This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy. Close this notification
Paper The following article is Open access

A transmon quantum annealer: decomposing many-body Ising constraints into pair interactions

, and

Published 16 December 2016 © 2016 IOP Publishing Ltd
, , Citation Martin Leib et al 2016 Quantum Sci. Technol. 1 015008 DOI 10.1088/2058-9565/1/1/015008

2058-9565/1/1/015008

Abstract

Adiabatic quantum computing is an analogue quantum computing scheme with various applications in solving optimisation problems. In the parity picture of quantum optimization, the problem is encoded in local fields that act on qubits that are connected via local four-body terms We present an implementation of a parity annealer with Transmon qubits with a specifically tailored Ising interaction from Josephson ring modulators.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Among superconducting qubits, Transmons are a leading platform with respect to decay and dephasing times [14]. They are charge qubits operated in a regime with strong resilience to the ubiquitous charge noise in any superconducting qubit devices. With this insensitivity to radio frequencies and tunability by fast microwave signals, they are considered ideal candidates for gate-based quantum computing. For analogue applications such as adiabatic quantum computing (AQC), Transmons have not been considered so far.

Adiabatic quantum computing [5] has recently gained considerable attention because of the prospect to solve problems that can be mapped to classical optimisations [69] efficiently [1016]. In the spin-glass formulation of adiabatic quantum annealing [5], the system is slowly switched from the ground state of a trivial initial Hamiltonian in ${\sigma }_{x}$-basis to an infinite range Ising model in ${\sigma }_{z}$-basis [17]. The optimisation problem is encoded in the interactions between the spins in the final state. There are major challenges related to this protocol: a fundamental challenge in this scheme is the required all-to-all connectivity [18, 19]. Infinite range connectivity is required, while natural qubit interactions are finite range [20, 21]. Another fundamental question is the possible quantum speedup due to the scaling of the minimal gap [16] and the sensitivity to errors [2224]. The parity adiabatic quantum optimization scheme [25] has been recently proposed to encode arbitrary connected models in a larger Hilbert space with local constraints on a square lattice. The optimisation problem is, in contrast to minor embedding schemes [20, 21, 2628], encoded in local fields that act on the qubits and the interactions are four-body constraints that are independent of the problem.

In this paper, we propose an implementation of the parity adiabatic quantum optimization scheme [25] with Transmons. We focus on a coherent protocol that operates in the rotating frame and unlike in previous annealing schemes, the dynamics do not include the interplay between incoherent thermal noise and quantum fluctuations [14]. To implement the required four-body constraints for the parity scheme, we use a general recursive decomposition of classical k-local Ising terms The construction consists of two steps: a recursive decomposition of k-local terms to three-body terms and a decomposition of each three-body term into two-body terms The latter decomposition is similar to ancilla-based, non-perturbative decompositions in [9, 2931]. We present an implementation where all pair interactions have the same sign, are next neighbour and do not cross each other. Note that, the k-body terms in the parity scheme are constraints which only require the ground state to be reproduced. In the parity scheme, the ${\sigma }_{z}{\sigma }_{z}$ interactions do not require individual time-dependent tuning and can in principle be always on. However, an always-on protocol is more challenging as the initial state is not a trivial product state as in the standard ramp protocol.

The paper is structured as follows. In section 2.1, we briefly review the parity adiabatic quantum computing scheme. In section 2.2, we present the general decomposition of Ising-type constraints which we apply in section 2.3 to the four-body terms of the parity adiabatic quantum computing (PAQC) scheme. In section 2.4, we present numerical calculations for a single plaquette to validate that the minimal gap during the annealing sweep is not altered substantially by the introduction of the ancilla spins. In section 3, we present the implementation of parity adiabatic quantum computing with decomposed Ising constraints with Transmon qubits.

2. Parity adiabatic quantum computing with pair interactions

AQC maps the solution to a complicated optimisation problem into the ground state of a specifically designed Hamilton operator ${H}^{({\rm{final}})}$ [17]. Rather than cooling into the ground state of a system that is described by ${H}^{({\rm{final}})}$, which is too slow, one prepares the system at a different point in parameter space described by ${H}^{({\rm{initial}})}$ in its ground state. Afterward, a adiabatic sweep in parameter space prepares the system in the ground state of the problem Hamilton operator ${H}^{({\rm{final}})}$. This is only advantageous if there is a known efficient technique to prepare the system in the ground state of ${H}^{({\rm{initial}})}$ and if the maximal sweep speed scales polynomial with the problem size. One can think of AQC as the slow deformation of a complex high-dimensional energy landscape. The state of the system is prepared in a trivial valley of the energy landscape and the adiabatic sweep deforms the energy landscape while the system tries to stay in the instantaneous valley. To unlock the true potential of quantum mechanics, we have to enable the system to tunnel through energy barriers to the deepest valley. This is accomplished by choosing initial and final Hamilton operators that do not commute $[{H}^{({\rm{final}})},{H}^{({\rm{initial}})}]\ne 0$

In the spin-glass paradigm of AQC, one uses a spin-glass with on-site fields and all-to-all Ising interactions (c.f. figure 2(a)),

Equation (1)

Here, ${\tilde{\sigma }}_{\{x,y,z\}}^{(i)}$ are the Pauli operators of spins in the spin-glass annealing paradigm. The parameter space in spin-glass annealing is given by the set of all on-site fields (${h}_{z,i}$ and ${h}_{x,i}$) and Ising interaction strengths (${J}_{i,j}$). The initial point in parameter space could be the point of dominating on-site transverse fields $| {h}_{x,i}| \gg | {h}_{z,j}| $ and $| {h}_{x,i}| \gg | {J}_{i,j}| $ $\forall i,j$ and the corresponding ground state would be the fully separable state $| 1\rangle ={\bigotimes }_{i}(| \downarrow \rangle -| \uparrow )/\sqrt{2}$. The on-site fields and Ising interaction strengths are slowly deformed to prepare the spin-glass in the ground state of ${H}^{({\rm{final}})}$ given in parameter space by  ${h}_{z,i}\gg | {h}_{x,j}| $ and $| {J}_{i,j}| \gg | {h}_{x,i}| $ $\forall i,j$. The actual values of ${h}_{z,i}$ and ${J}_{i,j}$ encode the optimisation problem. The spin-glass remains in its ground state if the sweep is slow enough, and a subsequent readout of all spin states provides the solution to the given optimisation problem. A fundamental challenge for the construction of a spin-glass annealer is the implementation of tunable all-to-all Ising interactions.

