosephson junction arrays may have. The amplitude ofthe drive is ϕac = 0.1 rad corresponding to an oscillat-ing flux bias of amplitude 0.1×Φ0 at the SQUID. Theseparameters give rise to a strength for star and plaque-tte interactions of Js = Jp = h × 1 MHz, which exceedscurrently achievable dissipation rates of transmon qubitssignificantly [26, 55–58]. The strength of the stabilizerinteractions can be enhanced further by increasing theJosephson energies of the SQUIDs EJ and/or the ampli-tudes of the oscillating drives ϕac.

For the above parameters, the total oscillation fre-quency, i.e. the sum of the frequencies of the qubitoperators and the frequency of the external flux, is forany remaining two-body interaction term at least 10-30times larger than the corresponding coupling strength.To leading order, these terms give rise to frequency shiftsof the qubits which can be compensated for by a modifi-cation of the drive frequencies. All interaction terms con-tained in higher order correction are negligible comparedto the four-body interactions forming the stabilizers. Wenote that the experimental verification of the topologicalphase is further facilitated by its robustness against localperturbations, such as residual two-body interactions or

8

FIG. 4. Pattern of transition frequencies for an implemen-tation of a large lattice. The qubits are indicated by smallcolored squares, where the coloring encodes their transitionfrequencies. Here four principal transition frequencies areneeded. These differ by a few GHz and are indicated bythe filling colors brown, blue, red and purple. To suppresstwo-body interactions between next nearest neighbor qubits,each pair of these needs to be detuned from each other by∼ 100 MHz. The resulting frequency differences are indicatedby the colored frames around the respective squares. Thegray diamond highlights the frequency set which is periodi-cally repeated in both lattice directions.

shifts of the qubit transition frequencies, c.f. [59, 60].In common with all two-dimensional lattices, the ar-

chitecture we envision requires an increasing number ofcontrol lines for the qubits, SQUIDs and the read-out asone enhances the lattice size. Whereas solutions to thiscomplexity challenge are already being developed, e.g. byplacing the control lines on a second layer [29, 61, 62], aminimal realization of the Hamiltonian (1) requires onlyeight qubits that form four stars and four plaquettes, seeFig. 3. Moreover, these eight physical qubits are arrangedon a 4 × 2 lattice so that every qubit is located on itsboundary, facilitating the control access.

Our scheme also allows to realize the Toric Code on alarger lattice. The pattern of transition frequencies forthis case is sketched in Fig. 4 and can be realized withsimilar circuit parameters as given above.

For an eight qubit realization, controlling single qubitstates and two qubit correlation measurements would re-quire a charge line and a readout resonator for each qubit.Our scheme, where a single cell features four qubits withdifferent transition frequencies, has the beneficial prop-erty that four readout resonators can be connected to asingle transmission line and the qubit states read out us-ing frequency division multiplexing [63–65]. Such a jointreadout is capable of measuring the state of all four qubitsand qubit pairs even when their transition frequenciesare separated by several GHz. An improvement over the

readout through a common transmission line would be toemploy a common Purcell filter [66, 67] with two-modes,one mode placed at 3.5 GHz and the second at 12.5GHz. Such scheme would not only allow a joint readoutof four qubits but also protect them against the Purcelldecay. Frequency division multiplexing can also be usedfor qubit excitation where a single charge line can be splitor routed to four qubits in the cell. If a common trans-mission line is used for the readout the same line can alsobe used as a feedline for qubit state preparation [14].

Larger and more realistic Toric code realizations willsuffer from increased complexity. This could be drasti-cally simplified when adding on-chip RF switches [68] forrouting microwave signals to different cells and using se-lective broadcasting technique which would also reducethe number of microwave devices such as signal genera-tors and arbitrary waveform generators [69].

IV. ADIABATIC PREPARATION OFTOPOLOGICAL ORDER

On the surface of a torus, i.e. for a two dimensional lat-tice with periodic boundary conditions in both lattice di-rections, the Hamiltonian (1) has four topologically pro-tected, degenerate eigenstates |ψj〉 (j = 1, 2, 3, 4), whichare joint eigenstates of the stabilizers As and Bp witheigenvalue +1, see appendix E for a possible choice of the|ψj〉 for an eight-qubit lattice. A state in this manifold ofprotected states can either be prepared via measurementsof the stabilizer operators or via an adiabatic sweep thatstarts from a product state [70]. Measurements of sta-bilizers can in our architecture be performed via a jointdispersive readout of all qubits in a star or plaquette, seeSec. V.

For the adiabatic sweep, one starts in our realisationwith the oscillating driving fields turned off, ϕac = 0and the qubit transition frequencies blue-detuned by adetuning ∆ with respect to their final values ωj . In thecryogenic environment of an experiment, this initializesthe system in a product state, where all qubits are intheir ground state, |ψ0〉 =

⊗j |0j〉. Then the amplitudes

of the external oscillating fluxes, ϕac, are gradually in-creased from zero to their final values ϕac,f , i.e. ϕac(t) =λ(t)ϕac,f , where λ(t = 0) = 0, λ(TS) = 1, and TS is theduration of the sweep. The qubit transition frequenciesare decreased to ωj , i.e. ω′j(t) − ωj = ∆[1 − λ(t)] > 0,with ω′j(0) − ωj = ∆ and ω′j(TS) − ωj = 0. For thissweep, the Hamiltonian of the circuit can be written inthe rotating frame as,

H(λ) =∑j

(1−λ)∆

2σzj −λJs

∑s

As−λJp∑p

Bp, (18)

where As and Bp are as in Eq. (1). Thus the initialHamiltonian is H(λ = 0) =

∑j ∆σzj /2 and the unique

initial ground state is the vacuum with no qubit exci-tations (all spins down), i.e. |ψ0〉 =

⊗j |0j〉. The fi-

nal Hamiltonian is the targeted Toric Code Hamiltonian

9

correlations are built upcorrelations are built up

MHz

Thursday, June 25, 15

FIG. 5. The energy eigenvalues of H(λ) as a function of λ forthe eight-qubit lattice with ∆ = 2π × 5MHz and Js = Jp =2π × 2MHz. The initial ground state is the unique vacuum|ψ0〉 =

⊗8j=1 |0j〉. At λ ∼ 0.6 the ground state starts to

become degenerate but the degenerate ground state manifoldis always separated by a gap ∼ Js ∼ Jp from higher excitedstates.

given in Eq. (1). Provided the parameters are tuned suf-ficiently slow, the resulting sweep will drive the systemadiabatically into the manifold of protected states. Thespectrum of the minimal eight-qubit lattice, c.f. Fig. 3, isshown in Fig. 5 as a function of the tuning parameter λ.One sees that there is one unique ground state at λ = 0which eventually becomes degenerate with three otherstates to form the degenerate four-dimensional manifoldof topologically ordered ground states at λ = 1. For anyvalue of λ there is however a finite energy separation ofthe order of Js or Jp to all other states which makes thesweep robust. Moreover the initial detuning is needed toavoid a kick of the system by a sudden quench as oneswitches to the rotating frame.

Figure 6a shows the fidelity F =∑4j=1〈ψj |ρ(TS)|ψj〉

for being in the protected qubit subspace as a function ofthe sweep time TS and the initial detuning ∆. Here ρ(t)is the actual state of the system in the rotating frame andthe |ψj〉 (j = 1, . . . , 4) form the subspace of degenerateprotected ground states. To test that the prepared stateis pure, we also plot the purity P = Tr(ρ2) as a functionof TS and ∆ in Fig. 6b. We choose Js = Jp = h ×2 MHz, see Sec. II D, and relaxation and dephasing timesT1 = T2 = 50µs. For these values, a protected state cansuccessfully be prepared.

V. MEASURING TOPOLOGICALCORRELATIONS

Our scheme allows to experimentally explore and ver-ify topological order. To this end one prepares the latticein one of the topologically ordered states |ψj〉 and mea-sures the state of individual physical qubits as well as acorrelation 〈

∏l∈v σ

xl 〉 for a non-contractible path v along

one of the lattice directions, which is closed in a loopto implement the periodic boundary conditions. For the

minimal eight qubit realisation this is merely a measure-ment of a two-spin correlation 〈σxmσxn〉. Non-contractibleloops of σy operations, such as

∏l∈v′ σ

yl , transform the

protected states |ψj〉 into one another [5]. Here, these canbe realized via strings of π-pulses applied to the qubitsalong the loop. After applying a string of π-pulses alongone lattice direction, which transforms |ψj〉 into a differ-ent topologically correlated state |ψj′〉, one again mea-sures individual spins as well as the correlation 〈

∏l∈v σ

xl 〉.

Whereas the states of the individual spins should be thesame for |ψj〉 and |ψj′〉 because the topologically corre-lated states are locally indistinguishable, the results forthe correlations should be distinct. This experimentalverification of topological correlations can even be per-formed for the minimal lattice consisting of eight qubits,where it only requires measuring individual qubits andtwo-qubit correlations. An example of possible outcomesfor these measurements are given in table III and dis-cussed in appendix E.

