<<

Stoquasticity in circuit QED

A. Ciani QuTech, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, The Netherlands

B.M. Terhal QuTech, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, The Netherlands and JARA Institute for , Forschungszentrum Juelich, D-52425 Juelich, Germany (Dated: March 30, 2021) We analyze whether circuit QED Hamiltonians are stoquastic, focusing on systems of coupled flux . We show that scalable sign-problem-free path integral Monte Carlo simulations can typically be performed for such systems. Despite this, we corroborate the recent finding [1] that an effective, non-stoquastic Hamiltonian can emerge in a system of capacitively coupled flux qubits. We find that if the capacitive coupling is sufficiently small, this non-stoquasticity of the effective qubit Hamiltonian can be avoided if we perform a canonical transformation prior to projecting onto an effective qubit Hamiltonian. Our results shed light on the power of circuit QED Hamiltonians for the use of quantum adiabatic computation and the subtlety of finding a representation which cures the sign problem in these systems.

CONTENTS I. INTRODUCTION

I. Introduction 1 An important subject in quantum computational com- plexity theory is the study of the computational power of II. Circuit QED Hamiltonians 3 quantum Hamiltonians, in particular the hardness of es- timating its ground-state energy. Estimating the ground- A. Time-Reversal Invariance and Stoquasticity 3 state energy of quite general quantum Hamiltonians with 1. Qubit 4 sufficiently high accuracy is known to be a hard problem 2. Non-time reversal invariant Hamiltonians 5 for quantum computers (QMA-hard) [2]. 3. Gauge transformations 5 There is however an important subclass of so-called stoquastic Hamiltonians, first introduced in Ref. [3], for III. Stoquasticity of effective flux qubit which the ground-state problem is believed to be easier: Hamiltonians 5 rather than being QMA-complete it is StoqMA-complete [4, 5]. The class StoqMA is not very well understood, but A. Two coupled flux qubits with a it is known that NP ⊆ StoqMA ⊆ QMA, suggesting that non-stoquastic effective qubit Hamiltonian 6 ground-state energy problem for stoquastic Hamiltonians B. Flux qubits with weak-strength capacitive is at least as hard at the ground-state energy problem for and inductive couplings 7 classical Hamiltonians (NP-complete) but not as hard as C. Monte Carlo simulations 9 the problem for quantum Hamiltonians which do have a sign problem (QMA-complete). IV. Discussion 10 A quantum Hamiltonian H is stoquastic in a certain basis B = {|xi} if its entries are real and its off-diagonal V. Acknowledgements 11 elements are all non-positive, ı.e., hx|H|yi ≤ 0 for x 6= y [3]. If H is stoquastic, it can easily be shown that the

arXiv:2011.01109v3 [quant-ph] 29 Mar 2021 Gibbs matrix exp(−βH) is entrywise non-negative for all A. PIMC in the flux basis 11 β > 0 in the basis B and the partition function Z(β) = 1. Metropolis-Hastings reviewed 12 Tr exp(−βH) can be written as a sum of products of non- 2. Details of the Monte Carlo simulations 12 negative weights. In addition, the ground-state of H has non-negative amplitudes in the basis |xi. B. Stoquasticity of the Cooper-pair box 12 The term stoquastic was introduced to capture that these systems avoid the sign problem: the estimation of C. Flux qubit Hamiltonians 13 the partition function or the energy expectation value in the Gibbs state are amenable to stochastic Monte Carlo 1. The flux qubit reviewed 13 methods. Of particular interest is the use of stoquastic 2. Two-flux qubit Hamiltonian with symmetric Hamiltonians for quantum adiabatic computation. It has double well potentials 14 been shown that adiabatic computation using only sto- quastic frustration-free Hamiltonians can be efficiently References 16 classically simulated [6], but there are more general adia- 2 batic stoquastic computations whose output can only be tonian or on the effective qubit Hamiltonian. Previ- obtained using a sub-exponential, hence inefficient, num- ously, the effect of curing the sign problem by local ba- ber of classical queries [7],[8]. This shows that even when sis changes for qubit Hamiltonians was studied in [23– one avoids a sign problem, the power of stochastic Monte 27]. It has been an open question whether a stoquas- Carlo methods can be limited. On the other hand, such tic high-energy ‘master’ Hamiltonian can have a non- methods provide heuristic, often well-performing, classi- stoquastic effective low-energy Hamiltonian, –obtained cal simulation strategies. We refer the reader to [9] for using Schrieffer-Wollf perturbation theory–, even allow- a general review on adiabatic quantum computation in- ing for local basis changes in the basis of the low-energy cluding the use of stoquastic Hamiltonians. Hamiltonian [28]. Important physical realizations of quantum adiabatic The capacitively- and inductively coupled flux-qubit computation in the form of use induc- Hamiltonians thus seem to provide new examples of such tively coupled flux qubits [10]. These coupled flux qubits, stoquastic master Hamiltonians. Our aim in this paper described by the formalism of circuit QED, give rise to is not to prove this with full mathematical rigor however: the effective transverse field Ising model (TIM) of quan- we caution that even the perturbation theory for a simple tum annealing [11–13]. Since the TIM Hamiltonian is anharmonic oscillator can be subtle in its convergence stoquastic and hence amenable to quantum Monte Carlo [29]. methods, the power of the quantum annealing method is not well understood [9, 14]. Some research has been The contents of this paper are as follows. In Sec. II devoted to the use of additional non-stoquastic terms in we study the general form of circuit QED Hamiltonians the quantum annealing schedule [15–18], usually referred and their Lagrangians showing that they never give rise to as non-stoquastic catalysts. Again, it should be clear to a sign problem if we work in the continuous flux basis that the use of purely stoquastic Hamiltonians does not (details in Appendix A). We also discuss a generally non- preclude a quantum computational advantage, in partic- stoquastic Hamiltonian based on a nonreciprocal electric ular when the computation evolves through states other circuit for contrast and the role of gauge transformations. than the ground state via so-called diabatic quantum an- Sec. III deals with the stoquasticity of effective Hamilto- nealing [19]. nians for coupled flux qubits. The flux qubit itself is re- In this paper we study general Hamiltonians that viewed in Appendix C 1. In Sec. III A we show that two emerge in circuit QED [20–22] and their stoquasticity. capacitively coupled flux qubits can be described by an In circuit QED one starts with a Lagrangian, and then effective non-stoquastic Hamiltonian (we also show when one constructs a Hamiltonian, expressed in terms of elec- this does not happen in Appendix C 2). trical degrees of freedom, such as fluxes and charges which are by definition conjugate variables. Quanti- In Sec. III B we then show how in flux qubit sys- zation of such system results in a Hamiltonian which tems with weak capacitive and inductive coupling, we is the electric equivalent of a quantum mechanical sys- can always project onto an approximate effective qubit tem with conjugate variables of momentum and posi- transverse-field Ising Hamiltonian which is stoquastic. tion. Such continuous-variable Hamiltonian is then rep- Crucial to our derivation is the application of an efficient resented in its low-energy discrete sub-space, using per- canonical transformation before obtaining this effective turbative methods, leading to an effective Hamiltonian model, and the identification of a suitable qubit basis. which emulates a system. This procedure is inequivalent to applying local unitaries What we find in this paper is that very general flux to cure non-stoquasticity once an effective model is ob- qubit Hamiltonians, –and even transmon qubit Hamilto- tained and it explicitly exploits the structure of the ini- nians from a certain perspective–, can be called stoquas- tial Hamiltonian. This result is not in contradiction with tic: their thermal properties are directly simulatable us- that of Ref. [1], since in [1] the capacitive coupling is not ing classical Monte Carlo methods. Curiously, this does weak, and thus the derivation does not directly apply. not imply that a corresponding low-energy effective qubit In Sec. III C we apply the path integral Monte Carlo Hamiltonian is also stoquastic, even allowing for local method to estimate the thermal energy of two capaci- basis changes on the qubits. Such example of a non- tively coupled flux qubits. In particular, for weak ca- stoquastic qubit coupler for a pair of capacitively and pacitive coupling, we perform path integral Monte Carlo inductively-coupled flux qubits was first presented in [1]. (PIMC) simulations both in the original flux basis and us- However, we also show that for weak coupling, if ing the effective stoquastic Hamiltonians, showing good we apply a canonical transformation on the continuous- agreement with direct numerical diagonalization. For variable Hamiltonian before projecting down to a qubit strong capacitive coupling only PIMC in the flux basis space, the resulting qubit Hamiltonian is a transverse can be used without suffering from the sign problem. Us- field Ising model and thus no longer non-stoquastic. ing this method, the average thermal energy can still be These results show that apparent sign problems can accurately estimated, even if the system is described by a be cured by transformations and that the power of such low energy effective Hamiltonian which is non-stoquastic. transformations can depend at what level they are ap- We finally provide some discussion and perspective in plied, ı.e., on the global (continuous-variable) Hamil- Sec. IV. 3