2.1. Parity adiabatic quantum computing

In PAQC, a system with a larger number of spins but only local interactions is used to mimic the spin-glass system with all-to-all Ising interactions at the end of the adiabatic sweep. This is possible because there exists an unambiguous mapping of eigenstates of the spin-glass system to the lowest-energy states of the PAQC system, which are separated by an energy gap from the remaining states of the enlarged Hilbert space of the PAQC system.

The mapping between states of the spin-glass system with all-to-all connectivity to the states of the PAQC system is accomplished as follows. The Ising interaction ${\sigma }_{z,i}{\sigma }_{z,j}$ has two two-fold degenerate eigenvalues: +1 and −1 corresponding to parallel ($| \uparrow \uparrow \rangle $ and $| \downarrow \downarrow \rangle $) and anti-parallel ($| \uparrow \downarrow \rangle $ and $| \downarrow \uparrow \rangle $) alignment of the spins. In the PAQC system, the state of this Ising interaction is represented by a single spin, say state $| \uparrow \rangle $ corresponds to a parallel configuration and $| \downarrow \rangle $ to a anti-parallel configuration of spins. This already provides us with an unambiguous mapping from states of the spin-glass with all-to-all connectivity with N spins to the states of the PAQC system with $K=N(N-1)/2$ spins. However, the larger Hilbert space of the PAQC system already prohibits an unambiguous mapping of all states to corresponding states of the spin-glass Hamilton operator. The effective reduction of the Hilbert space of the PAQC system is accomplished with constraints. Constraints are terms in the Hamilton operator that inflict energy penalties, larger than any other energy scale in the system, on configurations of the PAQC systems that do not have a counterpart in the spin-glass system. The PAQC system is a spatial arrangement of the spins that enables only local constraint, i.e. energy penalties involving only neighbouring spins (c.f. figure 2(b)) and [25]). It consists of a cubic lattice of spins with overall pyramid geometry where the constraints act locally on every unit cell of the cubic lattice. The Hamilton operator of PAQC has the form

Equation (2)

Here, ${\sigma }_{\{x,y,z\}}^{(i)}$ are the Pauli operator spins in the PAQC system. The terms $A(t)$, $B(t)$ and $C(t)$, are three independent schedule functions that define the annealing protocol. $A(t)$, the schedule function of the initial Hamiltonian, is tuned from its maximal value ${A}_{{\rm{\max }}}$ to 0 during the adiabatic sweep. $B(t)$, the schedule function of the Hamiltonian that encodes the optimization problem, is tuned from 0 to its maximal value ${B}_{{\rm{\max }}}$ during the adiabatic sweep. $C(t)$ is the schedule function of the constraint term. $C(t)$ can be switched from 0 to ${C}_{{\rm{\max }}}$, with ${C}_{{\rm{\max }}}$ dominating every other energy scale in the system. However, PAQC also opens the possibility to keep the constraints 'always on'. The constraints are implemented with four-body Ising interactions involving the north $(l,n)$, east $(l,e)$, south $(l,s)$ and west $(l,w)$ spins of every plaquette l. They are constraints because at least at the end of the adiabatic sweep they are the dominating energy contribution and all dynamics is restricted to reside in their ggroundstate manifold. The scheme features error correction capabilities for uncorrelated noise models [32]. However, in [33], it is pointed that this does not correct for correlated errors as expected from a error model based on quantum Monte Carlo simulations, where minor embedding schemes show larger robustness.

In the following, we show how to decompose a general many-body Ising constraints to two-body Ising interactions with the help of ancilla spins.

2.2. Decomposition of constraints

The constraint decomposition follows a two-step process: (i) We present a general recursive technique to decompose many-body Ising constraints to three-body Ising constraints with ancilla spins and use this technique to decompose the four-body Ising constraint of PAQC to two three-body Ising constraints, and (ii) we show a decomposition of a three-body Ising constraint with two-body Ising interactions with one ancilla spin. This constraint decomposition technique follows a similar logic to the embedding described above. However, instead of embedding the whole eigensystem of an operator in another, we only need to embed the ggroundstate manifold of the constraint and make sure that the ground state manifold of the embedding constraint is still separated by a sufficient gap from the rest of the states.

Decomposing many- to three-body constraints. The many-body Ising interaction $C{\prod }_{i=1}^{M}{\sigma }_{z}^{(i)}$, has two multiply degenerate eigenvalues 1 and −1 corresponding to product states of an even or odd number of spins in the spin down state, respectively. In the following, we call them even- or odd-parity states, respectively. Depending on the sign of the constraint strength C, either the states with even or odd parity are the ground state manifold, i.e. are 'allowed' under the constraint. Given a set of qubits in eigenstates of their respective ${\sigma }_{z}$ operators, we could detect the parity of the state by either detecting the parity of the whole set or subdividing the set into two subsets, A (spins 1 through to k) and B (spins $k+1$ through to N), and detecting the parity of the two subsets. If the two subsets have the same parity, the parity of the set is even and vice versa. We can translate this principle into the domain of Ising interactions with the help of an ancilla spin $(a)$,

Equation (3)

The equivalence '$\equiv $' is here defined as the existence of a one-to-one mapping between the ground state manifolds in the sense that the states are identical for the non-ancilla spins. Additionally, all remaining states of the embedding constraint have to be separated by a gap from the ground state manifold. The overall sign of the embedding constraint $\pm | C| $ above is actually arbitrary because the states of the ground state manifolds for the system with ancilla are in both cases identical for the non-ancilla spins. The constraint decomposition technique is valid for odd $(C\gt 0)$ and even $(C\lt 0)$ parity constraints. The four-body Ising constraint of PAQC is a even party constraint, which is why we have chosen to further illustrate even parity decomposition. The reasoning for odd parity constraints parallels the reasoning given in the following.

The ground state manifold of the embedding constraint for an even parity constraint $(C\lt 0)$ is characterised by a odd parity for the subset A including the ancilla spin $(a)$ and the subset B including the same ancilla spin $(a)$. If the ancilla spin $(a)$ is in the spin-up state, then both subsets A and B have to have odd parity in order for the state to be in the ground state manifold. If the ancilla spin $(a)$ is in the spin down state, both subsets have to have even parity. The ancilla spin $(a)$ therefore 'communicates' the parity information between subsets. Therefore, the set of spins not including the ancilla spin has the same ground state manifold as for the many-body Ising interaction. The gap of the embedding constraint above, c.f. Equation (3), is actually of the same size as the gap of the constraints involving the subsets A and B by virtue of the non-perturbative nature of the decomposition. This opens the door to very strong effective Ising constraints.

