This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: [email protected]thanks: Current affiliation: Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

Accurate methods for the analysis of strong-drive effects in parametric gates

Alexandru Petrescu Institut Quantique and Département de Physique, Université de Sherbrooke, Sherbrooke, Québec, J1K 2R1, Canada    Camille Le Calonnec Institut Quantique and Département de Physique, Université de Sherbrooke, Sherbrooke, Québec, J1K 2R1, Canada    Catherine Leroux Institut Quantique and Département de Physique, Université de Sherbrooke, Sherbrooke, Québec, J1K 2R1, Canada    Agustin Di Paolo Institut Quantique and Département de Physique, Université de Sherbrooke, Sherbrooke, Québec, J1K 2R1, Canada    Pranav Mundada Department of Electrical Engineering, Princeton University, Princeton, NJ 08544, USA    Sara Sussman Department of Physics, Princeton University, Princeton, NJ 08544, USA    Andrei Vrajitoarea Department of Electrical Engineering, Princeton University, Princeton, NJ 08544, USA    Andrew A. Houck Department of Electrical Engineering, Princeton University, Princeton, NJ 08544, USA    Alexandre Blais Institut Quantique and Département de Physique, Université de Sherbrooke, Sherbrooke, Québec, J1K 2R1, Canada Canadian Institute for Advanced Research, Toronto, Ontario M5G 1M1, Canada
Abstract

The ability to perform fast, high-fidelity entangling gates is an important requirement for a viable quantum processor. In practice, achieving fast gates often comes with the penalty of strong-drive effects that are not captured by the rotating-wave approximation. These effects can be analyzed in simulations of the gate protocol, but those are computationally costly and often hide the physics at play. Here, we show how to efficiently extract gate parameters by directly solving a Floquet eigenproblem using exact numerics and a perturbative analytical approach. As an example application of this toolkit, we study the space of parametric gates generated between two fixed-frequency transmon qubits connected by a parametrically driven coupler. Our analytical treatment, based on time-dependent Schrieffer-Wolff perturbation theory, yields closed-form expressions for gate frequencies and spurious interactions, and is valid for strong drives. From these calculations, we identify optimal regimes of operation for different types of gates including iiSWAP, controlled-Z, and CNOT. These analytical results are supplemented by numerical Floquet computations from which we directly extract drive-dependent gate parameters. This approach has a considerable computational advantage over full simulations of time evolutions. More generally, our combined analytical and numerical strategy allows us to characterize two-qubit gates involving parametrically driven interactions, and can be applied to gate optimization and cross-talk mitigation such as the cancellation of unwanted ZZZZ interactions in multi-qubit architectures.

I Introduction

With considerable advances in state preparation, gate operation, measurement fidelity, and coherence time, superconducting qubits have become one of the leading platforms for quantum information processing Kjaergaard et al. (2020); Krantz et al. (2019); Blais et al. (2021). Systems consisting of up to a few dozen qubits have been recently deployed by a number of research groups Versluis et al. (2017); Arute et al. (2019); Corcoles et al. (2020). As these architectures are scaled up, an important challenge is to engineer two-qubit interactions to realize gates that are fast enough compared to the decoherence times of the qubits, while at the same time obtaining operation fidelities that are sufficiently high to satisfy a threshold for quantum error correction Gottesman (1997); Aharonov and Ben-Or (2008). To realize fast and high-fidelity two-qubit gates, precise modeling of the dynamics of small multi-qubit systems is necessary, but becomes computationally difficult as the number of degrees of freedom increases. Moreover, to achieve fast gates, drives that are strong in the sense of the rotating-wave approximation (RWA) are necessary, in which case beyond-RWA corrections become important.

Refer to caption
Figure 1: a) Graph representation of the model. The three bare modes have mutual capacitive couplings (light gray edges); mode cc is parametrically driven. b) Superconducting circuit implementation: modes aa and bb are transmon qubits; the coupler mode cc is implemented as a generalized capacitively shunted flux qubit (see text).