II. CIRCUIT QED HAMILTONIANS inductive energies of linear self- and mutual inductances, and by the contributions of the (nonlinear) Josephson In this section we consider typical Hamiltonians in cir- junctions. We refer the reader to Refs. [20], [35] for a cuit QED [20] and discuss their stoquasticity. We focus detailed description of how to obtain the Hamiltonian of on Hamiltonians without time-dependent external driv- a superconducting circuit and how this Hamiltonian can ing fields as we are interested in thermal and ground- be formally written as in Eq. (2.1). −1 state properties for quantum adiabatic computing. For If the inverse capacitance matrix C is a diagonal Hamiltonians subject to time-dependent driving, one can matrix, Hamiltonians of the form of Eq. (2.1) are clearly easily break-time reversal invariance, leading to gener- stoquastic in the flux basis when we consider the dis- ally complex Hamiltonians [30] which are thus not sto- cretized version of the Hamiltonian operator, i.e. we dis- quastic. In addition, electric circuits which are not in- cretize the flux basis. This can be seen from the fact that ˆ cluded in the discussion here and which may lead to the term U(Φ) is diagonal in this basis, while the kinetic ˆ2 ∂2 non-stoquastic Hamiltonians are ones where phase-slip term gives rise to terms Qk = − 2 . Discretizing the flux ∂Φk junctions are present [31, 32]. basis so that Φk = δm with integer m in some interval, We first consider classical Hamiltonians with N inde- the finite-difference second derivative is approximated as pendent degrees of freedom of the following form d2f 1 − dx2 | ≈ δ2 (−f(x + δ) − f(x − δ) + 2f(x)). Hence, as a matrix, this finite-difference negative Laplace operator is 1 H = K(Q) + U(Φ) = QT C−1Q + U(Φ), (2.1) real and has non-positive off-diagonal entries. This fact 2 is explicitly used in Ref. [36] to construct a path integral T Monte Carlo method. When C is not diagonal, the dis- where Φ = Φ ,..., Φ  is the vector of independent 1 N cretization becomes more awkward, but, as we show in  T fluxes, Q = Q1,...,QN is the vector of conjugate the next Subsection, we still obtain a sign-problem-free charges. This Hamiltonian originates from a Lagrangian representation of the partition function. of the form 1 L(Φ˙ , Φ) = Φ˙ T CΦ˙ − U(Φ). (2.2) A. Time-Reversal Invariance and Stoquasticity 2 The Hamiltonian is obtained via a Legendre transform Rather than using a discretization of of the flux degrees H = QT Φ˙ − L, with the vector of charges defined as of freedom we determine a non-negative path integral Q = ∂L . By definition, through a Poisson bracket, the ∂Φ˙ expression for exp(−βH) and Z = Tr exp(−βH) as an classical charges and fluxes are conjugate variables. The integral over non-negative weights for the Hamiltonians definition of Q for this Lagrangian gives in Eq. (2.1). This representation can be directly used to perform dΦ Q = C . (2.3) path integral simulations of a quantum adiabatic com- dt putation using Hamiltonians of this form. We use this The capacitance matrix C is a N ×N symmetric positive- representation in Sec. III A to simulate the thermal state definite matrix with diagonal entries and negative off- of a system of two capacitively coupled flux qubits. diagonal entries and it is thus invertible [33]. This implies Indeed, for the quantum Hamiltonian in Eq. (2.1) we that C−1 is a symmetric, entrywise non-negative matrix can write exp(−βH) in the flux basis using a Feynman [34]. path integral. It can be obtained in discretized form, When we quantize this electric system, we promote the using Trotterization, see Appendix A, leading to conjugate variables to quantum operators and they obey Z the canonical commutation relations by definition: Z = dΦ hΦ| e−βH |Φi ≈ ˆ ˆ Z [Φk, Ql] = i~δkl. (2.4) −βHc(Φ1,...,Φm) C dΦ1 . . . dΦM e (2.5)

We note that these conjugate operators Qˆk and Φˆ l, like momenta and positions in a quantum-mechanical system, with non-negative constant C, periodic boundary condi- tions Φ = Φ and classical Hamiltonian take values in R (see Subsection II A 1 and Appendix B M+1 1 for a discussion on the common switch to 2π-periodic M phase variables). κ X T T Hc = (Φ − Φ )C(Φs+1 − Φs)+ The term 1 QT C−1Q in Eq. (2.1) represents the elec- 2 s+1 s 2 s=1 trostatic energy stored in the capacitors of the system M 1 X and, in a mechanical analogy, it has the interpretation U(Φ ), (2.6) of a kinetic term. The term U(Φ) represents the induc- M s s=1 tive, ‘potential’, contribution to the energy. We assume M no particular form of U(Φ) as our discussion will be gen- with coupling coefficient κ = 2 2 assuming large Trotter ~ β eral. In circuit QED U(Φ) will be given by the sum of the parameter M  1. This expresses the well-known map- 4 ping from the partition function of a quantum Hamilto- It is clear that the Lagrangian Lshift is not time- nian with its N-dimensional phase space onto the parti- reversal invariant due the presence of the charge vector ∂ tion function of a N+1-dimensional classical Hamiltonian Qg. When we use Q = −i~ ∂Φ , the Hamiltonian is com- [37]. Taking a continuum limit we can introduce the vari- plex. able τ, taking values τ = sβ , with s = 1,...,M and for It is also apparent that a canonical transformation M 0 0 large enough number of Trotter slices M, the integrand Q = Q − Qg, Φ = Φ can bring this Hamiltonian to on the r.h.s in Eq. (2.5) equals the form in Eq. (2.1). At a quantum level, this transfor- mation preserves the commutation relations between Qˆ ! Z β  1 ∂ΦT ∂Φ  and Φˆ and corresponds to the basis change exp − dτ 2 C + U(Φ) . (2.7) 0 2~ ∂τ ∂τ T |Φ0i = eiΦ Qg /~ |Φi . (2.11)

In this continuum limit we also have T 0 i Q /~ 0 0 Z Note that then e |Φ i = |Φ + i for some vector , as is expected. Since the basis change merely applies hΦ1| exp(−βH) |Φ0i = C DΦ × Φ0→Φ1 overall phases, one can verify, following the analysis in ! Z β  i dΦ  Appendix A for Hshift, that hΦ| exp(−βHshift) |Φi, ı.e., exp dτL , Φ) . (2.8) using the original basis, still has a Monte-Carlo path inte- 0 ~ dτ gral representation with non-negative weights as we start and begin at the same state |Φi. ˙ with Lagrangian L(Φ, Φ). Clearly, if the integrand L on However, if we use the original basis |Φi then the r.h.s of Eq. (2.8) is real-valued for all Φ, then the hΦ1| exp(−βHshift) |Φ0i for Φ1 6= Φ0 is complex, and r.h.s. is a path integral over non-negative weights and hence Hshift cannot be called stoquastic in this basis. suffers no sign problem. For the time-reversal invariant The rather trivial basis change to |Φ0i in Eq. (2.11) cures Lagrangian of Eq. (2.2) the integrand is real-valued (as this, but since the path integral expression for Z uses the can also be seen from the finite Trotter parameter expres- same initial and final state one could also omit it. In any sions in Eq. (2.5)-(2.6)). For more general Lagrangians case, it follows that for the Hamiltonian in Eq. (2.9) one of the form L = K(Φ˙ ) − U(Φ), –assuming that they lead can apply the path integral Monte Carlo method without to a well-defined Hamiltonian–, the path integral seems sign problem to study the thermal expectation value of less useful as the integral over momenta (as in Appendix H and any diagonal operator in Φ. A) cannot necessarily be executed [38]. In such cases one ∂ could discretize the flux basis and express Qˆk = −i ~ ∂Φk as a finite-difference operator. In case H can be ex- 1. Transmon Qubit panded as a Taylor series in Qˆ, the Hamiltonian will be ˆ2n real when it only contains terms Q with n ∈ N, ı.e., A Cooper-pair box or transmon qubit coupled to an ex- only containing terms which are invariant under time- ternal voltage source, inducing an offset charge Q , pro- ˆ ˆ ˆ ˆ g reversal of operators Qk → −Qk, Φk → Φk. However, vides a simple example of the Hamiltonian in Eq. (2.9) βK(Q) this does not seem sufficient to let hΦ1| I − M |Φ0i [39]. In that case we have a single flux Φ and its conjugate be non-negative as the finite-difference expression of, say, charge Q. By the basis change in the previous paragraph 4 a fourth-derivative Qk has alternating signs on the off- the transmon qubit Hamiltonian is thus stoquastic and diagonal (An example is the Lagrangian of a relativistic, its thermal state a non-negative matrix (assuming dis- but non-causal, particle, expressed in circuit QED coor- cretization). However, the transmon qubit Hamiltonian q 2 Φ˙ 2 is often stated in a rotor subspace of the oscillator space dinates as L = −mc 1 − 2 − U(Φ) with Hamiltonian c which is spanned by a compact 2π-periodic supercon- p 2 2 2 H = c m c + Q + U(Φ)). ducting phase basis. This rotor subspace is fixed by the . operator SQ = exp(iπQ/eˆ ) taking a certain phase eigen- In circuit QED one also encounters Hamiltonians such value, see a detailed analysis in Appendix B. Physical as processes which affect the support of the 1 in these rotor subspaces are the tunneling of single or H = (Q − Q )T C−1(Q − Q ) + U(Φ). (2.9) shift 2 g g fractional charges through the Josephson junc- tion: these are energetically suppressed due to supercon- where Qg is a vector of classical (gate) charges. This ductivity. Even though changes in the support are ener- Hamiltonian originates from a Lagrangian of the form getically suppressed, an initial state of a transmon qubit device could well be one with support in multiple rotor 1 T T subspaces. The upshot of these considerations is this. Lshift = Φ˙ CΦ˙ + Q Φ˙ − U(Φ) (2.10) 2 g Whether the transmon qubit can be called stoquastic or not depends on whether one considers the Hamiltonian in using the definition Q = ∂Lshift and the Legendre trans- ∂Φ˙ a rotor subspace or the full oscillator space and whether T form Hshift = Q Φ˙ − Lshift. one is physically interested in the thermal state in the 5 full oscillator space or the thermal state in a single rotor Thus the gauge freedom expressed in f can lead to a subspace. In all but one rotor subspace the Hamilto- Hamiltonian which seems (at first sight) non-stoquastic. nian is not stoquastic with respect to the phase basis in ˙ C ˙ 2 An example is the case k = 1 with L(Φ, Φ) = 2 Φ −U(Φ) this subspace and the ground-state wave function is not and let’s take, say, f(Φ) = Φ3. We get H0(Q0, Φ) = a non-negative function of phase. 1 0 2 2 2C (Q −3Φ ) +U(Φ) which is not manifestly stoquastic. The flux-type qubits used in quantum annealing [10, Nonetheless, this still does not lead to a sign problem as 40], ı.e., the focus of this paper, include self- and mutual Eq. (2.8) is still satisfied with the new Lagrangian. inductances which makes a switch to a rotor subspace One can view gauge transformations as potential cur- not correct as the dynamics induced by the Hamiltonian ing transformations. In fact we can observe that for a is not confined to such subspace. Hamiltonian of the form 1 H = (Q − A(Φ))T C−1(Q − A(Φ)) + U(Φ). (2.14) 2. Non-time reversal invariant Hamiltonians 2 where A(Φ) is a N-dimensional vector field depending on As another class of examples, we consider the La- ∂f Φ, we can gauge away this field when A = ∇f(Φ) = ∂Φ grangian of a so-called non-reciprocal electric circuit for some f(Φ). which involve gyrators or circulators [41–43]. Such el- Considering the non-time reversal invariant La- ements can be obtained through active driving [44] or grangian of the previous section, taking f(Φ) = ΦT MΦ coupling to a magnetic field [45]. The Lagrangian is then would indeed lead to ∂f = MΦ, but the anti-symmetry of general form ∂Φ of the matrix M immediately implies that f(Φ) = 1 ΦT MΦ = 0. Said differently, we cannot gauge away L = Φ˙ T CΦ˙ + Φ˙ T MΦ − U(Φ), (2.12) gyr 2 these time-reversal symmetry breaking terms, similar as one cannot gauge away the vector potential A~ = ∇×~ B~ in with real, anti-symmetric matrix M [41]. Applying a a minimal coupling Hamiltonian of a particle in a mag- Legendre transformation, one obtains the Hamiltonian netic field. The effect of gauge transformations on the form of ef- 1 T −1 Hgyr = (Q − MΦ) C (Q − MΦ) + U(Φ). (2.13) fective Rabi model Hamiltonians has been recently dis- 2 cussed for instance in [46, 47] and references therein. Our analysis of how canonical transformations on the full cir- It is clear that the Lagrangian Lgyr is not time-reversal invariant due to the term Φ˙ T MΦ, making the Hamil- cuit QED Hamiltonian can affect stoquasticity of the ef- tonian complex and hence not stoquastic. Following the fective qubit Hamiltonian bears some resemblance to this discussion, although the focus is different. path integral analysis in Appendix A for exp(−βHgyr), one finds complex expressions due to the presence of QT Φ terms. III. STOQUASTICITY OF EFFECTIVE FLUX In this general case the application of canonical (sym- QUBIT HAMILTONIANS plectic) transformations, –possibly mixing ‘positions and momenta’ but preserving their commutations relations–, In this section we prefer to work with dimensionless cannot even bring Hgyr to a real, time-reversal invari- variables and we thus introduce dimensionless charges ant form, that is, a form in which it is invariant un- Q 2πΦ h 0 0 q = , and fluxes φ = , with Φ0 = the super- der Q → −Q . Thus the sign problem for such non- 2e Φ0 2e reciprocal Hamiltonians can generically not be cured by conducting flux quantum. Then, for k, l = {1,...,N} we a canonical transformation. then have ˆ ˆ ˆ [φk, φl] = [ˆqk, qˆl] = 0, [φk, qˆl] = iδkl, (3.1) 3. Gauge transformations The quantum Hamiltonian of Eq. (2.1) in terms of the rescaled operators equals Here we also like to comment on the well-known fact T that the Lagrangian L does not determine the (quan- H = 4qˆ EC qˆ+ U(φˆ). (3.2) tum) dynamics uniquely, i.e. one can always add a to- tal time derivative of an arbitrary function to the La- where we defined the charging energy matrix ˙ ˙ df(Φ,t) grangian: L(Φ, Φ, t) → L(Φ, Φ, t) + dt . Namely, assuming (for simplicity) that f (nor L) has no explicit e2 df(Φ) PN ∂f −1 time-dependence, we have = Φ˙ k. This EC = C . (3.3) dt k=1 ∂Φk 2 implies that the conjugate variables Qk get changed to 0 ∂f Q = Q + ∂Φ and the Hamiltonian is invariant under In the previous section we have shown that a general the transformation: H(Q, Φ) = H0(Q0, Φ) where H0 is a circuit QED Hamiltonian of the form in Eq. (3.2) is sto- different function. quastic and free of the sign problem and can be simulated 6