To summarise, with the ancilla spin $(a)$, one can decompose a constraint of order N, i.e. involving N spins, into two constraints of order $k+1$ and $N-k+1$. This decomposition can be iterated recursively with both subsystems, thus one can decompose any constraint of arbitrary order N into a set of coupled three-body constraints. The recursive sequence of decomposition steps generates a tree structure which is a highly desirable feature for two-dimensional setups like the ones used in circuit quantum electrodynamics because qubits are connected without any crossings and no air-bridges are needed, not including control and readout circuitry.

Figure 3 depicts an example of this recursive algorithm for the decomposition of a five-body constraint. We start by splitting the five-body term into a right branch with subset $\{1,2\}$ and a left branch with subset $\{3,4,5\}$. In each branch, we add the ancilla spin $({a}_{1})$ resulting in two terms, a three-body constraint in the right branch and a four-body constraint in the left branch. The left branch is finished, as three-body constraints cannot be further decomposed with this scheme. We continue with the left branch and split the four-body constraint into two subsystems containing $\{4,5\}$ and $\{{a}_{1},3\}$, respectively. Adding the ancilla qubit $({a}_{2})$ on both sides results in a three-body constraint in all leafs of the tree which terminates the procedure. After joining the leafs, the final decomposed five-body constraint is depicted in figure 3(b).

Decomposing three- to two-body constraints. The second decomposition step (ii) aims at replacing the remaining three-body constraints by pair interactions including another ancilla spin. The above recursive decomposition cannot be applied to three-body terms, as one subbranch would again result in a three-body constraint. However, it is possible to construct a set of interactions that feature the same degenerate ground state as the three-body Ising constraint if we ignore the state of the ancilla spin for the moment [30, 31, 34]. There are actually many possible combinations of pair interactions that share the same degenerate ground state. We chose the particular solution, where (i) all interactions have the same sign and (ii) there are no crossings (depicted in figure 1(b))

Equation (4)

Here, the symbol '$\equiv $' quotes the definition of equivalence for constraints given above. Using equation (4) together with equation (6), the parity quantum annealing scheme [25] can be implemented with pair interactions only. The full layout of interactions and ancillas is shown in figure 1(c).

Figure 1.

Figure 1. (a) Four-body constraints (red) of the parity quantum annealing scheme are decomposed into two three-body terms (blue) with one ancilla qubit (orange). (b) In a second step, the three-body terms are decomposed into pair interactions with an additional ancilla (yellow). The interaction strengths are either $+1{\sigma }_{z}{\sigma }_{z}$ (blue) or $+2{\sigma }_{z}{\sigma }_{z}$ (red). Local fields acting on ancilla qubits and programmable qubits are −2 and −1, respectively. (c) The architecture based on these plaquettes consists of $K=N(N-1)/2$ programmable qubits and $M={(N-2)}^{2}/2$ ancilla qubits.

Standard image High-resolution image
Figure 2.

Figure 2. Illustration of the PAQC encoding technique. (a) The system of spins for the spin-glass annealing paradigm with all-to-all Ising Interactions. Ising pair interactions are symbolised by lines and labelled by the coupled spins. (b) The pyramid shaped, cubic lattice of spins of the PAQC scheme where the parallel or anti-parallel alignment of spins i and j of the spin-glass annealing paradigm is encoded in the state of the spin $(i\,j)$. Four-body Ising interactions act on the unit cells of the PAQC spin system, symbolised by red circles.

Standard image High-resolution image
Figure 3.

Figure 3. (a) Example of the recursive decomposition of five-body Ising constraints into a 'tree' of three-body Ising constraints. The example starts with a plaquette of 5 qubits that interact via the five-body Ising even parity constraint $-C{\sigma }_{z}^{(1)}{\sigma }_{z}^{(2)}{\sigma }_{z}^{(3)}{\sigma }_{z}^{(4)}{\sigma }_{z}^{(5)}$ (top left). Using equation (3), we split the system into two parts $\{1,2\}$ (right branch) and $\{3,4,5\}$ (left branch). Adding the ancilla $({a}_{1})$ results in a three-body Ising constraint and a four-body Ising constraint (bottom middle). The three-body Ising constraint cannot decomposed any further with this technique and is a 'leaf' of the tree. However the four-body Ising constraint is further decomposed into two three-body Ising constraints with another ancilla spin $({a}_{2})$ (top right).

Standard image High-resolution image

2.3. PAQC with pair Ising interactions

Let us now apply the above algorithm to decompose all constraints in the PAQC scheme. The first step (i) is a trivial tree in this case. We simply split the four-body constraint into two three-body constraints,

Equation (5)

Plugging equation (5) into equation (2) results in the adiabatic protocol

Equation (6)

Here, we add one ancilla spin for each plaquette which makes a total of $M={(N-2)}^{2}/2$ spins as shown in figure 1. Each of the ancilla spins is shared by two three-body Ising interactions. Compared with the implementation with four-body Ising constraints, the number of constraints doubled in the above realisation with three-body Ising constraints. In the second step (ii), each three-body term in equation (6) is replaced by the right hand side of equation (4), which results in the final layout with pair interaction only depicted in figure 1(c). Note that this scheme may be optimised by combining pairs of spins (figure 1(c) (yellow)), which results in a layout with less spins but introduces crossings of interactions.

2.4. Adiabatic protocol

The main source of errors in our annealing scheme are Landau-Zener transitions at the minimal gap, as we are considering a coherent annealing sweep. Therefore, it is vital that the constraint decomposition technique presented above does not considerably decrease the minimal gap during the annealing sweep. In the following, we illustrate the differences in the instantaneous eigenenergies of one plaquette for the original PAQC scheme with the four-body Ising constraint (c.f. figure 1(a)) and the above presented decomposition with three ancillas and pair interactions (c.f. figure 1(b)). We start by diagonalizing the instantaneous Hamiltonian of a plaquette with the decomposed four-body Ising constraint. We consider a protocol that starts with a strictly local transverse field Hamiltonian, A = 1, B = 0 and C = 0 in equation (2). Then the local longitudinal fields and constraint terms are adiabatically switched on ($B\to 1$, $C\to 1$ ), while the the ${\sigma }_{x}$ term is switched adiabatically off ($A\to 0$).