In recent years several strategies for state tomographyin superconducting circuits have been demonstrated ex-perimentally [14, 33, 34, 71]. These either proceed byreading out the qubit excitations via an auxiliary phase-qubit or resonator. Both concepts are compatible withour circuit design. In particular correlations betweenmultiple qubits, including the stabilizers As and Bp canbe measured in a joint dispersive read-out via a commoncoplanar waveguide resonator [14, 24, 33, 34] or severalreadout resonators connected to a single transmission lineusing frequency multiplexing [63–65].

VI. BRAIDING AND DETECTION OF ANYONS

Our approach is also ideally suited for probing the frac-tional statistics of excitations in the Toric Code. Oncethe lattice has been prepared in a topologically orderedstate, for example via the sweep described in Sec. IV, theanyonic character of the excitations can be verified in anexperiment using single qubit rotations and the measure-ment of one stabilizer only. We here discuss a protocolthat has originally been proposed for cold atoms [72] butis very well suited for superconducting circuits.

For the prepared topologically ordered state |ψ〉, theprotocol considers applying a single qubit rotation σy

to one of the qubits. This operation, which in our ap-proach can be implemented via a microwave pulse in acharge line to the qubit, creates a pair of so called e-particles [5] on the neighboring vertices and puts the sys-tem into a state |ψ, e〉. Alternatively, half a σy-rotation

creates the state (|ψ〉 + |ψ, e〉)/√

2. Similarly, with aσx-rotation on another qubit, one creates a pair of m-particles. Then one can move one of these m-particlesaround one of the e-particles via successive σx-rotationson the qubits along the chosen path and fuse both m-particles again. After this sequence, the system is inthe state (|ψ〉 − |ψ, e〉)/

√2 due to the fractional phase

π acquired from braiding the e- and m-particles. This

10

0.71.5

0.75

0.8

1000

0.85

800

0.9

1600

0.95

4000.5 200

0.81.5

0.85

1000

0.9

8001600

0.95

4000.5 200 � [MHz]

TS [µs]

F P

� [MHz]TS [µs]

FIG. 6. Adiabatic preparation of an eight qubit lattice in the protected subspace. (a) Fidelity F for being in the protectedsubspace and (b) purity P of the state. Js = Jp = h× 2 MHz and coherence times T1 = T2 = 50µs, see Sec. II D.

phase flip can for instance be detected with another√σy-operation that maps (|ψ〉 − |ψ, e〉)/

√2 → |ψ〉 and

(|ψ〉 + |ψ, e〉)/√

2 → |ψ, e〉 and a subsequent measure-ment of the corresponding stabilizer As. We note thatin our approach all single qubit operations can be imple-mented via pulses applied through charge lines and thestabilizer measurement can be realized via the techniquedescribed in Sec. V.

VII. CONCLUSION AND OUTLOOK

We have introduced a scheme and architecture for aquantum simulator of topological order that implementsthe central model of this class of quantum many-bodysystems, the Toric Code, on a two dimensional super-conducting circuit lattice. Our approach is based on ascheme that allows to realize a large variety of many-bodyinteractions by coupling multiple qubits via dc SQUIDsthat are driven by a suitably oscillating flux bias. Thescheme is thus highly versatile and not restricted to im-plementations of the Toric Code only. For example, bytuning the time-independent flux bias of the couplingSQUIDs away from the operating point for the ToricCode, one can generate additional two-body interactionsof the form σzjσ

zl between neighboring qubits which can

eventually become stronger than the stabilizer interac-tions. Since the transition frequencies of the qubits maybe tuned as well, these control handles make our ap-proach a very versatile quantum simulator for exploringquantum phase transitions between the topologically or-dered and other phases of two-dimensional spin lattices.

The advanced status of experimental technology for su-perconducting circuits makes available an ample toolboxfor exploring the physics of topological order. All physi-

cal qubits and coupling SQUIDs can be individually con-trolled with high precision, topologically ordered statescan be prepared via an adiabatic ramp of the stabilizerinteractions, multi-qubit correlations including stabiliz-ers and correlations along non-contractible loops can bemeasured and anyons can be braided and their fractionalstatistics can be verified. Moreover, superconducting cir-cuit technology allows for realizing periodic boundaryconditions, a feature that not easily available in otherplatforms but crucial for exploring topological order.

We expect our work to open up a realistic path towardsan experimental investigations of this intriguing area ofquantum many-body physics that at the same time willenable access for detailed measurements.

VIII. ACKNOWLEDGEMENTS

M. S. and M. J. H. are grateful for discussionswith Martin Leib at an early stage of this project.A. P. and A. W. acknowledge support by ETH Zurich.M. J. H. was supported by the EPSRC under Grant No.EP/N009428/1. M. S. and M. J. H. were supported inpart by the National Science Foundation under GrantNo. NSF PHY11-25915 for a visit to the program “Many-Body Physics with Light” at KITP Santa Barbara wherea part of this work has been done.

Appendix A: Lagrangian and Hamiltonian of entirecircuit

In this appendix we present a detailed derivation of theHamiltonian in Eq. (5) that describes the considered cir-cuit lattice. We start with the full Lagrangian describingthe lattice depicted in Fig. 1, which reads,

11

L =

N,M∑n,m=1

[φ2

0

Cq;n,m2

ϕ2n,m + EJq;n,m cos(ϕn,m)

](A1)

−M,N∑n,m=1

φ20

2L

[(ϕa;n,m − ϕn,m)2 + (ϕa;n,m − ϕn,m+1)2 + (ϕb;n,m − ϕn+1,m)2 + (ϕb;n,m − ϕn+1,m+1)2

]+

M,N∑n,m=1

φ20

[Cg2ϕ2a;n,m +

Cg2ϕ2b;n,m + CJ(ϕa;n,m − ϕb;n,m)2

]+

M,N∑n,m=1

2EJ cos(ϕext

2) cos(ϕa;n,m − ϕb;n,m).

As explained in the main text, ϕn,m is the phase at node(n,m), Cq;n,m is the capacitance and EJq;n,m the Joseph-son energy of the qubit at that node. ϕa;n,m and ϕb;n,mare the phases at the auxiliary nodes next to the SQUIDlocated between nodes (n,m), (n,m+ 1), (n+ 1,m), and(n + 1,m + 1). The Cg are the capacitances of theseauxiliary nodes with respect to ground and CJ is thecapacitance of each junction in the SQUID including ashunt capacitance parallel to the junctions. EJ is theJosephson energy of each junction in the SQUID.

As the Josephson junctions of the dc-SQUIDs only cou-ple to the difference between the phases of the two adja-cent auxiliary nodes, we define the modes,

ϕ±;n,m = ϕa;n,m ± ϕb;n,m, (A2)

and describe the dc-SQUIDs in terms of ϕ+;n,m andϕ−;n,m from now on. We now turn to derive the Hamil-tonian associated to the Lagrangian in Eq. (A1).

As there is neither capacitive coupling between thequbits in the lattice nor between the qubits and the cou-pling SQUIDs, the kinetic energy can be decomposed intothe contribution of individual qubits and that of the in-dividual coupling SQUIDs,

K =

N,M∑n,m=1

φ20

2

(Cq;n,mϕ

2n,m + ~ΦTs;n,mCs ~Φs;n,m

), (A3)

where ~Φs;n,m = (ϕ+;n,m, ϕ−;n,m)T and

Cs =

(2CJ +

Cg2 0

0Cg2

)(A4)

is the capacitive matrix associated with one SQUID.

The Hamiltonian of the lattice is obtained in the stan-dard way via a Legendre transform of the Lagrangian Lin Eq. (A1) by defining the qubit and SQUID momenta,

πn,m =∂L

∂ϕn,m, and π±;n,m =

∂L∂ϕ±;n,m

. (A5)

The kinetic energy thus reads,

K =

N,M∑n,m=1

[π2n,m

2~φ20Cq;n,m

+π2−;n,m

4~φ20CJ

+π2

+;n,m

~φ20Cg

](A6)

where we have used that Cg � CJ in inverting Cs. Wethus arrive at the Hamiltonian,

H =

N,M∑n,m=1

[4ECq;n,m

~2π2n,m + 4ELϕ

2n,m − EJq;n,m cos(ϕn,m)

](A7)

+

N,M∑n,m=1

[4EC+

~2π2

+;n,m + 4EC−~2

π2−;n,m + ELϕ

2+;n,m + ELϕ

2−;n,m − 2EJ cos

(ϕext2

)cos(ϕ−;n,m)

]

−M,N∑n,m=1

EL [ϕq+;n,mϕ+;n,m + ϕq−;n,mϕ−;n,m] ,

where we have introduced the notation

ϕq±;n,m = ϕn,m + ϕn,m+1 ± ϕn+1,m ± ϕn+1,m+1. (A8)

EL = φ20/(2L) is the inductive energy associated with an

inductor L and the charging energies of the qubits and

12

SQUID modes are

ECq;n,m =e2

2Cq;n,m, EC+ =

e2

Cgand EC− =

e2

4CJ,

where e is the elementary charge. The first line inEq. (A7) describes the qubits, the second line theSQUIDs and the third line the interactions betweenqubits and SQUIDs via the inductances L.