as

2 X 1 H = 4E qˆ2 + E φˆ2 − Eeff cosφˆ + φx  Ck k 2 Lk k Jk k qk k=1 ˆ ˆ + 8EC12qˆ1qˆ2 + EL12φ1φ2, (3.4)

where ECk and EC12 are the diagonal and off-diagonal FIG. 1: Electric circuit of two capacitively and entries of the charging matrix, respectively. The charging inductively coupled flux qubits. energy matrix is in turn directly related via Eq. (3.3) to the capacitance matrix

C + C −C  C = 1 c c . (3.5) by the PIMC algorithm. We note that the Hamiltonian in −C C + C Eq. (3.2) also models various other ‘modern’ flux qubits c 2 c such as the fluxonium [48]. In addition, the ELk and EL12 are the diagonal and off- For superconducting circuits one usually wants to rep- diagonal entries of the inductive energy matrix, respec- resent the problem using an effective qubit Hamiltonian eff x  tively, while EJk φcjj is the effective Josephson energy that describes the behaviour of a discrete number of low- defined in Eq. (C.2). We assume that EC12 and EL12 are lying energy levels. It is thus natural to ask the ques- much smaller than the gap of the qubit subspace and any tion of whether these effective Hamiltonians on qubits other energy level, so that we can treat the coupling at are stoquastic or not. Weak inductive coupling in flux lowest-order perturbation theory. For concreteness, con- qubits gives rise to the Hamiltonian of an effective TIM, sidering a flux qubit with parameters as in Table I in the ı.e., with X, Z and ZZ interactions (see Appendix C 1), symmetric configuration this gap is 5.9 GHz. which is stoquastic (by applying Pauli Z basis changes We can immediately note that if we project the Hamil- so the X terms are negative). tonian in Eq. (3.4) onto a tensor product of qubit spaces, It was shown in [1] that by adding a capacitive cou- each qubit space associated with an uncoupled Hamil- pling between flux qubits, the effective two-qubit Hamil- tonian with conjugate variables qk, φk, the presence of tonian is non-stoquastic and the non-stoquasticity can- both inductive and capacitive couplings will typically not be cured by local unitaries, according to the crite- lead to Pauli interactions of rank equal to 2. Said dif- ria of Ref. [24]. This finding was further confirmed in ferently, the projected two-qubit interaction term of the Ref. [49], where the authors put forward a more refined P3,3 form H = i=1,j=1 βijPi ⊗ Pj is such that the 3 × 3 analysis based on the perturbative Schrieffer-Wolff (SW) matrix β has rank 2. transformation [50]. But when either one of the couplings EC12 or EL12 is While a higher-order SW transformation is usually nec- zero, we note that the rank of matrix β is 1. It can be essary in order to achieve good accuracy of all the param- proved quite directly that a two-qubit Hamiltonian with eters in the problem, for the purpose of studying stoquas- a rank 1 β-matrix and arbitrary single-qubit terms can ticity, and build intuition, we start by considering effec- be locally sign-cured [24]. tive qubit Hamiltonians that are obtained by simply pro- Let us look at this in detail here. We can project onto jecting the initial Hamiltonian onto the computational the qubit subspace for each flux qubit as in Appendix C 1 subspace. This is the SW transformation at lowest-order we obtain and it is the common way to obtain an effective qubit model in systems of flux qubits [51]. ∆ ε ∆ ε H /h = − 1 X − 1 Z − 2 X − 2 Z We review the basics of flux qubits in Appendix C 1. 2q 2 1 2 1 2 2 2 2 In the next section, we explain why and under which + JYY Y1Y2 + JZZ Z1Z2, (3.6) conditions, a system of two coupled flux qubits can give rise to an effective non-stoquastic Hamiltonian. where E J = 8 C12 h0|qˆ |1i h0|qˆ |1i , (3.7a) YY h 1 1 2 2 A. Two coupled flux qubits with a non-stoquastic E effective qubit Hamiltonian J = L12 h0|φˆ |0i h0|φˆ |0i . (3.7b) ZZ h 1 1 2 2