In the absence of local programmable fields Ji, the result is a superposition of all eight constraint-satisfying configurations. The time-dependent spectrum of this ideal sweep is shown in figure 4(a). In the presence of local programable fields, the final ground state is the solution of the optimization problem (c.f. figure 4(b)). The local field terms Ji lift the degeneracy of the final state. All unphysical, or not constraint-satisfying states, are separated by a energy gap of size C at the end of the sweep, thereby validating our above decomposition of the four-body Ising constraint.

Figure 4.

Figure 4. (a) Time-dependent spectrum of a four-body term represented by pair interactions (figure 1(b)). In the absence of a programmable local field, all constraint-satisfying states (red) collapse to the degenerate ground state (black). Constraint-violating states (blue) are separated by C. The finial state is the superposition of all constraint-satisfying configurations. (b) Spectrum during the sweep with local random fields with strength $| J| \lt 0.2$. The degeneracy of the ground state is lifted and the solution of the optimization problem is the ground state. Note that the constraint-violating terms are shifted by at least C.

Standard image High-resolution image

Instead of adiabatically switching on the constraints, there is also the possibility of always-on constraints, $C(t)=\mathrm{const}.=1$, which circumvents difficulties in the implementation. Either there is an efficient protocol to generate the ground state of the initial Hamiltonian with A = 1 and C = 1, or one has dominating local transversal fields $A\gg C$ and initialises the system in the product state of all local transverse field Hamiltonians. The instantaneous eigenenergies of the two protocols, the ramping up of constraints and always-on constraints (A = C = 1 at the beginning of the sweep), are illustrated in figure 5 using the same parameters as in figure 4.

Figure 5.

Figure 5. Time-dependent spectrum of the 'always on' switching protocol (black) and a ramp of interactions (red). Parameters as in figure 4 with random local fields of strength $| J| \lt 0.2$ and $C(t)=\mathrm{const}.\,=\,1$.

Standard image High-resolution image
Figure 6.

Figure 6. (a) Scatter plot of the minimal gap in the annealing process for a plaquette with three ancillas ${{\rm{\Delta }}}_{\min }^{a}$ against the minimal gap with an idealised four-body interaction ${{\rm{\Delta }}}_{\min }^{4\mathrm{body}}$ of N = 500 random instances with $| J| \lt 0.5$. (b) Scatter plot comparing the minimal gap in the ramp protocol ${{\rm{\Delta }}}_{\min }^{a}$ against the always-on protocol ${{\rm{\Delta }}}_{\min }^{c}$. The same instances as in (a) are shown.

Standard image High-resolution image

To quantify the efficiency of the protocols, we compare the minimal gaps from the idealised model with four-body Ising interaction and the ancilla-based approach. Figure 6(a) shows a comparison of the minimal gap of a plaquette using N = 500 random instances with $| J| \lt 0.5$. The minimal gap in the proposed three-ancilla setup $({{\rm{\Delta }}}_{\min }^{a})$ is compared with the minimal gap (${{\rm{\Delta }}}_{\min }^{4{\rm{body}}}$) of the model with idealised four-body interaction. The gap in the idealised model is in general larger than in the ancilla-based implementation, but remarkably, for a large number of instances, they are almost identical. We further compared the gap in the ramp protocol (${{\rm{\Delta }}}_{\min }^{a}$) as in figure 6(a) with the gap $({{\rm{\Delta }}}_{\min }^{c})$ in the always-on protocol starting from a superposition state (c.f. figure 6(b)). The gap in the always-on protocol is smaller for most instances, with only a small systematic difference between the two in favour of the ramp protocol.

3. Transmon implementation

We aim to exploit the large coherence times of Transmon qubits [14] for applications in adiabatic quantum computing. With T1 times of up to 60μs [35], Transmons are a suitable candidate for fully coherent quantum annealing protocols. With the aim to implement the PAQC scheme with Transmons, one needs a spin Hamiltonian with (i) individually tunable transverse and (ii) longitudinal fields; and (iii) Ising pair interactions that are large compared with the on-site energies. These terms are not naturally available in Transmons. With a microwave drive, we introduce (i) fully tunable longitudinal fields and (ii) transverse fields in a frame rotating with the microwave drive. (iii) The required Ising pair interaction is implemented based on Josephson ring modulators (JRM) [36], which improves the attainable coupling strengths with respect to a previous proposal [37, 38], making the interaction the dominant energy scale. In the following, we derive (i)–(iii) in detail.

3.1. Transmon

The Transmon was developed on the basis of charge qubits as a improvement on their coherence times. However, while the charge qubit provides a natural mapping to a spin Hamilton operator with longitudinal and transverse fields, the Transmon is more appropriately described as a harmonic oscillator with small non-linearity. In the following, we show how the longer coherence times of the Transmon are a direct consequence of the lost transverse field term and how to combine the transverse field term and the long coherence times in the following subsection.

A charge qubit is an arrangement of two superconducting islands that are connected by a Josephson junction. The state of the Cooper pair condensate on the two islands is completely described by a canonical conjugate pair of operators: the number of Cooper pairs N that have tunnelled through the Josephson junction starting from the electrical neutral state and the phase difference of the Cooper pair condensate between the two islands ϕ. The Hamilton operator of the charge qubit is given by the sum of the electrical energy stored in the capacitor (C) that is formed by the two islands and the inductive energy of the Josephson junction,

Equation (7)