The Hamiltonian (A7) is quantized in the standardway by imposing the canonical commutation relations

[ϕj , πl] = i~δj,l and [ϕ±;j , π±;l] = i~δj,l. (A9)

The external flux applied through the SQUID loops andhence the associated phase is composed of a constant dc-part and a time dependent ac-part as given in Eq. (6).We choose ϕdc = π and ϕac � ϕdc = π, and obtain toleading order in ϕac,

cos(ϕext

2

)≈ −ϕac

2F (t). (A10)

Appendix B: Effective time-dependent Hamiltonian

In this appendix we present details of the derivationof the effective Hamiltonian for the qubits only, whichwill be dominated by the four-body interactions corre-sponding to the stabilizer operators in the Toric CodeHamiltonian, see Eq. (1).

As the modes ϕ±;n,m of the SQUIDs oscillate at fre-quencies that strongly differ from the transition frequen-cies of the qubits and the oscillation frequencies of theapplied flux biases, they remain in their ground statesto a high degree of precision and can be adiabaticallyeliminated from the description. To this end we applythe Schrieffer-Wolf transformation introduced in Eq. (7)with the generator

S =i

2~

N,M∑n,m=1

(ϕq+;n,mπ+;n,m + ϕq−;n,mπ−;n,m), (B1)

and expand the transformed Hamiltonian up to fourthorder in S,

H = eSHe−S = H+ [S,H] +1

2![S, [S,H]] +

1

3![S, [S, [S,H]]] +

1

4![S, [S, [S, [S,H]]]] + . . . (B2)

Using the canonical commutation relations (A9), thecommutators in the expansion (B2) read,

[S,H] =

=

N,M∑n,m=1

ELϕq+;n,mϕ+;n,m + ELϕq−;n,mϕ−;n,m

− 4

N,M∑n,m=1

ECq;n,mπn,m (πS−;n,m + πS+;n,m)

−N,M∑n,m=1

EL2

(ϕ2q+;n,m + ϕ2

q−;n,m)

+

N,M∑n,m=1

EJ cos(ϕext

2

)sin(ϕ−;n,m)ϕq−;n,m

[S, [S,H]] =

=

N,M∑n,m=1

EL2

(ϕ2q+;n,m + ϕ2

q−;n,m)

+ 2

N,M∑n,m=1

ECq;n,m (πS−;n,m + πS+;n,m)2

+

N,M∑n,m=1

EJ2

cos(ϕext

2

)cos(ϕ−;n,m)ϕ2

q−;n,m

[S, [S, [S,H]]] =

= −N,M∑n,m=1

EJ4

cos(ϕext

2

)sin(ϕ−;n,m)ϕ3

q−;n,m

[S, [S, [S, [S,H]]]] =

= −N,M∑n,m=1

EJ8

cos(ϕext

2

)cos(ϕ−;n,m)ϕ4

q−;n,m

where we have here introduced the abbreviations,

πS±;n,m = π±;n,m + π±;n,m−1 ± π±;n−1,m ± π±;n−1,m−1.(B3)

We note that the terms [S, [S, [S,H]]] and[S, [S, [S, [S,H]]]] do not contain any contributionsfrom the modes ϕ+;n,m which are just harmonic os-cillators. They would vanish for an entirely harmoniccoupling circuit and only the Josephson terms of thecoupling SQUIDs that are nonlinear in ϕ−;n,m lead tocontributions.

Truncating the expansion (B2) at fourth order is agood approximation since the amplitudes of the termsϕn,mπ±;n,m/~ are much smaller than unity, see Eq. (B16)for details. Moreover, as we explain in the sequel, the fre-quencies contained in the flux bias φext that is appliedto the SQUIDs can be chosen such that interactions con-tained in the fourth order terms [S, [S, [S, [S,H]]]] becomethe dominant terms of the expansion (B2).

As stated in Eq. (8) of Sec. II A, the transformed

Hamiltonian H can be written as a sum of a part H0q

describing the individual qubits, a part HIq describing

13

pure qubit-qubit interactions, a part HS describing thecoupling SQUIDs and a part HqS describing the residualinteractions between qubits and SQUIDs,

H = H0q +HIq +HS +HqS . (B4)

The Hamiltonian of the individual qubits, H0q, and thequbit-qubit interactions, HIq, are given in Eqs. (8) and(9). We here repeat them for completeness,

H0q =

N,M∑n,m=1

Hq;n,m with (B5)

Hq;n,m =4ECq;n,m

~2π2n,m + 2ELϕ

2n,m − EJq;n,m cos(ϕn,m)

and

HIq = −2EL

N,M∑n,m=1

ϕn,mϕn,m+1 (B6)

The term describing the SQUIDs reads,

HS =

M,N∑n,m=1

[4EC+;n,m

~2π2

+;n,m + ELϕ2+;n,m

](B7)

+

M,N∑n,m=1

[4EC−;n,m

~2π2−;n,m + ELϕ

2−;n,m

]

− 2EJ cos(ϕext

2

) M,N∑n,m=1

cos(ϕ−;n,m),

where EC+;n,m and EC−;n,m are the renormalized charg-

ing energies of the + and−modes, EC±;n,m = EC±;n,m+(Eq;n,m + Eq;n,m+1 + Eq;n+1,m + Eq;n+1,m+1)/4. Thequbit-SQUID interactions read,

HqS =− 4

N,M∑n,m=1

ECq;n,m~2

πn,mπS−;n,m (B8)

+ EJ

N,M∑n,m=1

cos(ϕext

2

)sin(ϕ−;n,m)ϕq−;n,m

+1

2!

EJ2

N,M∑n,m=1

cos(ϕext

2

)cos(ϕ−;n,m)ϕ2

q−;n,m

− 1

3!

EJ4

N,M∑n,m=1

cos(ϕext

2

)sin(ϕ−;n,m)ϕ3

q−;n,m

− 1

4!

EJ8

N,M∑n,m=1

cos(ϕext

2

)cos(ϕ−;n,m)ϕ4

q−;n,m,

where again we have neglected terms ∝ πn,mπS+;n,m be-cause |ω+ − ωn,m| � |ω− − ωn,m|, see Eq. (B15), whichmakes these terms highly off-resonant so that we onlykeep the dominant contributions ∝ πn,mπS−;n,m.

1. Second-quantized form

The free Hamiltonian of the qubits can be written insecond-quantized form by defining the ladder operators

qi and q†i via

ϕj = ϕj(qj + q†j ), (B9)

πj = − i~2ϕj

(qj − q†j ), with

ϕj =

(2ECq;j

EJq;j + 4EL

)1/4

.

Here, qj (q†j ) is the annihilation (creation) operator for

qubit j [j = (1, 1), . . . , (N,M)] and the effective Joseph-son energy of the qubits is EJq;j +4EL due to the contri-butions of the four adjacent inductors. In terms of thesecreation and annihilation operators, H0q reads,

H0q =

(N,M)∑j=(1,1)

[~ωjq†jqj − ~Ujq†jq

†jqjqj

](B10)

where we have approximated cos(ϕj) by its expansionup to fourth order in ϕj and the transition frequenciesωj and nonlinearities Uj are given by

ωj =~−1√

8ECq;j(EJq;j + 4EL), and (B11)

Uj =ECq;j

2~EJq;j

EJq;j + 4EL. (B12)

For the SQUID modes we introduce the annihilation (cre-

ation) operators s± (s†±) according to,

ϕ±;j = ϕ±(s±;j + s†±;j), (B13)

π±;j = − i~2ϕ±

(s±;j − s†±;j), with

ϕ± =

(EC±EL

)1/4

,

and get

HS =

(N,M)∑j=(1,1)

[~ω+s

†+;js+;j + ~ω−s†−;js−;j

], (B14)

where

ω± = 4~−1

√EC±EL (B15)

and we have neglected the last term of Eq. (B7) sincecos (ϕext/2) ≈ −(ϕac/2)F (t) and the frequencies con-tained in F (t) strongly differ from ω+ and ω−. Note

that CJ � Cg and thus EC+ � EC−, so that ω+ � ω−.

14

2. Accuracy of truncating the transformation

After having introduced the amplitudes ϕj and ϕ± inEqs. (B9) and (B13), we can now estimate the amplitudeof the Schrieffer-Wolff generator S, see Eqs. (7) or (B1),to check whether a truncation of the expansion (B2) atfourth order provides a good approximation. For thescenario with at most a single excitation in each qubitand vanishingly small excitation numbers in the SQUIDmodes, which is of interest here, the generator S has amagnitude given by

1

4

ϕjϕ±

(B16)

for each site. Typical parameters of our scheme lead toϕj . 0.7 and ϕ± & 1.65 so that ϕj/4ϕ± ≈ 0.1 and thetruncation of the expansion in Eq. (B2) is well justified.