We consider a system of coupled flux qubits shown By definition, the tunnel couplings ∆1,2 are positive, see in Fig. 1. We show why the reduced Hamiltonian ob- Eq. (C.4), while the local fields ε1,2, see Eq. (C.9), are tained using the standard flux qubit basis is, in a certain real. parameter regime, non-stoquastic even if we allow for If we do not have capacitive coupling JYY = 0, the single-qubit unitary transformations. Hamiltonian in Eq. (3.6) is that of a TIM, which can al- The Hamiltonian of the circuit in Fig. 1 can be written ways be made stoquastic via single-qubit unitaries. If the 7 inductive coupling is absent, then JZZ = 0. The Hamil- mation for a single pair of qubits does not necessarily im- tonian is clearly non-stoquastic in the chosen basis as the ply the existence of a curing transformation which works Y ⊗ Y matrix has alternating signs on the off-diagonal for the entire set of qubits as the local basis changes have elements. However, as said above, we can always make it to be chosen to work for each two-qubit interaction. stoquastic via single-qubit unitaries in the following way. In the previous discussion, we have shown that the ef- In this case, we first perform a rotation around the Y - fective qubit Hamiltonian of two coupled flux qubits can axis, leaving the term YY unchanged, on each qubit that be non-stoquastic even after single-qubit unitary rota- transforms −∆kXk/2 − εkZk/2 7→ −∆˜ kXk/2, k = 1, 2. tions. However, there could be other ways to cure non- Now we can easily make the Hamiltonian stoquastic by stoquasticity. We show in the next Sec. III B that if performing the transformation Y1,2 ↔ Z1,2. the capacitive and inductive couplings are weak enough, More generally, we show in Appendix C 2 that if the effective flux qubit Hamiltonians can always be made ap- local fields ε1,2 are zero, even in the case in which we proximately stoquastic if we perform a canonical trans- have both capacitive and inductive couplings, the two- formation before obtaining the reduced qubit Hamilto- qubit Hamiltonian can always be made stoquastic and nian. While these transformations are highly non-local, this in fact holds at arbitrary order in SW perturbation they can still be implemented efficiently before reducing theory. to a qubit model. In particular, we show that the ad- If the local fields ε1,2 are non-zero, and we have capac- dition of capacitive couplings to flux qubit Hamiltonians itive and inductive coupling such that |JYY | > |JZZ | > 0 yields, to lowest perturbative approximation, a TIM with we conclude that the Hamiltonian cannot be made sto- modified parameters. While this derivation relies on the quastic by a product of two single-qubit unitaries, follow- fact that the coupling is weak, it has the appealing fea- ing the reasoning in Ref. [24], The basic (rough) idea is ture that it is valid for an arbitrary number of qubits. that in order for H to have non-positive off-diagonal ele- ments a term like Y1Y2 should be accompanied by a term B. Flux qubits with weak-strength capacitive and X1X2 of equal magnitude (which it is not) or be rotated away to the XZ-plane. In the latter case, one however inductive couplings also rotates the single-qubit X and Z terms into having a Y component, making the Hamiltonian complex and We consider flux-qubit systems where the Hamiltonian non-stoquastic. takes the following particular form We remark that, for simplicity of exposition, we T 1 ˆT ˆ present the discussion assuming the validity of the pro- H = 4qˆ EC qˆ + φ ELφ jection, in order to highlight the mechanism that leads to 2 N a non-stoquastic behaviour. By refining the perturbation X − Eeff cosφˆ + φx , (3.8) theory, ı.e., using higher order SW transformation for in- Jk k qk stance [49], the Hamiltonian can still be non-stoquastic k=1 eff x even in the absence of an inductive coupling. In partic- with EJk the effective Josephson energy and φqk the ex- ular, it is shown in Ref. [1] that in the case of strong ternal flux threading the loop formed by the SQUID loop capacitive coupling, the higher levels of the flux qubits and the corresponding shunting inductance of the kth generate an additional X1X2 term which can make the flux qubit. The inductive energy matrix equals Hamiltonian non-stoquastic under single-qubit unitaries. 2 Φ0 −1 Naturally, for a two-qubit Hamiltonian we can apply a EL = L , (3.9) two-qubit unitary basis change to diagonalize the Hamil- 4π2 tonian, hence there is always a basis change which re- with L−1 the inverse of the inductance matrix L (which moves the sign problem. However, one can readily extend can be assumed to be positive-definite). this two-qubit case to a line of N coupled flux qubits. By taking parameters such that all degrees of freedom When the capacitive and inductive couplings are suffi- are in the flux qubit regime, the Hamiltonian in Eq. (3.8) ciently weak (so that EC couples only nearest-neighbor models a system of capacitively and inductively coupled qubits on the line), we obtain a N flux qubit Hamiltonian flux qubits, where the inductive coupling is expressed in with YiYi+1,ZiZi+1 coupling between nearest-neighbor EL and the capacitive coupling is expressed in EC . See qubits i and i + 1 on the line. The same arguments then Fig. 1 for two such coupled qubits. apply as in the two-qubit case: when ∆i 6= 0, i 6= 0 We now show how this Hamiltonian, which has no sign the N-qubit Hamiltonian cannot be made stoquastic by problem in the flux basis as we discussed in Sec. II A, can a product of N single-qubit basis changes. be reduced to an effective qubit Hamiltonian which is also This is thus in sharp contrast with the fact that the stoquastic if the capacitive couplings are small, ı.e., the general master Hamiltonian in Eq. (3.2) can be called sto- off-diagonal elements of EC are much smaller than the di- quastic in the flux basis and was amenable to the PIMC agonal ones. In addition, the mutual-inductive couplings method, see the numerics in Sec. III C. between the flux qubits should also be small so that a Note also that when coupled flux qubits are non- projection onto the eigenbasis of the uncoupled qubits is identical in their parameters, finding a curing transfor- a good approximation. 8

In some sense this is not a surprising result as our sym- • the effective Hamiltonian of the kth flux qubit Hk plectic transformation removes the capacitive couplings, as leaving only the inductive couplings which lead, when 0 projected, to rank-1 β-matrices. (E )kk H = 4E qˆ2 + L φˆ2 We introduce the following canonical transformation k C0 k 2 k eff  ˆ x  − EJk cos Skkφk + φqk ; (3.16) qˆ0 = Sqˆ, φˆ0 = S−1φˆ, (3.10)

ind where we defined the matrix • the inductive coupling Hamiltonian Hc as X  1/2 Hind = (E0 ) φˆ φˆ ; (3.17) T EC c L kl k l S = S = , (3.11) hk,li EC0 with EC0 an arbitrary charging energy which just en- • the additional coupling due to the Josephson junc- jj sures the entries in S are dimensionless. S preserves tions Hc as the canonical commutation relations as S = ST , i,e. [φˆ0 , qˆ0] = [φˆ , qˆ ] = iδ . We will drop the primes from N N k l k k kl jj X eff  ˆ x X ˆ now on for these canonical variables. The Hamiltonian Hc = EJk sin Skkφk + φqk Sklφl. (3.18) in Eq. (3.8) becomes k=1 l=1 l6=k

T 1 T 0 H = 4E qˆ qˆ + φˆ E φˆ− We can now use the flux qubit Hamiltonians Hk to de- C0 2 L fine a local computational basis Bk = {|0i , |1i }, similar N  N  k k X eff ˆ x X ˆ to what is done in Appendix C 1. The global computa- EJk cos Skkφk + φqk + Sklφl , (3.12) tional basis B is obtained by taking all possible tensor k=1 l=1 NN l6=k products of these states, ı.e., B = k=1 Bk. By project- ing onto this basis we obtain a reduced N-qubit Hamil- where we have introduced an effective inductive energy tonian that can be written as matrix  N  X ∆k εk X 0 T Heff /h = − Xk + Zk + JklZkZl. (3.19) EL = S ELS. (3.13) 2 2 k=1 hk,li We now first show that we can map the Hamiltonian The parameters in this Hamiltonian are obtained as fol- in Eq. (3.12) to a transverse field Ising model when lows. With the definition of the flux qubit Hamiltonian the capacitive couplings between the flux qubits are not H defined in Eq. (3.16) the tunnel couplings ∆ are too large. This implies that matrix E ∝ C−1 has off- k k C given by diagonal elements which are small compared to its di- agonal elements and hence so will S when we treat the (k) (k) Ee − Eg capacitive coupling between flux qubits as a perturba- ∆k = , (3.20) tion. h In this case, we can expand each cosine term as (k) with Eg,e are the ground and first-excited eigenenergies x of Hk in the double well configuration φqk = π. By defin-  N  x x eff ˆ x X ˆ ing δqk = φqk −π the parameters εk and considering small − EJk cos Skkφk + φqk + Sklφl x δqk, similarly to Appendix C 1, we define l=1 l6=k eff  ˆ x  Eeff ≈ −EJk cos Skkφk + φqk + J x  ˆ  ˆ  εk = δqk h0| sin Skkφk |0ik − h1| sin φk|1ik = N h eff  ˆ x X ˆ Eeff EJk sin Skkφk + φqk Sklφl. (3.14) J x ˆ  2 δqk h0| sin Skkφk |0ik . (3.21) l=1 h l6=k Finally, neglecting the small corrections we get when x This allows us to rewrite Eq. (3.12) as δqk 6= 0, the exchange coupling Jkl reads

N E0 X ind jj Lkl ˆ ˆ H ≈ H + H + H , (3.15) Jkl = h0|φk|0i h0|φl|0i − k c c h k l k=1 Eeff J S h0| sinS φˆ |0i h0|φˆ |0i . (3.22) where we defined h kl kk k k l l 9

Eq. (3.19) is the Hamiltonian of a TIM, which is sto- quastic in the computational basis. Notice that this derivation is valid for an arbitrary number of flux qubits. The transverse field Ising model in Eq. (3.19) can also be mapped to a classical system and be studied using the PIMC method. We refer the reader to Ref. [52] for a derivation. The N + 1-dimensional classical Hamiltonian associated with the TIM reads

M  N N X X (s) (s+1) X εk (s) H /h = − J ⊥σ σ + σ − eff,c k k k 2 k s=1 k=1 k=1  X (s) (s) Jklσk σl , (3.23) hk,li

(s) where the variables σk are classical spins which can take FIG. 2: Average thermal energy using the PIMC in the value ±1 and we defined the parameter flux basis and the effective TIM model obtained in Sec. III B as a function of the number of Trotter steps M. −1 ⊥ M ∆kβ The temperature is taken to be (hβ) = 0.93 GHz. Jk = − ln tanh , (3.24) 2β 2M The coupling capacitance is chosen to be Cc = 10 fF, so that we obtain EC12/h = 0.008 GHz  EC /h. Both flux with M the number of Trotter slices as in Sec. II. qubits are operated in the symmetric double well x configuration with φq1,2 = π. The black dashed line corresponds to the exact thermal energy obtained from C. Monte Carlo simulations numerical diagonalization, while the ground-state energy is Eg = 7.675 GHz. In this subsection we perform sign-problem-free Monte Carlo simulations for the average thermal energy of a sys- tem of two capacitively coupled flux qubits. We begin by studying the problem using the path integral repre- sentation of the original Hamiltonian in the flux basis as discussed in Sec. II A for the case of weak capacitive cou- pling. We compare the result with the PIMC with those using the effective TIM discussed in Sec. III B. The goal is to compare the results using these two methods for weak capacitive coupling versus the exact results for the estimation of the average thermal energy

1   hHi = Tr He−βH , (3.25) β Z with Z = Tr[exp(−βH)] the partition function. We also study whether in the case of strong capacitive cou- pling, for which the effective qubit Hamiltonian is non- stoquastic as in Ref. [1], the PIMC using the origi- nal Hamiltonian provides reliable results. We consider two identical flux qubits with parameters as in Table I. FIG. 3: Average thermal energy using the PIMC in the Throughout this subsection the minimum of the poten- flux basis for the case of strong capacitive coupling. tial is taken as the zero of the energy in any parameter The coupling capacitance is Cc = 104 fF giving set. EC12/h = 0.062 GHz. The flux qubits are operated in x −4 We start by considering the case of a symmetric poten- an asymmetric configuration with δq1/2π = 10 , x −4 tial for both flux qubits and small coupling capacitance. δq2/2π = 2 × 10 . According to the color coding, the We provide some details of the Monte Carlo simulations dashed lines denote the exact thermal energy at the in Appendix A 2. The results are shown in Fig. 2. We see respective temperature. The ground-state energy is that the PIMC in the flux basis needs more Trotter slices Eg/h = 9.016 GHz, which is within 0.2% accuracy the to accurately estimate the average thermal energy com- thermal energy in the orange (light gray) dashed line. pared to the TIM. Qualitatively this happens because in the TIM we are using a basis in which the Hamiltonian 10