A dominant source of infidelity in gate operation consists of cross-Kerr interactions, or the ZZZZ terms in Pauli matrix notation. These terms are either static due to the connectivity of qubits, or dynamically generated by control drives. In the case of many two- and single-qubit gates, ZZZZ terms produce spurious entanglement that cannot be mitigated by local single-qubit operations. There are active experimental efforts to reduce the effect of ZZZZ interactions Mundada et al. (2019); Ku et al. (2020); Sung et al. (2020); Noguchi et al. (2020); Kandala et al. (2020); Zhao et al. (2020a). Moreover, the presence of nonlocal ZZZZ interactions, and of higher-order cross-Kerr terms, can indicate the onset of quantum chaotic behavior in systems of many coupled qubits Berke et al. (2020).

In this paper, we present a computationally efficient set of analytical and numerical tools to characterize and tailor gate Hamiltonians. As an example application of these tools, we consider flux-tunable parametric coupler architectures McKay et al. (2016); Collodo et al. (2020) schematically illustrated in Fig. 1. We develop two complementary approaches, both of which start from a treatment of the Floquet Hamiltonian which can capture non-RWA effects exactly Shirley (1965); Sambe (1973); Grifoni and Hänggi (1998). Our analytical approach starts from the quantization of the driven superconducting circuit. More specifically, while we adopt a normal-mode picture such as in black-box quantization Nigg et al. (2012) or energy-participation-ratio approaches Minev et al. (2020), the mode frequencies and impedances, and as a result self- and cross-Kerr interactions, depend on the strength of the drive. This dependence is accounted for in an expansion over the harmonics of the drive. From this, we obtain accurate expressions of ac-Stark shifted transition frequencies and interaction strengths. Importantly, as compared to previous work, normal modes are defined here by taking into account drive-induced corrections Verney et al. (2019) to the Josephson potential Petrescu et al. . Due to its similarity to black-box quantization, this analytical technique can be easily generalized to circuits containing multiple qubits and couplers.

To obtain corrections to the effective interaction strengths, our approach relies on a time-dependent Schrieffer-Wolff perturbation theory Theis and Wilhelm (2017); Petrescu et al. (2020); Malekakhlagh et al. (2020a), which consists of a hierarchy of unitary transformations applied to the time-dependent Floquet Hamiltonian Malekakhlagh et al. (2020b); Petrescu et al. (2020). We make the explicit choice to work in the transmon limit of small anharmonicity Koch et al. (2007), expressed in terms of the small dimensionless parameter 8EC/EJ\sqrt{8E_{\text{C}}/E_{\text{J}}}, whereas drive effects are included in a series expansion over the harmonics of the drive frequency and then integrated into the exact treatment of the normal-mode Hamiltonian. This approach allows us to identify the contribution of each driven normal mode to the different effective interaction constants.

Our formalism is equally applicable to strong anharmonicities, where one has to formulate the Hamiltonian in the energy eigenbasis. The cross-resonance gate Rigetti and Devoret (2010); Chow et al. (2011) has been accurately modeled Malekakhlagh et al. (2020a) with such methods, with the notable difference that drive effects were included in the perturbative expansion, something which requires the calculation of higher-order corrections as the drive strength is increased. In contrast, here we show that by effectively performing a series resummation over drive-amplitude contributions, we can model effects such as gate-rate saturation with drive power that are frequently observed (see e.g. Refs. Chow et al. (2011); Sheldon et al. (2016)) without the need to evaluate high-order terms in perturbation theory.

On the other hand, with our numerical approach, we show how gate parameters and, more precisely, the data from a two-tone spectroscopy experiment, can be extracted from a solution of the Floquet eigenproblem Shirley (1965); Sambe (1973). This is efficient by comparison to the simulation of Hamiltonian dynamics over the full duration of the gate protocol: Floquet methods rely on integrating the dynamics over one period of the parametric drive, on the order of 1ns1~{}\text{ns}, which is typically three orders of magnitude shorter than the gate duration. By construction, the parameters extracted from this approach account for renormalization by the drives. We are then able to benchmark the convergence of the analytical approach by direct comparison to the numerical result. In the context of superconducting circuit architectures, Floquet numerical methods have also been used to model instabilities in transmon qubits under strong drives Verney et al. (2019), to obtain corrections beyond linear-response theory for the bilinear interaction between two cavities mediated by a driven ancilla Zhang et al. (2019), to accurately model a strongly-driven controlled-phase gate between transmon qubits Krinner et al. (2020a), or to enhance the coherence of fluxonium qubits Mundada et al. (2020); Huang et al. (2021).