Appendix C: Effective Hamiltonian in the rotatingframe

To analyze which terms of the transformed Hamilto-nian (8) respectively (B4) provide the dominant contri-butions, we now move to a rotating frame, where allqubits rotate at their transition frequencies ωn,m and theSQUID modes rotate at their oscillation frequencies ω±.In transforming to this rotating frame, the raising andlowering operators transform as,

qj → e−iωjtqj , and s±;j → eiω±ts±;j (C1)

for j = (1, 1), . . . (N,M).

1. Adiabatic elimination of the SQUID modes

For the parameters of our architecture, the SQUID fre-quencies ω± differ strongly from the qubit transition fre-quencies ωj and the frequencies contained in the drivingfields. The SQUID modes ϕ+;n,m and ϕ−;n,m thereforeremain to a very good approximation in their groundstates, |0+;n,m〉 and |0−;n,m〉. They can thus be adiabat-ically eliminated from the description, where the domi-nant contributions come from the ϕ−;n,m modes becauseω+ � ω−.

The leading order of the adiabatic elimination is ob-tained by projecting HS and HqS onto the ground states

of the SQUIDs, |0SQUIDs〉 =∏N,Mn,m=1 |0−;n,m, 0+;n,m〉.

Whereas HS only gives rise to an irrelevant constant,

we get for H(0)qS = 〈0SQUIDs|HqS |0SQUIDs〉 the expression

given in Eq. (11), which we here repeat for completeness

H(0)qS =− 1

8EJF (t)

N,M∑n,m=1

ϕ2q−;n,m

+1

16

1

4!EJF (t)

N,M∑n,m=1

ϕ4q−;n,m,

(C2)

where,

EJ = χϕacEJ with (C3)

χ = exp

(−ϕ2−;n,m

2

)(C4)

In deriving Eq. (C2) we have used the ground state ex-pectation values

〈0−;n,m|π−;n,m |0−;n,m〉 = 0,

〈0−;n,m| cos(ϕ−;n,m) |0−;n,m〉 = χ,

〈0−;n,m| sin(ϕ−;n,m) |0−;n,m〉 = 0.

(C5)

The HamiltonianH(0)qS is a sum of contributions from each

cell formed by four qubits at its corners which interactvia one coupling SQUID.

When considering the terms of one cell only, as de-scribed in Sec. II A, and labelling its qubits by indices1,2,3 and 4, i.e. (n,m) → 1, (n + 1,m) → 2, (n + 1,m +1) → 3, (n,m + 1) → 4, one can divide the Hamiltonian

H(0)qS for one cell into three parts as described in Eq. (12),

H(0)qS

∣∣∣cell

= Hcell = H1 +H2 +H3, (C6)

where

H1 =1

16EJF (t)ϕ1ϕ2ϕ3ϕ4,

H2 = −1

8EJF (t)(ϕ1 + ϕ4 − ϕ2 − ϕ3)2,

H3 = −EJF (t)∑

(ijkl)

Cijklϕiϕjϕkϕl

Here, H1 contains the four-body interactions which willgive rise to stabilizer interactions, H2 contains all two-body interactions and H3 all fourth-order interactionsbeyond the four-body interaction, i.e. terms where atleast two of the indices i, j, k and l are the same as weindicate by the notation

∑(ijkl). The coefficients Cijkl

have the values Cijkl = ±1/32 for terms of the fromϕ2jϕkϕl with j 6= k, j 6= l and k 6= l, Cijkl = ±1/64 for

terms of the from ϕ2jϕ

2k with j 6= k, Cijkl = ±1/96 for

terms of the from ϕ3jϕk with j 6= k and Cijkl = ±1/384

for terms of the from ϕ4j . Here, “+” signs apply whenever

an even number of the indices i, j, k, l refer to the sameside of the SQUID (i.e. the same diagonal) and “−” signsotherwise. Whereas we discuss perturbations originatingfrom H2 and H3 in appendix D, we now turn to discussH1 in more detail.

2. Four-body interactions

Due to the nonlinearities of the qubits, we can hererestrict the description to the two lowest energy levels

of each node by replacing qj → σ−j (q†j → σ+j ). The

15

Hamiltonian H1 contains 16 terms of the form S1S2S3S4,where Sj ∈ {e−iωjtσ−j , eiωjtσ

+j } (see C6). Ignoring the

oscillating pre-factor F (t), these terms rotate at the an-gular frequencies as listed in table II. Denoting the sum

term frequency term frequency

σ+1 σ

+2 σ

+3 σ

+4 ν1 σ+

1 σ+2 σ

+3 σ−4 ν5

σ−1 σ−2 σ−3 σ−4 −ν1 σ+

1 σ+2 σ−3 σ

+4 ν6

σ+1 σ

+2 σ−3 σ−4 ν2 σ+

1 σ−2 σ

+3 σ

+4 ν7

σ+1 σ−2 σ

+3 σ−4 ν3 σ−1 σ

+2 σ

+3 σ

+4 ν8

σ+1 σ−2 σ−3 σ

+4 ν4 σ−1 σ

−2 σ−3 σ

+4 −ν5

σ−1 σ+2 σ

+3 σ−4 −ν4 σ−1 σ

−2 σ

+3 σ−4 −ν6

σ−1 σ+2 σ−3 σ

+4 −ν3 σ−1 σ

+2 σ−3 σ−4 −ν7

σ−1 σ−2 σ

+3 σ

+4 −ν2 σ+

1 σ−2 σ−3 σ−4 −ν8

frequencies:ν1 = ω1 + ω2 + ω3 + ω4

ν2 = ω1 + ω2 − ω3 − ω4

ν3 = ω1 − ω2 + ω3 − ω4

ν4 = ω1 − ω2 − ω3 + ω4

ν5 = ω1 + ω2 + ω3 − ω4

ν6 = ω1 + ω2 − ω3 + ω4

ν7 = ω1 − ω2 + ω3 + ω4

ν8 = −ω1 + ω2 + ω3 + ω4

TABLE II. The four-body terms contained in H1 of Eq. (C6)and their rotation frequencies.

of the terms in the first column of table II by V1 andthe sum of the terms in the second column by V2 it canreadily be verified that V1 + V2 = AS = σx1σ

x2σ

x3σ

x4 and

V1 − V2 = BP = σy1σy2σ

y3σ

y4 . Motivated by the specific

form of the Hamiltonian H1, see Eq. (C6), we thus definetwo oscillating fields FX and FY ,

FX(t) = f1,1,1,1 + f1,1,−1,−1 + f1,−1,1,−1 + f1,−1,−1,1

FY (t) = f1,1,1,−1 + f1,1,−1,1 + f1,−1,1,1 + f−1,1,1,1

(C7)

with fa,b,c,d = cos[(aω1 + bω2 + cω3 + dω4)t]. Each ofthe terms in FX (FY ) brings two of the four-body pro-cesses in the first (second) column of table II into res-onance (makes them non-rotating). Next to resonantterms, each field also generates rotating terms. The ro-tation frequency of these terms is the detuning betweendifferent components of FX and FY . As the least detun-ing is in the GHz range (∼ 0.4 GHz) and the strengthof four-body interaction is in the MHz range (∼ 1 MHz),the rotating terms can be neglected in a RWA. Hence theexternal oscillating field Fs = FX + FY gives rise to thestar interaction As and the external field Fp = FX − FYgives rise to the plaquette interaction Bp,

F (t) =

{Fs = FX + FY ⇒ H1 ∝ As = σx1σ

x2σ

x3σ

x4

Fp = FX − FY ⇒ H1 ∝ Bp = σy1σy2σ

y3σ

y4

(C8)and the strength of the stabilizer interaction is

Js = Jp =1

16EJ

4∏i=1

ϕi (C9)

Besides these stabilizer interactions, there are also two-body interactions between the qubits which may lead tocorrections in the effective Hamiltonian (1). We now turnto estimate their effect.

Appendix D: Perturbations

In this section we estimate the effects of the remain-ing terms of the transformed Hamiltonian H, see Eq. (8)or (B4). These include contributions contained in theHamiltonians HIq, see Eq. (B6), and HqS , see Eq. (B8).In our approach, the transition frequencies of the qubitsωn,m and the frequency spectrum of the oscillating fluxesνk are chosen such the all undesired terms are highly off-resonant and therefore strongly suppressed. They how-ever lead to a frequency shift

∆j = ∆(1)j + ∆

(2)j + ∆

(3)j + ∆

(4)j (D1)

for each qubit j, where ∆(1)j ,∆

(2)j ,∆

(3)j and ∆

(4)j are given

in Eqs. (D8), (D11),(D13) and (D18). These frequencyshifts can easily be compensating for by modifying thefrequency spectrum of the applied driving fields accord-ingly, so that we did not include them in our precedinganalysis. They can be calculated via the following timedependent perturbation theory.