Parameter GHz TIM Hamiltonian in III as we do not discuss the error in- EC /h 0.124 duced by only using the lowest-order Schrieffer-Wolff pro- EJ /h 1600 eff jection. For example, if the non-stoquasticity of Eq. (3.6) EJ /h 760 is of the same order of magnitude as the error induced E /h 704 L by perturbation theory, then one cannot draw any hard TABLE I: Parameters for simulations. The effective conclusions. Josephson energy is obtained by setting the external x fluxes to φcjj = 0.685550 × π. This choice of parameters We have observed that circuit QED Hamiltonians eff corresponds to EJ /EL = 1.08 and the tunnel coupling are generically stoquastic and thus amenable to Monte ∆/h = (Ee − Eg)/h = 1.36 GHz Carlo methods in their continuous variable representa- tion if they don’t contain explicit time-reversal invariance breaking terms due to driving or non-reciprocity in the is approximately diagonal and so the quantum effects are electric circuit. We have also recently become aware of already taken into account. Notice also that the simu- Ref. [36], where similar conclusions are drawn, although lations in the flux basis consistently underestimate the using a different Monte Carlo method that requires the average thermal energy for small M. This is due to the discretization of the flux degrees of freedom. In our case, fact that the Trotter break-up formula neglects the com- instead, we do not require this flux discretization, but we mutation relation, ı.e., quantum mechanical effects, and rely on a finite number of Trotter slices. so we expect to have lower zero-point energy compared to the exact quantum solution for small M. However, Naturally, these arguments do not directly apply to we see that as expected with 50 Trotter slices, PIMC in fermionic systems or fermionic field theories, in which the flux basis accurately estimates the average thermal the path integral is an integral over (non-commuting) energy. Grassmann variables. Alternatively, for fermions treated The PIMC method in the flux basis can also be used in first quantization, we can view the sign problem as to study the case of strong capacitive coupling as consid- arising from the fact that we are restricting the space of ered in Ref. [1], without fundamental limitations. This states to wave-functions which are fully anti-symmetric is shown in Fig. 3. For the given parameters, the ef- under the interchange of particles: a Gibbs state or a fective qubit Hamiltonian is non-stoquastic as discussed ground state in the full phase space is not the relevant in Subsec. III A. However, the PIMC using the original physical object to study. If we use second quantization, Hamiltonian is still able to estimate the average thermal we encapsulate the anti-symmetry constraint, –working energy accurately with similar number of Trotter slices in the subspace of anti-symmetric wavefunctions–, and for the same temperature as for the case of Fig. 2. How- generally see that the corresponding Hamiltonian, say a ever, for this case the Trotter error is clearly worse as we Hubbard model, is not stoquastic when expressed in a can see from the fact that at low M we have larger rela- fermionic Fock or qubit basis. Time-reversal does how- tive error compared to the case of weak coupling. This is ever play a role in some special cases when we avoid the simply due to the larger coupling capacitance and not a sign problem for fermionic systems [53]. signature of a fundamental obstruction. In addition, the orange (light gray) line also shows that similar accuracy can be achieved even if we reduce the temperature so As for complexity, it is important to note that it that the thermal energy gets closer to the ground-state is highly unlikely that one can find computationally- energy. This naturally comes at the price of increasing efficient curing transformations which map any (local) the number of Trotter slices by a factor of three, while Hamiltonian onto a stoquastic Hamiltonian as it would the number of Metropolis iterations was not changed be- have unlikely complexity-theoretic consequences. It was tween the two different temperatures. shown in Ref. [3] that the ground-state energy estimation problem for stoquastic qubit Hamiltonians is a problem contained in AM (and StoqMA ⊆ AM). The class AM is contained in the so-called polynomial hierarchy, while IV. DISCUSSION on the other hand, BQP, let alone QMA, is not believed to be contained in the polynomial hierarchy [54, 55]. Es- In this paper we have seen that qubit Hamiltonians tablishing the precise physical origin of the sign prob- which may appear to be non-stoquastic can have ‘master’ lem, and when it can be avoided, is important as the circuit QED Hamiltonians which are manifestly stoquas- sign problem is precisely what separates quantum from tic. We have used this observation to propose an efficient classical computation: At least, the sign problem neces- simulation method for quantum adiabatic computation sitates the use of quasi-probability distributions (which with such Hamiltonians, using path integral Monte Carlo lead to potentially-exponential variances in Monte Carlo methods. simulations) which can be used quite widely for the sim- It is not entirely straightforward to reconcile the pro- ulation of quantum computation by classical stochastic jected non-stoquastic Hamiltonian in Sec. III A with the means [56–58]. 11

V. ACKNOWLEDGEMENTS Inserting the identity Z Z We thank David DiVincenzo and Joel Klassen for dφ |Φi hΦ| = dQ |Qi hQ| = 1, (A.5) many insightful discussions on the topic of this paper. We thank Marios Ioannou and Sergey Bravyi for discussions in the flux basis M times we obtain on perturbative gadgets. We thank Gioele Consani for 1 Z useful discussions about the SW transformation for cou- hOi ≈ dΦ dΦ . . . dΦ hΦ |O|Φ i × pled flux qubits. We thank Adrian Parra-Rodriguez for a β Z 1 2 M 1 1 useful comment on canonical transformations in the pres- −βK/M −βK/M hΦ1|e |ΦM i ... hΦ2|e |Φ1i × ence of non-reciprocal elements. We are also thankful to −β/M PM U(Φ ) Jan Reiner for an important remark on single-qubit uni- e s=1 s , (A.6) taries curing non-stoquasticity. Our work was supported where we used the fact that O is diagonal in the flux by ERC grant EQEC No. 682726. basis. We thus need to evaluate the matrix element −βK/M hΦs+1| e |Φsi. Using Eq. (A.5) in the charge basis Appendix A: PIMC in the flux basis and [62]

1 T hΦ|Qi = eiQ Φ/~, (A.7) In this Appendix we explicitly derive the PIMC N/2 (2π~) method for the general Hamiltonian in Eq. (2.1) by per- forming a mapping to a classical model and its partition we obtain function. The derivation is a simple adaptation of those −βK/M − β QˆT C−1Qˆ that can be found in Refs. [37, 59], where the only ad- hΦs+1|e |Φsi = hΦs| e 2M |Φs+1i = ditional complication that is added is that the inverse 1 Z i T β T −1 Q (Φs+1−Φs) − 2M Q C Q of the capacitance matrix in Eq. (3.2) is not diagonal. . N dQe ~ e = (2π~) We consider the general task of computing the thermal N √   2   average of an observable O: M M 1/2 2 det C exp − |C (Φs+1 −Φs)| , ~22πβ 2~2β 1   hOi = Tr Oe−βH . (A.1) (A.8) β Z which is clearly positive. Eq. (A.6) becomes We will evaluate the trace in the flux |Φi basis and we Z will further assume that the observable O that we are hOiβ ≈ dΦ1dΦ2 . . . dΦM hΦ1|O|Φ1i p(Φ1,..., ΦM ), evaluating is diagonal in this basis. We remark that one can also evaluate the thermal average of H itself, even (A.9) though it has an off-diagonal kinetic term in the flux ba- where we defined the path probabilities sis. This follows by virtue of the quantum virial theorem [60], which states that that the average of the kinetic en- w(Φ1,..., ΦM ) p(Φ1,..., ΦM ) = = ergy K in any eigenstate of H, and thus also for thermal Z M NM averages, satisfies (det C) 2  M  2 2 exp(−βHc) (A.10) N Z ~ 2πβ 1 X ∂U  hKi = Φˆ . (A.2) β k with periodic boundary condition Φ = Φ , and clas- 2 ∂Φk β M+1 1 k=1 sical Hamiltonian given in Eq. (2.6) in the main text. We make use of this result to evaluate the thermal ener- Notice that all path probabilities are positive and they gies for the PIMC in the flux basis discussed in Subsec. are correctly normalized since by repeating the previous III C. derivation we can write the partition function as For general off-diagonal observables O there is no rig- Z orous relation between hOiβ and the evaluation of Z. Z = dΦ1dΦ2 . . . dΦM w(Φ1,..., ΦM ). (A.11) Let us start by rewriting Eq. (A.1) as Z We thus have written the thermal average of a diagonal 1 −βH/M −βH/M hOiβ ≈ dΦ1 hΦ1|O e . . . e |Φ1i , operator as the average of an estimator hΦ1|O|Φ1i over Z | {z } M times a classical probability distribution. We remark that also PM (A.3) m=1 hΦm|O|Φmi /M is a valid, unbiased estimator. where the integral is over RN and we compactly denote Eq. (A.9) is the basis for the PIMC method, where QN dΦ1 = k=1 dΦ1k. Assuming β/M  1 we can use we sample from the probability distribution over path Trotter’s break-up formula [61] and approximate configurations p(Φ1,..., ΦM ) for instance by using the Metropolis-Hastings algorithm [63, 64] that we detail in e−βH/M = e−β(K+U)/M ≈ e−βK/M e−βU/M . (A.4) the next Subsection A 1 for completeness. 12