where ${E}_{C}={e}^{2}/2C$ is the charging energy with the elementary charge e, EJ the Josephson energy and ${n}_{g}={C}_{g}{V}_{g}/(2e)$ the Cooper pair equivalent of an external potential Vg with capacitance Cg. Here, ng can be a externally applied classical field or the quantum electrical field of uncontrolled sources as well as quantum fields from coupled resonators or other charge qubits. If we ignore the Josephson energy for the moment, the states of the charge qubit as a function of the externally applied classical electric field ng are parabolas with origin at integer ng and degeneracies for half integer ng. Each parabola is the energy of a state $| n\rangle $ with exactly n Cooper pairs. The degeneracy at half integer ng is lifted by the introduction of the Josephson energy. This is possible because of the tunnelling of Cooper pairs between the islands through the Josephson junction. Around this avoided crossing in the spectrum, one can truncate the Hilbert space of the charge qubit to the two states of exactly defined Cooper pairs and thereby accomplish a mapping to an effective spin Hamilton operator with a longitudinal field given by the externally applied classical electrical field ${h}_{z}=2{E}_{C}(1-2{n}_{g})$ and transverse field given by the Josephson energy ${h}_{x}={E}_{J}/2$. The qubit states corresponding to spin 'up' and 'down' states are states of well-defined Cooper pairs $| n\rangle \to | \uparrow \rangle $ and $| n+1\rangle \to | \downarrow \rangle $ away from the avoided crossing. Exactly in the middle of the avoided crossing, the states are symmetric and anti-symmetric superpositions of states of well-defined charge, $(| n\rangle +| n+1)/\sqrt{2}\to | \uparrow \rangle $ and $(| n\rangle -| n+1)/\sqrt{2}\to | \downarrow \rangle $. Because the charge is not well-defined at the avoided crossing, the charge qubit shows increased resilience to external fluctuations in the electric field. The Transmon improves on this idea by generating a universal avoided crossing for a very strong Josephson tunneling which mixes all charge states $| n\rangle $, ${E}_{J}/{E}_{C}\gg 1$. In this parameter regime, the difference between the maximal and minimal energy as a function of ng decreases exponentially as a function of ${E}_{J}/{E}_{C}$, while the non-linearity, i.e. the difference between the energy ground state to the first excited state ${E}_{1}-{E}_{0}$ and the first to the second excited ${E}_{2}-{E}_{1}$ state decreases polynomially. Here, En are the eigenenergies of the Transmon. The Transmon thereby combines the advantages of a universal sweet spot with sufficient non-linearity, necessary for individual addressing of its eigenstates with microwave drives.

The eigenstates of the Transmon are characterised by small zero-point fluctuations in the phase by virtue of ${E}_{J}\gg {E}_{C}$, which is why it is appropriate to truncate the Josephson energy that is given by the cosine of the phase to fourth order,

Equation (8)

Equation (9)

The first two terms of the truncated Hamilton operator can be identified by the Hamilton operator of a harmonic oscillator. The externally applied ng represents a shift in space for the harmonic oscillator that does not change the eigenenergies which shows the existence of the universal 'sweet spot' [2]. We may therefore as well set ${n}_{g}\to 0$ in the following and introduce lowering and raising operators to describe the number of Cooper pairs and phase of the Transmon,

Equation (10)

Equation (11)

The truncated Hamilton operator of the Transmon can be approximated with the help of a rotating wave approximation that transforms the non-linear fourth-order phase term to a Kerr non-linearity,

Equation (12)

Equation (13)

Here, we shifted the ground state energy to zero. Here, $(\sqrt{8{E}_{J}{E}_{C}}-{E}_{C})$ is the Transmon energy including a renormalisation of the Transmon energy coming from the normal ordering procedure of the fourth-order phase term. The Transmon is therefore more appropriately described as a non-linear harmonic oscillator and as a consequence of the universal 'sweet spot' there is no mapping to a spin Hamilton operator with a transverse field. In the following subsection, we show how to reintroduce a transverse field in a rotating frame by a constant microwave drive.

3.2. Rotating frame

A viable way to reintroduce a transverse field is to excite the Transmon with a constant microwave drive of frequency ${\omega }_{d}$, strength A, and Hamilton operator ${H}_{{\rm{drive}}}=A({{ae}}^{i{\omega }_{d}t}+{a}^{\dagger }{e}^{-i{\omega }_{d}t})$. In a frame rotating with this microwave drive, $U=\exp (-i{\omega }_{d}{{ta}}^{\dagger }a)$, the transverse field term is reintroduced,

Equation (14)

where $2J=\sqrt{8{E}_{C}{E}_{J}}-{\rm{\hslash }}{\omega }_{d}$ is the energy equivalent of the frequency difference between Transmon and microwave drive. Here, ${\rm{\hslash }}$ is the reduced Planck constant. The symbol J is chosen in anticipation for the use in the PAQC scheme as the on-site longitudinal field that encodes the coupling strength of the associated spin-glass annealer. The quantum annealing processor is operated in a regime where we can neglect occupation of states higher than the first excited state, which is ensured by choosing a driving strength smaller than the non-linearity $A\lt {E}_{C}$. Therefore, we may project the Hamilton operator to the qubit subspace to get the effective Hamilton operator in the qubit subspace,

Equation (15)

To summarize, in a frame rotating with the microwave drive, the effective longitudinal field is defined by the energy equivalent of the frequency difference between the Transmon and the microwave drive J, which allows longitudinal fields smaller in magnitude than the Ising pair interaction presented below. Additionally, a transverse field given by the strength of the microwave drive A, is reintroduced.

3.3. Ising pair interaction

Transmon qubits are naturally coupled via their electric field [1]. First, we shortly illustrate why this capacitive coupling is linear in the Transmon field operators while the desired Ising pair interaction is of fourth order. Afterward, we present how to implement the Ising pair interaction with the non-linearity provided by Josephson junctions.

For capacitively coupled Transmons, ng, as given above in equation (8), of one Transmon $(a)$ is a function of the electric field of another Transmon $(b)$ and their mutual capacitance Cc. The interaction term in the Hamilton operator is

Equation (16)

where we assumed the validity of the Transmon approximation and performed a rotating wave approximation valid for small coupling strengths $g\lt {\omega }_{a}+{\omega }_{b}$. Here, ${\omega }_{a}$ and ${\omega }_{b}$ are the frequencies of the two Transmons, while a and b are their field operators. The natural capacitive coupling of two Transmons therefore provides us with an exchange interaction that is linear in the field operators of the two Transmons. However, the fundamental building block of the four-body Ising constraint that is needed for PAQC is, by virtue of the decomposition techniques described above (c.f. figure 1(c)), the Ising pair interaction ${\sigma }_{z}^{(a)}{\sigma }_{z}^{(b)}$. In terms of field operators of the Transmon, a and ${a}^{\dagger }$, every ${\sigma }_{z}$ operator is already a quadratic operator because $2({a}^{\dagger }a-1/2)$ corresponds to ${\sigma }_{z}$ in the qubit subspace and consequentially our desired Ising pair interaction term is of fourth order. Therefore, this interaction has to be implemented with the only non-linear element at our disposal for superconducting qubits, the Josephson junction. One possibility is to connect an island of the first Transmon to one of the islands of the other, which results in a coupling energy given by

Equation (17)

Here, we again assumed the validity of the Transmon approximation and performed a rotating wave approximation to get the linear exchange interaction. Additionally, we truncated the cosine of the Josephson junction at fourth order which is valid given the Transmon approximation. In addition to our desired fourth-order non-linear coupling, we get a linear exchange interaction. It is possible to cancel the linear terms provided by a Josephson junction with a parallel capacitive interaction [37, 38] because ${g}_{{\rm{cap}}}$ and ${g}_{{\rm{ind}}}$ have a different sign. However, this scheme limits the interaction energy to a small fraction of the Transmon non-linearity because the coupling capacitance has to be considerably smaller than the Transmon's capacitance to inhibit unwanted long-range interactions [38].