In the frame where all qubits rotate at their transi-tion frequencies, the transformed Hamiltonian H can bewritten as

H = HTC +∑α

hαeiωαt, (D2)

where HTC is the Toric Code Hamiltonian as introducedin Eq. (1) and the remaining terms have been written as

a sum of all their frequency components. Hence hα is thesum of all terms that rotate as eiωαt, where ωα containsall contributing frequencies, i.e. the oscillation frequen-cies of the operators and, for terms with a prefactor F (t),the frequency of the oscillating flux bias. Note that both,ωα > 0 and ωα < 0 contribute to the sum in Eq. (D2)

since H is Hermitian.Following approaches outlined in [73–75], the dynamics

generated by a Hamiltonian of the form as in Eq. (D2)

with ||hα|| � ~ωα can be approximated by a time in-dependent effective Hamiltonian that is an expansion inpowers of the interaction strength over oscillation fre-quency [74],

Heff = Heff,0 +Heff,1 +Heff,2 +Heff,3 + . . . , (D3)

where the individual contributions read

Heff,0 = HTC

Heff,1 =∑α,β

δ(ωα + ωβ)

2~ωβ

[hα, hβ

]

16

Heff,2 =∑α,β

δ(ωα + ωβ)

2~2ωαωβ

[[HTC, hα

], hβ

]+∑α,β,γ

δ(ωα + ωβ + ωγ)

3~2ωβωγ

[[hα, hβ

], hγ

]Heff,3 =

∑α,β,γ

δ(ωα + ωβ + ωγ)

3~3ωαωβωγ

[[[HTC, hα

], hβ

], hγ

]+∑

α,β,γ,δ

δ(ωα + ωβ + ωγ + ωδ)

4~3ωβωγωδ

[[[hα, hβ

], hγ

], hδ

]and δ(ω) is the Dirac δ-function.

Since ||hα|| � ~ωα, the first term of Heff,2 and thefirst term of Heff,3 are at least two orders of magnitudesmaller than HTC and we neglect them. We now turn todiscuss the other perturbations to HTC one by one.

1. Corrections to the adiabatic elimination of theSQUIDs

Besides the terms contained in H(0)qs that are found

by projecting the Hamiltonian H onto the ground statesof the SQUID modes, there are first-order correctionsto the adiabatic elimination. These are processes medi-ated by the creation and annihilation of a virtual ex-citation in a SQUID mode. To estimate their effect,we first note that terms proportional to cos(ϕ−;n,m) donot create single excitations in the SQUID modes since〈0−;n,m| cos(ϕ−;n,m) |1−;n,m〉 = 0 and calculate the ma-trix elements

〈0−;n,m| sin(ϕ−;n,m) |0−;n,m〉 = 0,

〈0−;n,m| sin(ϕ−;n,m) |1−;n,m〉 = χ,

〈1−;n,m| sin(ϕ−;n,m) |1−;n,m〉 = 0.

(D4)

In the subspace of at most one excitation, we may thusreplace

sin(ϕ−;n,m)→ χ(s−;n,m + s†−;n,m) =χ

ϕ−ϕ−;n,m, (D5)

to approximate ∆HqS = HqS −H(0)qS , c.f. Eq. (C6), as

∆HqS ≈− 4

N,M∑n,m=1

ECq;n,m~2

πn,mπS−;n,m (D6)

− EJ2ϕ−

F (t)

N,M∑n,m=1

ϕ−;n,mϕq−;n,m

where we have neglected the terms in Eq. (B8) that areproportional to ϕ3

q−;n,m as these are a factor ϕ2n,m/24�

1 smaller than the terms that are linear in ϕq−;n,m. Wenow consider corrections to HTC coming from second or-der processes of ∆HqS .

Using the formulation in second quantization, seeEqs. (B9) and (B13), one finds for the first term of ∆HqS

that the perturbation theory of Eq. (D3) is applicable if∣∣∣∣ ECq;j~(ω− − ωj)ϕjϕ−

∣∣∣∣� 1. (D7)

For this regime, the effective Hamiltonian Heff,1 accord-ing to Eq. (D3) for the first term of ∆HqS just containsa frequency shift

∆(1)j = −

4E2Cq;j

~2(ω− − ωj)ϕ2jϕ

2−. (D8)

for each qubit j = (n,m). Here, the prefactor 4 ac-counts for the fact that each qubit has 4 neighboringSQUIDs. For the set of parameters discussed in Sec. III,we find for the qubits with high transition frequencies

ECq;j/(hϕjϕ−) ∼ 500 MHz leading to ∆(1)j ∼ 200 MHz,

i.e. a shift of 50 MHz from each coupling to a SQUID.For qubits with low transition frequencies the shifts are

∆(1)j ∼ 100 MHz, i.e. a shift of 25 MHz from each cou-

pling to a SQUID. Besides frequency shifts, the first termof ∆HqS can also lead to interactions between two qubitsof one cell. These are ineffective provided any pair ofqubits in a cell is sufficiently detuned from each other,∣∣∣∣∣ ECq;jECq;l

~2(ωl − ωj)(ω− − ωj)ϕjϕlϕ2−

∣∣∣∣∣� 1, (D9)

which holds since any pair of qubits in one cell is at leastdetuned by 700 MHz from each other, see table I. Notethat the detuning between two neighboring qubits on adiagonal n that would couple via two SQUIDs is signif-icantly higher (∼ 10 GHz). The next higher order non-vanishing term of the expansion (D3) is here the secondterm of Heff,3 which is more than two orders of magnitude

smaller than ∆(1)j .

In a similar way as second-order terms can lead toqubit-qubit interactions in one cell, fourth-order termscould mediate interactions between two qubits in neigh-boring cells. For our parameters, these interactionswould have a strength of ∼ 0.5 MHz but are ineffectiveas qubits in neighboring cells are always detuned by atleast 100 MHz.

Via an analogous calculation, one finds for the secondterm of ∆HqS that the perturbation theory of Eq. (D3)is applicable if ∣∣∣∣∣ EJϕj

2~[(ω− ± ωj)± νk]

∣∣∣∣∣� 1 (D10)

and leads to the frequency shifts (taking into accountthat each qubit couples to 4 SQUIDs)

∆(2)j = −2

E2Jϕ

2j

~2

8∑k=1

∑x=0,1

ω− + (−1)xωj[ω− + (−1)xωj ]2 − ν2

k

(D11)

for each qubit j, where the frequencies νk that are con-tained in the driving field F (t) are defined in table II.

17

For the set of parameters as given in Sec. III, we find

|∆(2)j | . 20 MHz. The next higher order non-vanishing

term of the expansion (D3) is here the second term ofHeff,3 which is more than two orders of magnitude smaller

than ∆(2)j .

2. Perturbations contained in HIq

The transformed Hamiltonian H, see Eq. (8) or (B4)contains residual interactions between neighboring qubitson each diagonal n as described in Eq. (B6). Theirstrength is 2ELϕn,mϕn,m+1 for the interaction betweenqubits (n,m) and (n,m + 1) on diagonal n. Hence pro-vided that ∣∣∣∣ 2ELϕn,mϕn,m+1

~(ωn,m − ωn,m+1)

∣∣∣∣� 1 (D12)

the perturbation theory of Eq. (D3) is applicable andleads to the frequency shift

∆(3)n,m =

4E2Lϕ

2n,mϕ

2n,m−1

~2(ωn,m − ωn,m−1)+

4E2Lϕ

2n,mϕ

2n,m+1

~2(ωn,m − ωn,m+1),

(D13)for qubit (n,m). For the parameters discussed in Sec. III,we find 2ELϕn,mϕn,m+1/h ≈ 300 MHz whereas neighbor-ing qubits on a diagonal are detuned by almost 10 GHz so

that |∆(3)n,m| ∼ 18 MHz. In a higher order of the pertur-

bation theory of Eq. (D3), each qubit (n,m) can also me-diate an interaction between its two neighboring qubitson the diagonal n due to the couplings 2ELϕn,m−1ϕn,mand 2ELϕn,mϕn,m+1. To suppress these interactions ofstrength 10 MHz, next-nearest neighbor qubits need tobe sufficiently detuned from one another,∣∣∣∣∣ 4E2

Lϕn,m−1ϕ2n,mϕn,m+1

~2(ωn,m − ωn,m+1)(ωn,m−1 − ωn,m+1)

∣∣∣∣∣� 1, (D14)

which requires |ωn,m−1 − ωn,m+1| & 2π × 100 MHz asfulfilled by the frequencies in table I.

Furthermore, there could be interactions betweenqubits in neighboring cells, say qubit (n,m) and qubit(n,m + 2), since qubits (n,m) and qubit (n,m + 1)couple via 2ELϕn,mϕn,m+1 and qubits (n,m + 1) andqubit (n,m + 2) couple in a second order process viatheir interaction with a common SQUID as discussed inSec. D 1. This third-order interactions are strongly sup-pressed since their strength is ∼ 3 MHz and thus 30 timessmaller than the detuning between the respective qubits.