1. Metropolis-Hastings reviewed One can also come up with mixed strategies. As pointed out in [59] it is generally good to have a variety of up- The Metropolis-Hastings algorithm allows to sample date rules that we select with a certain probability. These from an arbitrary probability distribution, in our case considerations are however always dependent on the par- p(Φ), given the ability to compute a function f(Φ) pro- ticular system we are dealing with. portional to it, ı.e., f(Φ) = cp(Φ) for some c ∈ R. In our case 2. Details of the Monte Carlo simulations p(Φ) 0 0 = exp(−β(Hc(Φ) − Hc(Φ ))) (A.12) p(Φ ) We give some details of the Monte Carlo simulations The algorithm works as follows. discussed in Sec. III A. In Fig. 2, we initialize in both the PIMC and TIM simulations the corresponding classi- (k) (k) 1. Choose an initial configuration (Φ1 ,..., ΦM ), cal system in a random configuration. We let the system k = 0. The initial configuration can be chosen ran- equilibrate for 5 × 106 Metropolis iterations, after which domly, but this is not necessary. we start to sample the energy every 1000 iterations. We 6 0 0 continue to run the Metropolis algorithm until 30 × 10 2. Propose a new configuration (Φ1,..., ΦM ) accord- iterations are reached. In the calculation of the error ing to some probability distribution (transition bars we take into account the correction due to the cor- rule). Evaluate the variation of the Hamiltonian k 0 (k) relation between the samples by explicitly computing the (energy) ∆H = Hc(Φ ) − Hc(Φ ). It is assumed autocorrelation time of the samples. This explains why that the transition rules are chosen such that the the error bars are increasing with the number of Trotter probability for a transition from (Φ1,..., ΦM ) to 0 0 slices M in Fig. 2, since if we fix the number of iterations, (Φ1,..., ΦM ) is the same as that of transition from 0 0 we expect the autocorrelation time to increase with M. (Φ1,..., ΦM ) to (Φ1,..., ΦM ) (Markov chain is For the PIMC in the flux basis we apply local updates symmetric). with probability 0.9, while otherwise we attempt a global 3. Accept the new configuration and set update. In both cases, we attempt to modify the chosen (k+1) (k+1) flux variables by shifting them by a certain δ from a uni- (Φ ,..., Φ ) = (Φ0 ,..., Φ0 ) with 1 M 1 M form distribution in [−0.75, 0.75] (see discussion in the probability previous subsection). The same procedure is applied for   −β∆Hk Fig. 3. A similar update rule is applied for the PIMC p = min 1, e , (A.13) derived from the TIM model. With probability 0.9 we apply a local update where we suggest to flip a random (k+1) (k+1) (k) (k) spin. Otherwise, we attempt to flip all spins. otherwise (Φ1 ,..., ΦM ) = (Φ1 ,..., ΦM ). 4. Update k = k + 1 and go to 2. Appendix B: Stoquasticity of the Cooper-pair box 5. Halt the algorithm when a sufficient number of con- figurations have been generated from which we can compute the desired averages as arithmetic aver- The quantum Hamiltonian of a Cooper-pair box or ages. transmon qubit is We see that the Metropolis-Hastings algorithm gener- 1 H = (Qˆ − Q )2 − E (cos(2πΦˆ/Φ ), (B.1) ates a Markov chain whose equilibrium distribution can transmon 2C g J 0 be shown to be the desired probability distribution. h Thus, we should start to average only when equilibrium with Φ0 = 2e , as a special case of Eq. (2.9). The conju- is reached. Also, the performance of the algorithm is gate operators flux Φˆ and charge Qˆ take eigenvalues in R strongly influenced by the choice of the transition rule. so that this shifted Hamiltonian can be made manifestly These can be broadly distinguished into two main cate- stoquastic in the flux qubit basis by a simple transforma- gories: tion, as discussed in the main text, namely Eq. (2.11). When treating the Hamiltonian in Eq. (B.1), one often 1. local update: at step k a random particle s with s = moves to a rotor basis defined by a 2π-periodic phase ϕ 1,...,M in imaginary time is chosen and its config- ˆ 0 k and integer n ∈ Z [39]. We can indeed convert from Φ uration is randomly changed as Φs = Φs +δ where ˆ δ is a N-dimensional vector of random variables, and Q toϕ ˆ andn ˆ by defining the basis usually chosen uniformly within a range [−∆, ∆] X for some ∆; |ϕi = |φ = ϕ + 2πki , (B.2) k∈Z 2. global update: at step k all particles are shifted by the same N-dimensional vector δ of random vari- with φ = 2πΦ . This basis |ϕi is an eigenbasis for the Φ0 ables. subspace of the oscillator space defined by the operator 13

SQ = exp(iπQ/eˆ ) taking eigenvalue 1. In this (rotor) subspace we thus have that Qˆ = 2enˆ takes eigenvalues 2en with n ∈ Z, which is interpreted as there being an offset of n Cooper pairs with total charge 2en on the superconducting island defining the transmon qubit. In this subspace the transmon Hamiltonian of Eq. (B.1) equals

2 Htransmon,sub = 4EC (ˆn − ng) − EJ cos(ϕ ˆ), (B.3) with offset charge ng ∈ [0, 1) and Qg = 2eng. We could FIG. 4: Circuit of a flux qubit. The dynamical variable 2πΦ have picked another rotor subspace in which SQ takes φ is given by φ = with Φ the flux across the Φ0 i2πn˜g the eigenvalue, say, e for somen ˜g. The basis for inductor. Analogously, the external fluxes in the x this subspace is superconducting loops Φα, α ∈ {cjj, q} are given in 2πΦx terms of φx = α . X α Φ0 |ϕi = e2πin˜g k |φ = ϕ + 2πki . (B.4) n˜g k∈Z is shown in Fig. 4. Notice that in this circuit we are ne- i2πn˜g since SQ |ϕi = e |ϕi . n˜g n˜g glecting the small inductance of the SQUID loop. While The spectrum and eigenstates of Htransmon,sub in there are also other flux qubit designs [40, 67–69], we here Eq. (B.3) relate to eigensolutions of the Mathieu equa- focus on this simple circuit since it captures the funda- tion [39] and depend on ng. For ng 6= 0, the ground-state mental physics behind the flux qubit and it is also the R 2π |ψ0i = 0 dϕ ψ0(ϕ) |ϕi has a complex wavefunction design used in Ref. [1]. ψ0(ϕ) [39, 65]. For ng = 0, the wavefunction ψ0(ϕ) ≥ 0. The Hamiltonian of the circuit in Fig. 4 reads We can consider in which subspace the Hamiltonian has a ground-state with minimal energy overall. We ob- 2 EL ˆ2 eff x  ˆ x H = 4EC qˆ + φ − EJ φcjj cos φ + φq , (C.1) serve that by going to the subspace in whichn ˜g = ng, we 2 obtain a Hamiltonian as in Eq. (B.3) with ng = 0. The e2 standard spectrum of the transmon qubit [39] shows that where we defined the charging energy EC = 2C and the 2 Φ0 this choice achieves the lowest energy eigenvalue. Hence inductive energy EL = 4π2L . The external flux in the x the global ground-state is a non-negative wavefunction SQUID loop φcjj allows to control the effective Josephson in the subspace basis |ϕi . We observe that the ba- n˜g energy via the relation sis |ϕi is non-negatively related to the transformed n˜g =ng  x  φcjj 0 iΦˆQ / eff x  basis |Φ i = e g ~ |Φi in which the original Hamilto- EJ φcjj = EJ cos . (C.2) ˆ 2 nian was explicitly stoquastic: this holds as eiΦQg /~ |ϕi = |ϕi . x ng In what follows we will assume φ ∈ [−π, π). Thus we see that the fact that the ground-state wave- cjj The flux qubit is operated in the regime E ,Eeff  function is complex in some rotor subspace is entirely L J E , φx ≈ π and Eeff /E 1 [40, 70]. compatible with the stoquasticity of the Hamiltonian C q J L & With these conditions the potential becomes a dou- (when considered in the full space and in the right basis). ble well potential. The computational qubit basis is de- On a separate note, the convergence and accurate pre- fined by considering the case of a symmetric potential dictions of the Monte Carlo path integral simulation of obtained for φx = π. In this case the eigenstates obey a the transmon qubit in the subspace labeled byn ˜ can q g parity symmetry and consequently they are either even be examined. It can depend on whether the numerical or odd in the flux representation. An example of the simulation varies the winding number or not [66]. Here wave-functions for the ground-state |gi and first-excited the winding number is the number of times the phase ϕ state |ei is shown in Fig. 5a. The computational ba- wraps around 2π in the path integral. sis is defined by taking symmetric and anti-symmetric superpositions of |gi and |ei as Appendix C: Flux qubit Hamiltonians 1 |0i = √ (|gi + |ei) (C.3a) 2 1. The flux qubit reviewed 1 |1i = √ (|gi − |ei). (C.3b) 2 We briefly review the Hamiltonian of the flux qubit circuit and its mapping to a qubit model. A similar dis- As we see from Fig. 5b, the computational basis states cussion can be found in Refs. [40, 51]. The basic circuit |0i, |1i are localized on the left and right well, respec- of a compound Josephson junction rf-SQUID flux qubit tively. They correspond to anti-clockwise and clockwise 14

Let us now consider the asymmetric case in which we x x x slightly bias φq away from π, ı.e., we take φq = π + δq . x By expanding the cosine in Eq. (C.1) to first order in δq we obtain

E H ≈ 4E qˆ2 + L φˆ2 + Eeff φx  cos φˆ C 2 J cjj eff x ˆ − EJ δq sin φ = Hsym + V. (C.6)

The Hamiltonian is now given by the Hamiltonian in the symmetric case plus a perturbation

eff x ˆ V = −EJ δq sin φ. (C.7)

By projecting V onto the computational subspace we ob- (a) tain a term that is (by design) diagonal in the compu- tational basis, since h0| sin φˆ|1i = 0. In addition, we can also neglect the coupling that V induces to other energy levels with higher energy. The projected qubit Hamilto- nian in the asymmetric case then becomes

∆ ε H /h = − X − Z, (C.8) q 2 2 where we defined

Eeff ε = J δxh0| sin φˆ|0i − h1| sin φˆ|1i = h q Eeff 2 J δx h0| sin φˆ|0i . (C.9) h q

x For relatively large asymmetry of the potential δq and (b) small ∆ the computational basis states also become the x eigenbasis. In addition, the parameter δq can be used to FIG. 5: Flux qubit with symmetric potential. a) control the strength of the parameter ε, independently of Ground and first-excited wave-functions in φ, and their the tunnel coupling ∆. The independent tunability of ∆ energies as dashed lines. b) Computational basis states and ε by means of external fluxes is one of the features obtained as symmetric and anti-symmetric combination that makes flux qubits suited for quantum annealing al- of the first two eigenstates. The relevant parameters are gorithms [71]. In addition, an inductive coupling between taken as in Table I. two flux qubits with index k, l would give a term in the ˆ ˆ Hamiltonian proportional to φkφl, which when projected onto the computational basis gives a term ∼ ZkZl, thus average persistent currents in the loop formed by the in- realizing a TIM. ductor and the SQUID in Fig. 4. The energy difference between ground and first-excited state in case of a sym- metric potential 2. Two-flux qubit Hamiltonian with symmetric E − E double well potentials ∆ = e g (C.4) h We analyze a two-flux qubit Hamiltonian with induc- is usually called the tunnel coupling. tive and capacitive couplings, as in Fig. 1, and choose By projecting onto the computational subspace the symmetric double wells for both flux qubits. As we Hamiltonian with symmetric potential is will see this implies parity symmetry of the Hamilto- ∆ nian. Based on this symmetry we show that the effective H /h = − X. (C.5) qubit Hamiltonian is always stoquastic by a simple ba- q 2 sis change, at any order in SW perturbation theory [72]. x The external flux in the SQUID loop φcjj can be used This shows that asymmetry in the flux qubit potential is to control the height of the barrier and thus, the tunnel necessary to get an effective non-stoquastic qubit Hamil- coupling ∆. tonian. 15