We present another possibility to couple Transmons with Josephson junctions in the form of a JRM [36] where the linear coupling vanishes for symmetry reasons. The interaction energies exceed the coupling energies attainable by the capacitively shunted Josephson junction scheme by at least an order of magnitude without introducing unwanted long-range interactions.

Four superconducting islands joined to a ring by identical Josephson junctions define a JRM (c.f. appendix A). Its energy is proportional to the product of the cosines of three orthogonal modes if the flux threaded through the JRM loop vanishes. The JRM's modes comprise two differential modes involving opposing superconducting islands ${\varphi }_{x}={\varphi }_{1}-{\varphi }_{3}$, ${\varphi }_{y}={\varphi }_{2}-{\varphi }_{4}$ and a third mode involving all islands ${\varphi }_{z}={\varphi }_{1}-{\varphi }_{2}+{\varphi }_{3}-{\varphi }_{4}$. Here, ${\varphi }_{i}={\int }_{-\infty }^{t}{V}_{i}{dt}$ is the flux variable defined as the time integral of the electrical potential Vi of island i. It is connected to the phase variable introduced above by ${\varphi }_{i}=\phi {\varphi }_{0}$ with the reduced magnetic flux quantum ${\varphi }_{0}={\rm{\hslash }}/(2e)$.

We associate the two differential modes ${\varphi }_{x}$ and ${\varphi }_{y}$ of the JRM with the modes that register the two qubits ${\varphi }_{a}$ and ${\varphi }_{b}$ by connecting them with conducting leads. For the sake of simplicity, we assume all JRM junctions to be equal with Josephson energy ${E}_{{\rm{jrm}}}$, although our setup tolerates the usual fabrication inaccuracies (c.f. appendix B). The qubit capacitances CJ are also assumed to be equal, although unequal qubit capacitances do not alter the main result. The qubit Josephson junctions can be implemented as direct current superconducting quantum interference devices (dc-SQUID) and are modelled as tunable Josephson energies EJa and EJb. Note that the Transmons might as well be implemented with fixed frequencies, provided that we can individually change the drive frequencies. They are the 'nobs' of the quantum annealing processor that need to be adjusted to encode the problem we want to solve. Contrary to the typical use case of the JRM with ${\varphi }_{{ext}}\ne 0$, we are interested in the case ${\varphi }_{{ext}}=0$. The two-Transmon Lagrangian with the corresponding JRM contribution reads (for a full derivation of the Lagrangian of the JRM c.f. 4),

Equation (18)

Equation (19)

Notice that the third mode ${\varphi }_{z}$ of the JRM does not possess any capacitive term. In a mechanical picture, this mode is a massless particle moving in a one-dimensional potential that depends on the position of other massive particles. Therefore, it will immediately adjust itself to the potential minimum. Even a small capacitance for the ${\varphi }_{z}$-mode which will be present in every setup does not change this qualitative picture. In analogy to the elimination of the coupling degree of freedom for the g-mon [39], we find

Equation (20)

Equation (21)

${\varphi }_{a}$ and ${\varphi }_{b}$ are the flux variables of Transmons, i.e. their zero-point fluctuations are small and centred around 0. Therefore, ${\varphi }_{z}=0$ is the only possible solution to the constraint given by the Euler–Lagrange equation for mode ${\varphi }_{z}$. We Legendre transform the resulting Lagrangian and quantise the theory to get the Hamilton operator,

Equation (22)

Equation (23)

where ${N}_{a/b}=C{\dot{\varphi }}_{a/b}/(2e)$ and ${\phi }_{a/b}={\varphi }_{a/b}/{\varphi }_{0}$. In preparation for the Transmon approximation, we introduce bosonic lowering and raising operators with the following dependency on the Cooper pair and phase operators,

Equation (24)

Equation (25)

where $x\in \{a,b\}$. Truncating the Hamilton operator to quartic order results in the canonical Transmon approximation of a harmonic oscillator mode with a quartic non-linearity given by the charging energy of the Transmon,

Equation (26)

with

Equation (27)

Equation (28)

As a last step, we add the individual microwave drives introduced above for both Transmons and transform into a frame rotating with the microwave drives with ${U}_{x}=\exp (-i{\omega }_{d,x}{{tx}}^{\dagger }x)$ for $x\in \{a,b\}$. Note, that the ${\sigma }_{z}$ coupling strength in equation (28) is of the order of the non-linearity of the Transmon. The validity of our projection to the individual qubit subspaces still only requires the driving strengths Aa and Ab to be small compared with the on-site non-linearity EC. The effective Hamilton operator in the rotating frame projected to the qubit subspace is, therefore,

Equation (29)

where ${J}_{x}={E}_{x}-\delta -{\rm{\hslash }}{\omega }_{d,x}$ for $x\in \{a,b\}$. With δ a renormalization of the Transmon qubit energy that originates in the normal ordering process of the qubit non-linearity as well as the non-linear Ising coupling.

With the longitudinal and transverse field terms and the Ising pair interaction, we have all the ingredients to build a Transmon quantum annealer with the PAQC scheme. Errors on the local field terms ${\sigma }_{z}$ and ${\sigma }_{x}$ are highly reduced in the rotating frame due to ns-precision of state-of-the-art microwave drives used in recent experiments [40]. As the optimization problem is encoded in the z-direction of the Hamiltonian, any error in the ${\sigma }_{z}$ terms must be smaller than the range of programmability $C\gg | J| \gg {\delta }_{z},{k}_{B}T$. Errors arise from the second-order processes neglected in the derivation of equation (29) that result in additional ${\sigma }_{x}{\sigma }_{x}$ couplings (c.f. appendix B). Let us also note additional opportunities for implementations with our scheme. The presented approach allows for an implementation of a Transmon quantum annealer with fixed frequency qubits [41] that show increased resilience with respect to environmental noise. By driving each Transmon individually ${\omega }_{d,a}\ne {\omega }_{d,b}$ and upon changing into the rotating frame, one gets an effective time-independent Hamilton operator. One might even conceive of a Transmon quantum annealer where each Transmon is replaced by a capacitor and the JRMs alone provide the inductive energy for the on-site Transmons, as the effective Josephson energy of both Transmons is the sum of the Josephson energy of the Transmon junction and the JRM junction equation (27).