3. Residual two-body interactions in H(0)qS

When projected onto the ground states of the SQUIDmodes in the zeroth order of adiabatic elimination, the

Hamiltonian H(0)qS = 〈0SQUIDs|HqS |0SQUIDs〉 contains

terms that are quadratic in ϕq−;n,m and thus contain two-body interactions, see Eq. (C2). In the rotating frame,the two-body interaction Hamiltonian of one cell as givenby H2 in Eq. (C6) can be written as,

H2 =

4∑j,l=1

8∑k=1

Bj,l

{[ei(ωj−ωl+νk)t + ei(ωj−ωl−νk)t

]q†jql

+[ei(ωj+ωl+νk)t + ei(ωj+ωl−νk)t

]q†jq†l + H.c.

},

(D15)

where the coupling coefficients read Bj,l = ∓EJϕjϕl/8,where the “−” sign applies for j = l and (j, l) = (1, 4) or(2, 3), and the “+” sign applies otherwise. The pertur-bation theory of Eq. (D3) applies to H2 provided that∣∣∣∣∣ EJϕjϕl

8~(ωj ± ωl ± νk)

∣∣∣∣∣� 1 (D16)

for all j, l and k, c.f. Eq. (D15). Hence the spectrumof the driving fields must be chosen such that conditions(D10) and (D16) are met.

To estimate the corrections due to H2 as written inEq. (D15), one thus needs to consider the commutationrelations,

[qjq†l , q†jql] = (1− δj,l)

(q†l ql − q

†jqj

), (D17)

[qjql, q†jq†l ] = (1 + δj,l)

(q†jqj + q†l ql + 1

),

[qjql, qjq†l ] = (1 + δj,l) qjqj ,

[qjq†l , q†jq†l ] = (1 + δj,l) q

†l q†l ,

as these are the only ones that, together with their os-cillating prefactors Fs or Fp, lead to non-rotating terms.The two commutators in the last two lines of Eq. (D17)would lead to transitions out of the qubit subspace andcan thus be ignored because of the nonlinearity of thequbits.

Hence, to leading order the Hamiltonian H2 will leadto shifts of the transition frequencies of the qubits. Tak-ing into account that each cell is embedded into the to-tal lattice and hence each qubit experiences frequencyshifts from the two-body interactions in all four cells it ispart of, we find that the qubit transition frequencies areshifted by

∆(4)j = − E

2J

8

∑{l:|j−l|=1}

8∑k=1

ϕ4j

2ωj4ω2

j − ν2k;j,l

(D18)

− E2J

32

∑{l:|j−l|=1}

8∑k=1

ϕ2jϕ

2l [ωj + ωl]

[ωj + ωl]2 − ν2k;j,l

,

+E2J

32

∑{l:|j−l|=1}

8∑k=1

ϕ2jϕ

2l [ωj − ωl]

[ωj − ωl]2 − ν2k;j,l

,

where the sum∑{l:|j−l|=1} runs over all qubits l that are

nearest neighbors to qubit j and νk;j,l denotes the drive

18

frequencies at the SQUID that connects qubits j and l.For the set of parameters as given in Sec. III, we find

|∆(4)j | . 30 MHz.

Appendix E: Expected Measurement Outcomes

In this section we discuss a set of possible measurementout comes for the topologically ordered states on a eight-

qubit lattice, c.f. Fig. 3. One ground state of the ToricCode Hamiltonian on this lattice can be written as

|Ψ1〉 =1

4√

2(1 + σy1σ

y4σ

y5σ

y3 )(1 + σy2σ

y3σ

y6σ

y4 )(1 + σy5σ

y8σ

y1σ

y7 )(1 + σy6σ

y7σ

y2σ

y8 )

8⊗j=1

|Xj〉 , (E1)

where |Xj〉 is the eigenstate of σxj with eigenvalue 1,σxj |Xj〉 = |Xj〉. The other three ground states can be

found by applying the loop operators σy5σy6 , σy4σ

y8 and

σy4σy8σ

y5σ

y6 ,

|Ψ2〉 = σy5σy6 |Ψ1〉 (E2)

|Ψ3〉 = σy4σy8 |Ψ1〉

|Ψ4〉 = σy4σy8σ

y5σ

y6 |Ψ1〉 .

Whereas the reduced density matrices of individual spinsare all completely mixed (i.e. proportional to an identitymatrix) for all these four states, they can be distinguishedvia their two spin correlations, see table III.

[1] H. L. Stormer, D. C. Tsui, and A. C. Gossard, Thefractional quantum Hall effect, Rev. Mod. Phys. 71, 298(1999).

[2] L. Balents, Spin liquids in frustrated magnets, Nature464, 199 (2010).

[3] C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S.Das Sarma, Non-Abelian anyons and topological quantumcomputation, Rev. Mod. Phys. 80, 1083 (2008).

[4] A. Kitaev, Fault-tolerant quantum computation byanyons, Ann. Phys. 303, 2 (2003).

[5] A. Kitaev and C. Laumann, Topological phases and quan-tum computation, arXiv:0904.2771

[6] A. G. Fowler, M. Mariantoni, J. M. Martinis, and A.N. Cleland, Surface codes: Towards practical large-scalequantum computation, Phys. Rev. A 86, 032324 (2012).

[7] E. Dennis, A. Kitaev, A. Landahl, and J. Preskill, Topo-logical quantum memory J. Math. Phys. 43, 4452 (2002).

[8] The convention we use here is motivated by the usualnotation for superconducting qubits but differs from con-ventions mostly used when discussing computational as-pects of the Toric Code.

[9] L. B. Ioffe, M. V. Feigel’man, A. Ioselevich, D. Ivanov, M.Troyer, and G. Blatter, Topologically protected quantumbits using Josephson junction arrays, Nature 415, 503(2002)

[10] S. Gladchenko, D. Olaya, E. Dupont-Ferrier, B. Doucot,L. B. Ioffe, M. E. Gershenson, Superconducting nanocir-cuits for topologically protected qubits, Nat. Phys. 5, 48(2009).

[11] X.-C. Yao, T.-X. Wang, H.-Z. Chen, W.-B. Gao, A. G.Fowler, R. Raussendorf, Z.-B. Chen, N.-L. Liu, C.-Y.Lu, Y.-J. Deng, Y.-A. Chen, and J.-W. Pan, Experimen-

tal demonstration of topological error correction, Nature482, 489 (2012).

[12] D. Nigg, M. Muller, E. A. Martinez, P. Schindler, M.Hennrich, T. Monz, M. A. Martin-Delgado, R. Blatt,Quantum computations on a topologically encoded qubit,Science 345, 302 (2014).

[13] J. Kelly, R. Barends, A. G. Fowler, A. Megrant, E. Jef-frey, T. C. White, D. Sank, J. Y. Mutus, B. Campbell,Yu Chen, Z. Chen, B. Chiaro, A. Dunsworth, I.-C. Hoi,C. Neill, P. J. J. O’Malley, C. Quintana, P. Roushan, A.Vainsencher, J. Wenner, A. N. Cleland, and J. M. Mar-tinis, State preservation by repetitive error detection in asuperconducting quantum circuit, Nature 519, 66 (2015).

[14] D. Riste, S. Poletto, M.-Z. Huang, A. Bruno, V. Vester-inen, O.-P. Saira, and L. DiCarlo, Detecting bit-flip er-rors in a logical qubit using stabilizer measurements, Nat.Commun. 6, 6983 (2015).

[15] A. D. Corcoles, Easwar Magesan, Srikanth J. Srinivasan,Andrew W. Cross, M. Steffen, Jay M. Gambetta, andJerry M. Chow, Detecting arbitrary quantum errors viastabilizer measurements on a sublattice of the surfacecode, Nat. Comm. 6, 6979 (2015).

[16] D. Marcos, P. Rabl, E. Rico, and P. Zoller, Supercon-ducting Circuits for Quantum Simulation of DynamicalGauge Fields, Phys. Rev. Lett. 111, 110504 (2013).

[17] M. Hafezi, P. Adhikari, and J. M. Taylor, Engineeringthree-body interaction and Pfaffian states in circuit QEDsystems, Phys. Rev. B 90, 060503(R) (2014)

[18] G. K. Brennen, G. Pupillo, E. Rico, T. M. Stace, andD. Vodola, Loops and strings in a superconducting latticegauge simulator, arXiv:1512.06565 (2015)

19

〈σx1σ

x2 〉 〈σx

3σx4 〉 〈σx

5σx6 〉 〈σx

7σx8 〉 〈σx

3σx7 〉 〈σx

1σx5 〉 〈σx

4σx8 〉 〈σx

2σx6 〉

|Ψ1〉 0 1 0 1 0 1 0 1|Ψ2〉 0 1 0 1 0 -1 0 -1|Ψ3〉 0 -1 0 -1 0 1 0 1|Ψ4〉 0 -1 0 -1 0 -1 0 -1

TABLE III. Two-qubit correlations along non-contractible loops for the 4 ground states of the Toric Code on an 8-qubit lattice.