First, let the global parity operatorπ ˆ be a unitary, Yk = −i |gi he|k + i |ei hg|k , (C.15b) ˆ Hermitian operator, defined via its action on φk,q ˆk as

ˆ ˆ Zk = |gi hg|k − |ei he|k , (C.15c) πˆφkπˆ = −φk, πˆqˆkπˆ = −qˆk. (C.10)

† the effective qubit Hamiltonian reads Note thatπ ˆ = Π exp(iπa a ) with φˆ = √1 (a + k k k k 2 k † i † a ), qˆk = √ (a − ak). ∆1 ∆2 k 2 k H2q/h = P0HP0 = − Z1 − Z2+ Sinceπ ˆ2 = 1 the parity operator has eigenvalues ±1. 2 2 We call an operator O parity symmetric whenπO ˆ πˆ = O. JXX X1X2 + JYY Y1Y2, (C.16) The Hamiltonian of the two flux qubits in case φx = π q where the tunnel couplings ∆ are defined as in Eq. (symmetric double wells) reads 1,2 (C.4), JYY is given by 2 E X 2 1 2 eff C12 H = 4E qˆ + E φˆ + E cos φˆ JYY = − hg|qˆ1|ei hg|qˆ2|ei (C.17) Ck k 2 Lk k Jk k h 1 2 k=1 ˆ ˆ and the XX coupling JXX is + 8EC12qˆ1qˆ2 + EL12φ1φ2, (C.11) E and we can define the single flux qubit Hamiltonian J = L12 hg|φˆ |ei hg|φˆ |ei . (C.18) XX h 1 1 2 2 1 2 ˆ2 eff ˆ The Hamiltonian in Eq. (C.16) can always be made Hk = 4ECkqˆk + ELkφk + EJk cos φk, (C.12) 2 stoquastic by simple Clifford transformations. The con- ditions for stoquasticity is here that J ≤ −|J |. k = 1, 2, with ECk and ELk the diagonal elements of XX YY the charging energy and inductive energy matrix respec- If |JXX | ≥ |JYY | apply the transformation X1 7→ tively. −sign(JXX )X1,Z1 7→ −sign(JXX )Z1 and then the We can write Eq. (C.11) as Hamiltonian is stoquastic. If |JXX | < |JYY | apply the transformation that exchanges X and Y on both qubits, H = H0 + V, (C.13) and use the previous transformation. The previous result relies on the validity of the projec- with H0 = H1 +H2, ı.e., the uncoupled flux qubit Hamil- tion onto the computational subspace. A natural ques- tonians, and tion to ask is whether the effective qubit Hamiltonian can still always be made stoquastic if the perturbation theory ˆ ˆ V = 8EC12qˆ1qˆ2 + EL12φ1φ2, (C.14) is refined. We here show this is indeed the case by using a Schrieffer-Wolff transformation [50]. Note that this was where EC12 and EL12 denote the off-diagonal element of also used in Ref. [49] to obtain the effective Hamiltonian EC and EL respectively. Clearly, both H0 as well as V of two capacitively and inductively coupled flux qubits, are invariant under the global parity transformationπ ˆ. and to study its stoquasticity. In the remaining part of this section the index k will To properly discuss the SW transformation we recall always be k = 1, 2. The Hamiltonians Hk admit only some notions from Ref. [50]. Let λmin and λmax be the bound states as eigenstates. As a consequence the av- P0 P0 minimum and maximum eigenvalues of H0 with eigenvec- erage ofq ˆk in any eigenstate |ψi of Hk is zero, ı.e., min max k tors in P0, respectively, and let I0 = [λP , λP ] ⊆ R. hψ|qˆ |ψi = 0 [73]. 0 0 k k We define the energy gap Λ = λmin −λmax, where λmin is Due to parity symmetry, the eigenstates of Hamiltoni- Q0 P0 Q0 the minimum eigenvalue of H0 whose eigenvectors is in ans H0 and H can be chosen as eigenstates of the par- the complement subspace Q0 of P0. We introduce a new ity operator with eigenvalues ±1. This implies that the interval I = [λmin − Λ/2, λmax + Λ/2] ⊆ , and the sub- eigenstate wave-functions in flux are either even or odd P0 P0 R ˆ space P with projector P spanned by the eigenvectors of functions, and that hψ|φk|ψik = 0. H with eigenvelue in I. Our general goal is to obtain an We now assume the validity of first order perturbation effective Hamiltonian that is block-diagonal with respect theory in the eigenbasis of the Hk and obtain an effective to P0 and Q0, ı.e., P0Heff Q0 = Q0Heff P0 = 0 and has two-qubit Hamiltonian by projecting Eq. (C.11) onto the the same spectrum as H. In particular, by projecting subspace P0 spanned by the first two levels of each sub- the effective Hamiltonian onto P0 we obtain a reduced system |gi , |ei . This consists in applying the projector k k Hamiltonian H0eff = P0Heff P0 with spectrum in I. The SW transformation is a unitary transformation defined P = P ⊗ P = |gi hg| + |ei he|  ⊗ |gi hg| + |ei he|  0 1 2 1 1 2 2 as to the Hamiltonian in Eq. (C.11). By defining our Pauli U = exp(S) = p(1 − 2P )(1 − 2P ), (C.19) operators in the eigenbasis as 0 where S is a block-off-diagonal, anti-hermitian opera- X = |gi he| + |ei hg| , (C.15a) k k k tor with respect to P0,Q0. In order for the SW to be 16 uniquely defined we require the condition kSk < π/2 (C.15) as [50]. The (exact) effective Hamiltonian H0eff is πXˆ kπˆ = −Xk, πYˆ kπˆ = −Yk, πZˆ kπˆ = Zk. (C.22) † H0eff = P0UHU P0. (C.20) Thus, the only terms allowed in H0eff in order to satisfy the parity symmetry, and the fact that the Since the Hamiltonians H and H0 are invariant under Hamiltonian is real, are local Z1,2 and the interactions parity transformations, also the projectors P , P0 satisfy X1X2,Y1Y2,Z1Z2. the parity symmetry. Consequently also the unitary U Hence, compared to the lowest-order SW projection in Eq. (C.19) and the generator S are parity symmetric, only the Z1Z2 term can be added, and since this term and thus also H0eff , ı.e., is diagonal we can employ the same Clifford transforma- tions as before to cure the non-stoquasticity. πHˆ 0eff πˆ = H0eff . (C.21) We note that this observation does not immediately generalize to multiple flux qubits as the SW transfor- Sinceπ ˆ |gik = +1 |gik andπ ˆ |eik = −1 |eik, the parity mation may introduce k-local terms and it is not clear operatorπ ˆ acts on the Pauli operators defined in Eq. whether the parity symmetry would suffice in that case.

[1] I. Ozfidan et al., Demonstration of a nonstoquastic of the efficiency of quantum annealing by non-stoquastic Hamiltonian in coupled superconducting flux qubits, Hamiltonians, Frontiers in ICT 4, 2 (2017). Phys. Rev. Applied 13, 034037 (2020). [17] T. Albash, Role of nonstoquastic catalysts in quantum [2] A. Y. Kitaev, A. H. Shen, and M. N. Vyalyi, Classical adiabatic optimization, Phys. Rev. A 99, 042334 (2019). and Quantum Computation (American Mathematical So- [18] T. Albash, Validating a two-qubit nonstoquastic Hamil- ciety, Boston, MA, USA, 2002). tonian in quantum annealing, Phys. Rev. A 101, 012310 [3] S. Bravyi, D. P. DiVincenzo, R. I. Oliveira, and B. M. (2020). Terhal, The complexity of stoquastic local Hamiltonian [19] E. J. Crosson and D. A. Lidar, Prospects for quantum problems, Quant. Inf. Comp 8, 0361 (2008). enhancement with diabatic quantum annealing (2020), [4] S. Bravyi, A. J. Bessen, and B. M. Terhal, Merlin-Arthur arXiv:2008.09913. Games and Stoquastic Complexity (2006), arXiv:quant- [20] U. Vool and M. Devoret, Introduction to quantum elec- ph/0611021. tromagnetic circuits, International Journal of Circuit [5] T. S. Cubitt, A. Montanaro, and S. Piddock, Univer- Theory and Applications 45, 897 (2017). sal quantum hamiltonians, Proceedings of the National [21] A. Blais, A. L. Grimsmo, S. Girvin, and A. Wall- Academy of Sciences 115, 9497 (2018). raff, Circuit Quantum Electrodynamics (2020), [6] S. Bravyi and B. Terhal, Complexity of stoquastic arXiv:2005.12667. frustration-free Hamiltonians, SIAM Journal on Comput- [22] X. Gu, A. F. Kockum, A. Miranowicz, Y. xi Liu, ing 39, 1462 (2010). and F. Nori, Microwave photonics with superconducting [7] M. B. Hastings, The power of adiabatic quantum com- quantum circuits, Physics Reports 718-719, 1 (2017). putation with no sign problem (2020), arXiv:2005.03791. [23] M. Marvian, D. A. Lidar, and I. Hen, On the computa- [8] A. Gily´enand U. Vazirani, (Sub)Exponential advantage tional complexity of curing non-stoquastic Hamiltonians, of adiabatic quantum computation with no sign problem Nature Comm. 10, 1571 (2019). (2020), arXiv:2011.09495. [24] J. Klassen and B. M. Terhal, Two-local qubit Hamiltoni- [9] T. Albash and D. A. Lidar, Adiabatic quantum compu- ans: when are they stoquastic?, Quantum 3, 139 (2019). tation, Rev. Mod. Phys. 90, 015002 (2018). [25] J. Klassen, M. Marvian, S. Piddock, M. Ioannou, I. Hen, [10] R. Harris et al., Compound Josephson-junction coupler and B. M. Terhal, Hardness and Ease of Curing the for flux qubits with minimal crosstalk, Phys. Rev. B 80, Sign Problem for Two-Local Qubit Hamiltonians (2019), 052506 (2009). arXiv:1906.08800. [11] T. Kadowaki and H. Nishimori, Quantum annealing in [26] M. Ioannou, S. Piddock, M. Marvian, J. Klassen, and the transverse Ising model, Phys. Rev. E 58, 5355 (1998). B. M. Terhal, Sign-curing local Hamiltonians: termwise [12] G. E. Santoro, R. Martoˇn´ak,E. Tosatti, and R. Car, The- versus global stoquasticity and the use of Clifford trans- ory of quantum annealing of an Ising spin glass, Science formations (2020), arXiv:2007.11964. 295, 2427 (2002). [27] D. Hangleiter, I. Roth, D. Nagaj, and J. Eisert, Eas- [13] M. W. Johnson et al., Quantum annealing with manu- ing the monte carlo sign problem, Science Advances 6, factured spins, Nature 473, 194 (2011). 10.1126/sciadv.abb8341 (2020). [14] S. Isakov et al., Understanding quantum tunneling [28] One can easily construct an example of a 3-qubit stoquas- through quantum Monte Carlo simulations, Phys. Rev. tic Hamiltonian which has a non-stoquastic two-qubit Lett. 117, 180402 (2016). low-energy effective Hamiltonian without permitting ad- [15] L. Hormozi, E. W. Brown, G. Carleo, and M. Troyer, ditional local basis changes. Nonstoquastic Hamiltonians and quantum annealing of [29] C. M. Bender and T. T. Wu, Anharmonic oscillator, an Ising spin glass, Phys. Rev. B 95, 184416 (2017). Phys. Rev. 184, 1231 (1969). [16] H. Nishimori and K. Takada, Exponential enhancement [30] J. Koch, A. A. Houck, K. L. Hur, and S. M. Girvin, 17