4. Conclusion

With the strong resilience to noise, the Transmon qubit opens the possibility to go beyond temperature driven annealing protocols and to study the influence of coherent versus incoherent processes in an adiabatic sweep. On the implementation level, we showed how to introduce tunable ${\sigma }_{x}$ and ${\sigma }_{z}$ terms in Transmons as well as ${\sigma }_{z}{\sigma }_{z}$-interactions from Josephson ring modulators that can be larger than the local field terms On the encoding level, we demonstrated a recursive constraint decomposition method that allows one to break any combination of classical k-local constraints into coupled three-body constraints with ancillas. Due to the recursive nature of the algorithm, the resulting graph is a binary tree and therefore does not feature any crossings or junctions. The number of ancillas scales linear with $k-3$ and the recursion terminates at three-body terms To further decompose three-body terms to pair interactions, we followed a different strategy where only the ground state is degenerate. The result is a system where (i) constraint-satisfying terms are degenerate, (ii) the constraint-violating terms are higher in energy, (iii) the layout has no crossings and (iv) all interactions have the same sign.

The decomposition in combination with the parity adiabatic quantum computing scheme [25] and the rotating frame Transmon qubit aims at implementing a quantum annealer with full all-to-all from physical two-body interactions where the problem is encoded in local fields alone. In the D-Wave architecture, the problem is encoded in quasi-local interactions and full connectivity is achieved by minor embedding techniques [26, 27]. The overhead in the present scheme is three ancilla qubits per plaquette. The Transmon quantum annealer provides an alternative to the flux qubit annealer with the prospect of large coherence times that could improve the annealing success probability considerable. Quantum annealing in the rotating frame adds additional flexibility which allows for studies of the influence of coherence in quantum annealing. However, the proposed scheme cannot rely on the interplay of thermal and quantum fluctuations and also poses questions on how incoherent processes influence the success probability which we plan to address in future work.

During the writing of this paper, we became aware of related work. In reference [42], the authors present a PAQC encoding that introduces odd instead of even parity constraints that are easier to implement with pair Ising interactions and a single ancilla spin per constraint. This reduces the number of ancilla spins, however the connectivity graph is not flat. In reference [43], a PAQC implementation based on flux qubits is introduced with four ancilla qubits per plaquette. The proposal introduced here realises a full embedding of the Ising interaction, however at the cost of one additional ancilla qubit per plaquette and an additional energy scale that needs to be larger than the constraint energy scale.

Acknowledgements

We acknowledge fruitful discussions with Gerhard Kirchmair. ML acknowledges support from the Lise-Meitner project M 1972-N27. WL was supported by the Austrian Science Fund (FWF): P 25454-N27 and PZ was supported by ERC Synergy Grant UQUAM and SFB FoQuS (FWF Project No. F4016-N23).

Appendix A.: Josephson ring modulator

The Josephson ring modulator (JRM) as applied in the main text as a coupling device is originally employed in the regime ${\varphi }_{\circ }=\pi {\varphi }_{0}/4$ as a three-wave mixing device for readout purposes [36]. Here, we provide a derivation for the main JRM characteristics and evaluate the performance of the JRM with unequal Josephson junctions. The JRM is a square composed of large area Josephson junctions ${E}_{{JRM}}\gg {E}_{C}$ c.f. figure 7. The Lagrangian reads

Equation (A.1)

Equation (A.2)

with ${\varphi }_{0}={\phi }_{0}/(2\pi )={\rm{\hslash }}/(2e)$ the reduced quantum of flux and the flux enclosed by the ring ${\varphi }_{\circ }$. Because of the enclosed flux ${\varphi }_{\circ }$, a ring current is present even in a static- or ground state ${\dot{\tilde{\varphi }}}_{i}=0$ $\forall i$. Therefore, all node fluxes ${\tilde{\varphi }}_{i}$ may be expressed as the sum of a static ${\varphi }_{i}^{{\rm{DC}}}({\varphi }_{\circ })$ and dynamic ${\varphi }_{i}^{{\rm{AC}}}$ part $\tilde{{\varphi }_{i}}={\varphi }_{i}^{{\rm{DC}}}+{\varphi }_{i}^{{\rm{AC}}}$. While the static fluxes are parameters of the circuit, the dynamic node fluxes are the degrees of freedom of the JRM. To reveal the full role of the enclosed flux in the Lagrangian, it is necessary to determine the static node fluxes ${\varphi }_{i}^{{\rm{DC}}}$ as a function of the enclosed flux ${\varphi }_{\circ }$. The correct values for the static fluxes can be determined with the help of the Euler–Lagrange equations

Equation (A.3)

These equations are the Kirchhoff node rules involving only the inductive branches since no constant current can flow through a capacitor. The set of equations reads

Equation (A.4)

Equation (A.5)

Equation (A.6)

Equation (A.7)

This set of equations suggests a solution with equal static flux drops at all Josephson junctions of the JRM ${\varphi }_{1}^{{DC}}-{\varphi }_{2}^{{DC}}={\varphi }_{2}^{{DC}}-{\varphi }_{3}^{{DC}}={\varphi }_{3}^{{DC}}-{\varphi }_{4}^{{DC}}={\varphi }_{4}^{{DC}}-{\varphi }_{1}^{{DC}}-{\varphi }_{\circ }$. Additionally, all flux drops summed up around the loop should equal ${\varphi }_{\circ }$, which implies ${\varphi }_{2}^{{DC}}=(1/4){\varphi }_{\circ }$, ${\varphi }_{3}^{{DC}}=(2/4){\varphi }_{\circ }$ and ${\varphi }_{4}^{{DC}}=(3/4){\varphi }_{\circ }$. This corresponds to a situation of a static ring current of equal strength $I=({E}_{{JRM}}/{\varphi }_{0})\sin ({\varphi }_{\circ }/(4{\varphi }_{0}))$ at each Josephson junction of the JRM. The Lagrangian expressed in terms of dynamic node fluxes ${\varphi }_{i}^{{\rm{AC}}}$ is

Equation (A.8)

with

Equation (A.9)

Equation (A.10)

Here and in the following, we drop the superscript ${\rm{DC}}$ as there is no danger of confusion. In the main text, we apply the same labelling convention. With this shift to the minimum in the inductive potential, we restored the circular symmetry of the JRM and employ this symmetry to reformulate the Lagrangian to

Equation (A.11)