[19] W. Lechner, P. Hauke and P. Zoller, A quantum anneal-ing architecture with all-to-all connectivity from local in-teractions, Sci. Adv. 1, e1500838 (2015).

[20] N. Chancellor, S. Zohren, and P. A. Warburton, Cir-cuit design for multi-body interactions in superconductingquantum annealing system with applications to a scalablearchitecture, arXiv:1603:09521 (2016)

[21] M. Leib, P. Zoller, and W. Lechner, A Transmon Quan-tum Annealer: Decomposing Many-Body Ising Con-straints Into Pair Interactions, arXiv:1604.02359 (2016)

[22] Y. Chen, P. Roushan, D. Sank, C. Neill, Erik Lucero,Matteo Mariantoni, R. Barends, B. Chiaro, J. Kelly, A.Megrant, J. Y. Mutus, P. J. J. O’Malley, A. Vainsencher,J. Wenner, T. C. White, Yi Yin, A. N. Cleland, andJohn M. Martinis, Simulating weak localization using su-perconducting quantum circuits Nat. Commun. 5, 5184(2014)

[23] R. Barends, L. Lamata, J. Kelly, L. Garca-lvarez, A. G.Fowler, A. Megrant, E. Jeffrey, T. C. White, D. Sank, J.Y. Mutus, B. Campbell, Yu Chen, Z. Chen, B. Chiaro,A. Dunsworth, I.-C. Hoi, C. Neill, P. J. J. O’Malley, C.Quintana, P. Roushan, A. Vainsencher, J. Wenner, E.Solano, and John M. Martinis, Digital quantum simula-tion of fermionic models with a superconducting circuit,Nat. Commun. 6, 7654 (2015).

[24] Y. Salathe, M. Mondal, M. Oppliger, J. Heinsoo, P.Kurpiers, A. Potocnik, A. Mezzacapo, U. Las Heras, L.Lamata, E. Solano, S. Filipp, and A. Wallraff, Digitalquantum simulation of spin models with circuit quantumelectrodynamics, Phys. Rev. X 5, 021027 (2015).

[25] C. Eichler, J. Mlynek, J. Butscher, P. Kurpiers, K. Ham-merer, T.?J. Osborne, and A. Wallraff, Exploring Inter-acting Quantum Many-Body Systems by ExperimentallyCreating Continuous Matrix Product States in Supercon-ducting Circuits, Phys. Rev. X 5, 041044 (2015).

[26] P. J. J. O’Malley, R. Babbush, I. D. Kivlichan, J.Romero, J. R. McClean, R. Barends, J. Kelly, P.Roushan, A. Tranter, N. Ding, B. Campbell, Y. Chen,Z. Chen, B. Chiaro, A. Dunsworth, A. G. Fowler, E. Jef-frey, A. Megrant, J. Y. Mutus, C. Neill, C. Quintana,D. Sank, A. Vainsencher, J. Wenner, T. C. White, P. V.Coveney, P. J. Love, H. Neven, A. Aspuru-Guzik, and J.M. Martinis, Scalable Quantum Simulation of MolecularEnergies Phys. Rev. X 6, 031007 (2016).

[27] M. D. Reed, L. DiCarlo, S. E. Nigg, L. Sun, L. Frunzio, S.M. Girvin and R. J. Schoelkopf, Realization of three-qubitquantum error correction with superconducting circuits,Nature 482, 382 (2012).

[28] L. Steffen, Y. Salathe, M. Oppliger, P. Kurpiers, M.Baur, C. Lang, C. Eichler, G. Puebla-Hellmann, A. Fe-dorov, and A. Wallraff, Deterministic quantum telepor-tation with feed-forward in a solid state system, Nature500, 319 (2013).

[29] R. Barends, J. Kelly, A. Megrant, A. Veitia, D. Sank, E.Jeffrey, T. C. White, J. Mutus, A. G. Fowler, B. Camp-

bell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, C.Neill, P. O’Malley, P. Roushan, A. Vainsencher, J. Wen-ner, A. N. Korotkov, A. N. Cleland, and J. M. Marti-nis, Superconducting quantum circuits at the surface codethreshold for fault tolerance Nature 508, 500 (2014)

[30] R. Barends, A. Shabani, L. Lamata, J. Kelly, A. Mez-zacapo, U. Las Heras, R. Babbush, A. G. Fowler, B.Campbell, Yu Chen, Z. Chen, B. Chiaro, A. Dunsworth,E. Jeffrey, E. Lucero, A. Megrant, J. Y. Mutus, M. Nee-ley, C. Neill, P. J. J. O’Malley, C. Quintana, P. Roushan,D. Sank, A. Vainsencher, J. Wenner, T. C. White, E.Solano, H. Neven, and John M. Martinis, Digitized adia-batic quantum computing with a superconducting circuit,Nature 534, 222 (2016)

[31] N. Ofek, A. Petrenko, R. Heeres, P. Reinhold, Z. Leghtas,B. Vlastakis, Y. Liu, L. Frunzio, S. Girvin, L. Jiang, M.Mirrahimi, M.H. Devoret, and R.J. Schoelkopf, Extend-ing the lifetime of a quantum bit with error correction insuperconducting circuits, Nature, advance online publica-tion (2016).

[32] M. Fitzpatrick, N. M. Sundaresan, A. C. Y. Li, J.Koch, and A. A. Houck, Observation of a dissipativephase transition in a one-dimensional circuit QED lat-tice, arXiv:1607.06895 (2016).

[33] S. Filipp, P. Maurer, P. J. Leek, M. Baur, R. Bianchetti,J. M. Fink, M. Goppl, L. Steffen, J. M. Gambetta, A.Blais, and A. Wallraff, Two-Qubit State Tomography Us-ing a Joint Dispersive Readout, Phys. Rev. Lett. 102,200402 (2009).

[34] L. DiCarlo, M. D. Reed, L. Sun, B. R. Johnson, J. M.Chow, J. M. Gambetta, L. Frunzio, S. M. Girvin, M. H.Devoret, and R. J. Schoelkopf, Preparation and measure-ment of three-qubit entanglement in a superconductingcircuit, Nature 467, 574 (2010)

[35] P. Roushan, C. Neill, A. Megrant, Y. Chen, R. Bab-bush, R. Barends, B. Campbell, Z. Chen, B. Chiaro, A.Dunsworth, A. Fowler, E. Jeffrey, J. Kelly, E. Lucero,J. Mutus, P. J.J. O’Malley, M. Neeley, C. Quintana,D. Sank, A. Vainsencher, J. Wenner, T. White, E.Kapit, H. Neven, J. Martinis, Chiral groundstate cur-rents of interacting photons in a synthetic magnetic field,arXiv:1606.00077 (2016).

[36] M. Abuwasib, P. Krantz and P. Delsing, Fabrication oflarge dimension aluminum air-bridges for superconduct-ing quantum circuits J. Vac. Sci. Technol. B 31, 031601(2013).

[37] Z. Chen, A. Megrant, J. Kelly, R. Barends, J. Bochmann,Yu Chen, B. Chiaro, A. Dunsworth, E. Jeffrey, J. Y.Mutus, P. J. J. O’Malley, C. Neill, P. Roushan, D. Sank,A. Vainsencher, J. Wenner, T. C. White, A. N. Clelandand John M. Martinis. Fabrication and characterizationof aluminum airbridges for superconducting microwavecircuits, Appl. Phys. Lett. 104, 052602 (2014).

[38] J. Jin, D. Rossini, R. Fazio, M. Leib, and M. J. Hart-mann, Photon solid phases in driven arrays of non-

20

linearly coupled cavities, Phys. Rev. Lett. 110, 163605(2013).

[39] L. Neumeier, M. Leib, and M. J. Hartmann, SinglePhoton Transistor in Circuit Quantum Electrodynamics,Phys. Rev. Lett. 111, 063601 (2013).

[40] H. Weimer, M. Muller, I. Lesanovsky, P. Zoller, and H.P. Buchler, A Rydberg quantum simulator, Nat. Phys. 6,382 (2010).

[41] D. Becker, T. Tanamoto, A. Hutter, F. L. Pedrocchi, andD. Loss, Dynamic generation of topologically protectedself-correcting quantum memory Phys. Rev. A 87, 042340(2013).

[42] T. Tanamoto, V. M. Stojanovic, C. Bruder, and D.Becker, Strategy for implementing stabilizer-based codeson solid-state qubits, Phys. Rev. A 87, 052305 (2013).

[43] A. Mezzacapo, L. Lamata, S. Filipp, and E. Solano,Many-Body Interactions with Tunable-Coupling Trans-mon Qubits Phys. Rev. Lett. 113, 050501 (2014).

[44] J. Koch, T.M. Yu, J. Gambetta, A.A. Houck, D.I. Schus-ter, J. Majer, A. Blais, M.H. Devoret, S.M. Girvin, andR.J. Schoelkopf, Charge-insensitive qubit design derivedfrom the Cooper pair box, Phys. Rev. A 76, 042319(2007).