Time-reversal-symmetry breaking in circuit-QED-based Annals of Physics 326, 2793 (2011). photon lattices, Phys. Rev. A 82, 043811 (2010). [51] M. Willsch, D. Willsch, F. Jin, H. De Raedt, and [31] J. E. Mooij and Y. V. Nazarov, Superconducting K. Michielsen, Real-time simulation of flux qubits used nanowires as quantum phase-slip junctions, Nature for quantum annealing, Phys. Rev. A 101, 012327 (2020). Physics 2, 169 (2006). [52] R. Martoˇn´ak,G. E. Santoro, and E. Tosatti, Quantum [32] D. T. Le, A. Grimsmo, C. M¨uller,and T. M. Stace, Dou- annealing by the path-integral Monte Carlo method: The bly nonlinear superconducting qubit, Phys. Rev. A 100, two-dimensional random Ising model, Phys. Rev. B 66, 062321 (2019). 094203 (2002). [33] Using standard methods, the case when C is not invert- [53] C. Wu and S.-C. Zhang, Sufficient condition for ab- ible can be treated separately, ı.e., the modes with zero sence of the sign problem in the fermionic quantum energy are eliminated as they have no dynamics. Monte Carlo algorithm, Phys. Rev. B 71, 10.1103/phys- [34] A. Berman and R. J. Plemmons, Nonnegative Matrices revb.71.155115 (2005). in the Mathematical Sciences (Society for Industrial and [54] R. Raz and A. Tal, Oracle separation of BQP and PH, Applied Mathematics, 1994). in Proceedings of the 51st Annual ACM SIGACT Sympo- [35] G. Burkard, R. H. Koch, and D. P. DiVincenzo, Mul- sium on Theory of Computing, STOC 2019 (Association tilevel quantum description of decoherence in supercon- for Computing Machinery, New York, NY, USA, 2019) ducting qubits, Phys. Rev. B 69, 064503 (2004). p. 13–23. [36] T. Halverson, L. Gupta, M. Goldstein, and I. Hen, Ef- [55] S. Aaronson, BQP and the Polynomial Hierarchy, in Pro- ficient simulation of so-called non-stoquastic supercon- ceedings of the Forty-Second ACM Symposium on The- ducting flux circuits (2020), arXiv:2011.03831. ory of Computing, STOC ’10 (Association for Computing [37] D. Landau and K. Binder, A Guide to Monte Carlo Machinery, New York, NY, USA, 2010) p. 141–150. Simulations in Statistical Physics (Cambridge University [56] H. Pashayan, J. J. Wallman, and S. D. Bartlett, Esti- Press, New York, NY, 2005). mating outcome probabilities of quantum circuits using [38] H. Kleinert, Path integrals in quantum mechanics, statis- quasiprobabilities, Phys. Rev. Lett. 115, 070501 (2015). tics, polymer physics, and financial markets; 3rd ed. [57] M. Howard and E. Campbell, Application of a resource (World Scientific, River Edge, NJ, 2004). theory for magic states to fault-tolerant quantum com- [39] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. puting, Phys. Rev. Lett. 118, 090501 (2017). Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, [58] A. Mari and J. Eisert, Positive wigner functions ren- and R. J. Schoelkopf, Charge-insensitive qubit design de- der classical simulation of quantum computation efficient, rived from the cooper pair box, Phys. Rev. A 76, 042319 Phys. Rev. Lett. 109, 230503 (2012). (2007). [59] D. M. Ceperley, Path integrals in the theory of condensed [40] R. Harris et al., Experimental demonstration of a robust Helium, Rev. Mod. Phys. 67, 279 (1995). and scalable flux qubit, Phys. Rev. B 81, 134510 (2010). [60] V. Fock, Bemerkung zum Virialsatz, Zeitschrift f¨ur [41] M. Rymarz, The Quantum Electrodynamics of Singu- Physik 63, 855 (1930). lar and Nonreciprocal Superconducting Circuits, Master’s [61] M. Suzuki, Generalized Trotter’s formula and systematic thesis, RWTH Aachen (2018). approximants of exponential operators and inner deriva- [42] M. Rymarz, S. Bosco, A. Ciani, and D. P. DiVincenzo, tions with applications to many-body problems, Comm. Hardware-encoding grid states in a nonreciprocal super- Math. Phys. 51, 183 (1976). conducting circuit, Phys. Rev. X 11, 011032 (2021). [62] J. J. Sakurai and J. Napolitano, Modern Quantum Me- [43] A. Parra-Rodriguez, I. L. Egusquiza, D. P. DiVincenzo, chanics, 2nd ed. (Cambridge University Press, 2017). and E. Solano, Canonical circuit quantization with linear [63] N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, nonreciprocal devices, Phys. Rev. B 99, 014514 (2019). A. H. Teller, and E. Teller, Equation of state calculations [44] K. M. Sliwa, M. Hatridge, A. Narla, S. Shankar, L. Frun- by fast computing machines, The Journal of Chemical zio, R. J. Schoelkopf, and M. H. Devoret, Reconfigurable Physics 21, 1087 (1953). josephson circulator/directional amplifier, Phys. Rev. X [64] W. K. Hastings, Monte Carlo sampling methods using 5, 041020 (2015). Markov chains and their applications, Biometrika 57, 97 [45] G. Viola and D. P. DiVincenzo, Hall effect gyrators and (1970). circulators, Phys. Rev. X 4, 021019 (2014). [65] A. Cottet, Implementation d’un bit quantique dans un [46] D. De Bernardis, P. Pilar, T. Jaako, S. De Liberato, and circuit supraconducteur, Ph.D. thesis, UNIVERSITE P. Rabl, Breakdown of gauge invariance in ultrastrong- PARIS VI (2002). coupling cavity qed, Phys. Rev. A 98, 053819 (2018). [66] P. Henelius, S. M. Girvin, and A. W. Sandvik, Role of [47] O. Di Stefano, A. Settineri, V. Macr`ı, L. Garziano, winding numbers in quantum monte carlo simulations, R. Stassi, S. Savasta, and F. Nori, Resolution of gauge Phys. Rev. B 57, 13382 (1998). ambiguities in ultrastrong-coupling cavity quantum elec- [67] T. P. Orlando, J. E. Mooij, L. Tian, C. H. van der Wal, trodynamics, Nature Physics 15, 803 (2019). L. S. Levitov, S. Lloyd, and J. J. Mazo, Superconducting [48] L. B. Nguyen, Y.-H. Lin, A. Somoroff, R. Mencia, persistent-current qubit, Phys. Rev. B 60, 15398 (1999). N. Grabon, and V. E. Manucharyan, High-coherence flux- [68] J. Q. You, X. Hu, S. Ashhab, and F. Nori, Low- onium qubit, Phys. Rev. X 9, 041041 (2019). decoherence flux qubit, Phys. Rev. B 75, 140515 (2007). [49] G. Consani and P. A. Warburton, Effective hamiltoni- [69] F. Yan et al., The flux qubit revisited to enhance coher- ans for interacting superconducting qubits: local basis ence and reproducability, Nature Comm. 7, 12964 (2016). reduction and the Schrieffer–Wolff transformation, New [70] G. Wendin and V. S. Shumeiko, Superconducting Quan- Journal of Physics 22, 053040 (2020). tum Circuits, Qubits and Computing (2005), arXiv:cond- [50] S. Bravyi, D. P. DiVincenzo, and D. Loss, Schrieffer- mat/0508729. Wolff transformation for quantum many-body systems, [71] M. Johnson et al., A scalable control system for a su- 18

perconducting adiabatic quantum optimization proces- well configuration since the Hamiltonian is effectively the sor, Superconductor Science and Technology 23, 065004 same as for flux qubits, but in the parameter regime EL < eff (2010). EC < EJ . [72] The analysis applies also to fluxonium qubits in a double [73] A. Messiah, Quantum Mechanics (Dover Publications, Mineola, NY, 1999).