Equation (A.12)

where ${\varphi }_{x}={\varphi }_{1}-{\varphi }_{3}$, ${\varphi }_{y}={\varphi }_{2}-{\varphi }_{4}$ and ${\varphi }_{z}={\varphi }_{1}-{\varphi }_{2}+{\varphi }_{3}-{\varphi }_{4}$. In the main text, we associate modes ${\varphi }_{x}$ and ${\varphi }_{y}$ with Transmon modes to couple them via the JRM. Note the reduction in the degrees of freedom for the JRM $\{{\varphi }_{1},{\varphi }_{2},{\varphi }_{3},{\varphi }_{4}\}\to \{{\varphi }_{x},{\varphi }_{y},{\varphi }_{z}\}$, which stems from the fact that a fourth mode, ${\varphi }_{m}={\varphi }_{1}+{\varphi }_{2}$ $+{\varphi }_{3}+{\varphi }_{4}$, does not couple to the inductive potential given by the Josephon junctions and may therefore be ignored. The flux enclosed by the JRM loop ${\varphi }_{\circ }$ is defined by the externally applied flux up to multiples of the superconducting quantum of flux (${\varphi }_{\circ }=(2\pi {\varphi }_{0})n+{\varphi }_{{ext}}$ for $n\in {\mathbb{Z}}$). This implies that the JRM is likely to minimise its energy for various externally applied fields ${\varphi }_{{ext}}$ by trapping or releasing flux quanta. In total, there are four distinct states of the JRM corresponding to $n=\{0,1,2,3\}+4n$ for $n\in {\mathbb{Z}}$. Stable operation of the JRM is only possible in the state that minimises the total energy. This restricts the tunability of the JRM. However, we are only interested in a regime where ${\varphi }_{{ext}}=0$, and therefore get stable operation for ${\varphi }_{{ext}}={\varphi }_{\circ }=0$.

Figure 7.

Figure 7. Illustration of the circuit diagram for the two-qubit building blocks of the quantum annealing processor. As all interactions are the same, this is the only required building block. Two Transmons with charging energy EC and Josephson energies EJa and EJb, respectively are coupled by a Josephson ring modulator with Josephson energy ${E}_{{\rm{jrm}}}$.

Standard image High-resolution image

Appendix B.: Non-symmetric josephson ring modulator

The Josephson junctions cannot be made identical in a large setup like the quantum annealing chip we propose. Imperfections in the Josephson junctions defining the superconducting interference devices (SQUID) of the Transmons do not have any effects on the workings of the chip, as  the Transmon frequency is defined by the flux threaded through the SQUID and can be adjusted during the operation of the quantum annealer. Therefore, here we investigate implications of imperfect Josephson junctions of the JRM. Let us assume four different Josephson energies ${E}_{J1}$, ${E}_{J2}$, ${E}_{J3}$ and ${E}_{J4}$ with mean value EJRM and variance ${\rm{\Delta }}{E}_{i}$, then the Lagrangian reads

Equation (B.1)

Equation (B.2)

with the static flux jumps ${\delta }_{\mathrm{1,2}}={\varphi }_{1}^{{\rm{DC}}}-{\varphi }_{2}^{{\rm{DC}}}$, ${\delta }_{\mathrm{2,3}}={\varphi }_{2}^{{\rm{DC}}}-{\varphi }_{3}^{{\rm{DC}}}$, ${\delta }_{\mathrm{3,4}}={\varphi }_{3}^{{\rm{DC}}}-{\varphi }_{4}^{{\rm{DC}}}$ and ${\delta }_{\mathrm{4,1}}={\varphi }_{4}^{{\rm{DC}}}-{\varphi }_{1}^{{\rm{DC}}}+{\varphi }_{\circ }$. The static flux jumps can be determined with the help of the steady-state Kirchhoff rules of the JRM,

Equation (B.3)

Equation (B.4)

Equation (B.5)

Equation (B.6)

For the desired operation mode of very small flux enclosed by the JRM loop ${\varphi }_{\circ }/{\varphi }_{0}\ll 1$, we may linearise the above equations and find the solution ${\delta }_{\mathrm{1,2}}=(1/{E}_{J1})({\varphi }_{\circ }/\alpha )$, ${\delta }_{\mathrm{2,3}}=(1/{E}_{J2})({\varphi }_{\circ }/\alpha )$, ${\delta }_{\mathrm{3,4}}=(1/{E}_{J3})({\varphi }_{\circ }/\alpha )$ and ${\delta }_{\mathrm{4,1}}=(1/{E}_{J4})({\varphi }_{\circ }/\alpha )$ with $\alpha =({E}_{J1}^{-1}+{E}_{J2}^{-1}+{E}_{J3}^{-1}+{E}_{J4}^{-1})$. The Lagrangian for the JRM up to second order in the flux enclosed by the loop ${\varphi }_{\circ }/{\varphi }_{0}$ is expressed in terms of the collective modes ${\varphi }_{x}$, ${\varphi }_{y}$ and ${\varphi }_{z}$,

Here, the energies are renormalized slightly due to the non-zero enclosed flux, ${E}_{{Ji}}^{\prime }={E}_{{Ji}}{(1-(1/2)({\varphi }_{\circ }/({E}_{{Ji}}{\varphi }_{0}\alpha ))}^{2})$ and ${\rm{\Delta }}{E}_{i}^{\prime }={\rm{\Delta }}{E}_{i}{(1-(1/2)({\varphi }_{\circ }/({E}_{{Ji}}{\varphi }_{0}\alpha ))}^{2})$. In the main text, we associate modes ${\varphi }_{x}$ and ${\varphi }_{y}$ with the Transmons with qubit capacitance C. In realistic setups, the mode ${\varphi }_{z}$ has only a negligible capacitive shunt. We set the small shunt to zero and eliminate the mode ${\varphi }_{z}$ as a false degree of freedom before quantisation ${\varphi }_{z}\to 0$ (c.f. the main text for details). The last three terms in the above Lagrangian of the non-symmetric JRM are therefore strongly suppressed and can be neglected. The first non-symmetric contribution in the Lagrangian, however, reduces to a ${\sigma }_{x}{\sigma }_{x}$-coupling between the two transmons in the qubit subspace. This term is quadratic in the flux quadratures of the Transmons, while our desired ${\sigma }_{z}{\sigma }_{z}$ coupling term is quartic in the flux quadratures. The inaccuracies in the Josephson energies ${\rm{\Delta }}{E}_{i}$ may therefore not be larger than 10 percent for typical Transmons,

Equation (B.7)

Please wait… references are loading.
10.1088/2058-9565/1/1/015008