[45] D. Kafri, Ch. Quintana, Y. Chen, A. Shabani, J. M.Martinis, and H. Neven Tunable inductive coupling ofsuperconducting qubits in the strongly nonlinear regime,arXiv:1606.08382 (2016).

[46] Y. Nakamura, Yu. A. Pashkin, and J. S. Tsai, Coherentcontrol of macroscopic quantum states in a single-Cooper-pair box, Nature 398, 786 (1999)

[47] A. Kitaev, Anyons in an exactly solved model and beyond,Ann. Phys. 321, 2 (2006)

[48] S. Bravyi, D. DiVincenzo, D. Loss, and B. Terhal, Sim-ulation of Many-Body Hamiltonians using PerturbationTheory with Bounded-Strength Interactions, Phys. Rev.Lett. 101, 070503 (2008).

[49] R. Konig, Simplifying quantum double Hamiltonians us-ing perturbative gadgets, Quant. Inf. Comp. 10, 292(2010).

[50] C. G. Brell, S. Flammia, S.Bartlett, and A. Doherty,Toric codes and quantum doubles from two-body Hamil-tonians, New J. Phys. 13, 053039 (2011).

[51] A. O. Niskanen, K. Harrabi, F. Yoshihara, Y. Nakamura,S. Lloyd, and J. S. Tsai, Quantum Coherent TunableCoupling of Superconducting Qubits, Science 316, 723(2007).

[52] N. A. Masluk, I. M. Pop, A. Kamal, Z. K. Minev, andM. H. Devoret Microwave Characterization of JosephsonJunction Arrays: Implementing a Low Loss Superinduc-tance, Phys. Rev. Lett. 109, 137002 (2012).

[53] M. T. Bell, I. A. Sadovskyy, L. B. Ioffe, A. Yu. Ki-taev, and M. E. Gershenson, Quantum Superinductorwith Tunable Nonlinearity, Phys. Rev. Lett. 109, 137003(2012).

[54] N. Samkharadze, A. Bruno, P. Scarlino, G. Zheng,D.P. DiVincenzo, L. DiCarlo, and L.M.K. VandersypenHigh-Kinetic-Inductance Superconducting Nanowire Res-onators for Circuit QED in a Magnetic Field, Phys. Rev.Applied 5, 044004 (2016).

[55] C. C. Bultink, M. A. Rol, T. E. O’Brien, X. Fu, B. C. S.Dikken, C. Dickel, R. F. L. Vermeulen, J. C. de Sterke, A.Bruno, R. N. Schouten, and L. DiCarlo, Active resonatorreset in the nonlinear dispersive regime of circuit QED,arXiv:1604.00916 (2016)

[56] Z. Chen, J. Kelly, C. Quintana, R. Barends, B. Camp-bell, Yu Chen, B. Chiaro, A. Dunsworth, A. G. Fowler,E. Lucero, E. Jeffrey, A. Megrant, J. Mutus, M. Nee-ley, C. Neill, P. J. J. O’Malley, P. Roushan, D. Sank, A.Vainsencher, J. Wenner, T. C. White, A. N. Korotkov,and J. M. Martinis, Measuring and Suppressing QuantumState Leakage in a Superconducting Qubit, Phys. Rev.Lett. 116, 020501 (2016).

[57] C. M. Quintana, A. Megrant, Z. Chen, A. Dunsworth, B.Chiaro, R. Barends, B. Campbell, Yu Chen, I.-C. Hoi, E.Jeffrey, J. Kelly, J. Y. Mutus, P. J. J. O’Malley, C. Neill,P. Roushan, D. Sank, A. Vainsencher, J. Wenner, T. C.White, A. N. Cleland, and J. M. Martinis, Characteriza-tion and reduction of microfabrication-induced decoher-ence in superconducting quantum circuits, Appl. Phys.Lett. 105, 062601 (2014)

[58] J. M. Chow, S. J. Srinivasan, E. Magesan, A. D. Corcoles,D. W. Abraham, J. M. Gambetta, and M. Steffen, Char-acterizing a four-qubit planar lattice for arbitrary errordetection, Proc. SPIE 9500, Quantum Information andComputation XIII, 95001G (2015)

[59] S. Trebst, P. Werner, M. Troyer, K. Shtengel, and C.Nayak, Breakdown of a Topological Phase: QuantumPhase Transition in a Loop Gas Model with Tension,Phys. Rev. Lett. 98, 070602 (2007).

[60] S. Dusuel, M. Kamfor, R. Orus, K. P. Schmidt, and J. Vi-dal, Robustness of a Perturbed Topological Phase, Phys.Rev. Lett. 106, 107203 (2011).

[61] T. Brecht, W. Pfaff, Ch. Wang, Y. Chu, L. Frunzio, M.H. Devoret and R. J. Schoelkopf, Multilayer microwaveintegrated quantum circuits for scalable quantum comput-ing npj Quantum Information 2, 16002 (2016).

[62] J.H. Bejanin, T.G. McConkey, J.R. Rinehart, C.T.Earnest, C.R.H. McRae, D. Shiri, J.D. Bateman, Y. Ro-hanizadegan, B. Penava, P. Breul, S. Royak, M. Zap-atka, A.G. Fowler, and M. Mariantoni, The QuantumSocket: Three-Dimensional Wiring for Extensible Quan-tum Computing, arXiv:1606.00063 (2016).

[63] V. Schmitt, X. Zhou, K. Juliusson, B. Royer, A. Blais,P. Bertet, D. Vion, and D. Esteve, Multiplexed readoutof transmon qubits with Josephson bifurcation amplifiers,Phys. Rev. A 90, 062333 (2014)

[64] Yu Chen, D. Sank, P. O’Malley, T. White, R. Barends, B.Chiaro, J. Kelly, E. Lucero, M. Mariantoni, A. Megrant,C. Neill, A. Vainsencher, J. Wenner, Y. Yin, A. N. Cle-land and John M. Martinis Multiplexed dispersive readoutof superconducting phase qubits Appl. Phys. Lett. 101,182601 (2012)

[65] M. Jerger, S. Poletto, P. Macha, U. Hubner, E. Il’ichevand A. V. Ustinov, Frequency division multiplexing read-out and simultaneous manipulation of an array of fluxqubits, Appl. Phys. Lett. 101, 042604 (2012)

[66] M. D. Reed, B. R. Johnson, A. A. Houck, L. DiCarlo, J.M. Chow, D. I. Schuster, L. Frunzio, and R. J. SchoelkopfFast reset and suppressing spontaneous emission of a su-perconducting qubit, Appl. Phys. Lett. 96, 203110 (2010).

[67] Evan Jeffrey, Daniel Sank, J. Y. Mutus, T. C. White,J. Kelly, R. Barends, Y. Chen, Z. Chen, B. Chiaro, A.Dunsworth, A. Megrant, P. J. J. O’Malley, C. Neill, P.Roushan, A. Vainsencher, J. Wenner, A. N. Cleland, andJohn M. Martinis, Fast Accurate State Measurement withSuperconducting Qubits, Phys. Rev. Lett. 112, 190504(2014).

21

[68] M. Pechal, J.-C. Besse, M. Mondal, M. Oppliger,S. Gasparinetti, A. Wallraff, Superconducting switchfor fast on-chip routing of quantum microwave fieldsarXiv:1606.01031 (2016).

[69] S. Asaad, C. Dickel, S. Poletto, A. Bruno, N. K. Lang-ford, M. A. Rol, D. Deurloo, L. DiCarlo, Independent, ex-tensible control of same-frequency superconducting qubitsby selective broadcasting, arXiv:1508.06676 (2015).

[70] A. Hamma and D. Lidar, Adiabatic preparation of topo-logical order, Phys. Rev. Lett. 100, 030502 (2008).

[71] M Hofheinz, H. Wang, M. Ansmann, Radoslaw C. Bial-czak, Erik Lucero, M. Neeley, A. D. O’Connell, D. Sank,J. Wenner, J. M. Martinis, and A. N. Cleland, Synthe-sizing arbitrary quantum states in a superconducting res-

onator, Nature 459, 546 (2009).[72] Y.-J. Han, R. Raussendorf, and L.-M. Duan, Scheme

for Demonstration of Fractional Statistics of Anyons inan Exactly Solvable Model Phys. Rev. Lett. 98, 150404(2007).

[73] N. Goldman and J. Dalibard, Periodically Driven Quan-tum Systems: Effective Hamiltoninas and EngineeredGauge Fields, Phys. Rev. X 4, 031027 (2014).

[74] I. Scholz, B. H. Meier, and M. Ernst, Operator-basedtriple-mode Floquet theory in solid-state NMR, J. Chem.Phys. 127, 204504 (2007).

[75] D. F. V. James, Quantum Computation with hot and coldions: An assessment of proposed schemes, Fortschritteder Physik 48, 823 (2000).