The remainder of this paper is structured as follows. In Sec. II we introduce the circuit model, as well as a pedagogical toy model from which all qualitative features of the full theory can be extracted, and illustrate how to obtain the different gate Hamiltonians. In Sec. III we introduce the basic concepts for second-order RWA, based on a Schrieffer-Wolff transformation of the Floquet Hamiltonian. Section IV captures in more detail the complexity of the problem with an analysis of the three-mode theory derived from the full-circuit Hamiltonian. In Sec. V, we describe in detail a method to extract effective gate Hamiltonians from a Floquet analysis. In Sec. V.2, we compare all previous approaches using simulations based on the numerical integration of the Schrödinger equation. Finally, we summarize in Acknowledgments.

II Model Hamiltonian

As a concrete example of our approach, we consider a model for a parametric coupler consisting of three nonlinear bosonic modes interacting capacitively McKay et al. (2016), see Fig. 1a). The qubit modes a^\hat{a} and b^\hat{b} are assumed to be far-detuned, making the beam-splitter (or iiSWAP) qubit-qubit interactions negligible in the rotating-wave approximation. Those modes are capacitively coupled to a third mode, the coupler c^\hat{c}. The latter can be parametrically modulated in order to activate interactions between the two qubit modes, for example a iiSWAP-type gate on which we mostly focus here.

II.1 Superconducting circuit

A possible realization of this three-mode system is shown in Fig. 1b) and consists of two fixed-frequency transmon qubits interacting via a capacitively shunted flux qubit whose two branches contain one and NN Josephson junctions, respectively You et al. (2007); Steffen et al. (2010); Yan et al. (2016). In a single-mode approximation, this generalized flux qubit plays the role of coupler mode and the parametric drive is realized by modulating the reduced external flux φext=2πΦext/Φ0\varphi_{\text{ext}}=2\pi\Phi_{\mathrm{ext}}/\Phi_{0}, with Φext\Phi_{\mathrm{ext}} the flux threading the coupler’s loop and Φ0\Phi_{0} the flux quantum. For certain values of the static external flux, the coupler has a positive anharmonicity, which is important in obtaining gates with a vanishing ZZZZ interaction Ku et al. (2020); Zhao et al. (2020a); Xu and Ansari (2020); Zhao et al. (2020b). We stress that we use this specific circuit implementation for illustration purposes only, and that the methods presented here apply beyond the weakly-anharmonic regime.

Quantizing the circuit of Fig. 1c) using the standard approach Reynaud et al. (1997); Vool and Devoret (2017) yields the Hamiltonian (see Appendix B for a detailed derivation)

H^(t)=H^a+H^b+H^c(t)+H^g,\displaystyle\hat{H}(t)=\hat{H}_{a}+\hat{H}_{b}+\hat{H}_{c}(t)+\hat{H}_{g}, (1)

where the transmons and the coupler are described by

H^j=4ECj𝒏^j2EJjcos(𝝋^j),j=a,b,H^c(t)=4ECc𝒏^c2αEJccos[𝝋^c+μαφext(t)]βNEJccos[𝝋^cN+μβφext(t)].\displaystyle\begin{split}\hat{H}_{j}=4E_{\text{C}j}\hat{\bm{n}}_{j}^{2}&-E_{\text{J}j}\cos(\hat{\bm{\varphi}}_{j}),\;j=a,b,\\ \hat{H}_{c}(t)=4E_{\text{C}c}\hat{\bm{n}}_{c}^{2}&-\alpha E_{\text{J}c}\cos\left[\hat{\bm{\varphi}}_{c}+\mu_{\alpha}\varphi_{\text{ext}}(t)\right]\\ &-\beta NE_{\text{J}c}\cos\left[\frac{\hat{\bm{\varphi}}_{c}}{N}+\mu_{\beta}\varphi_{\text{ext}}(t)\right].\end{split} (2)

These expressions use pairs of canonically conjugate superconducting phase difference and Cooper pair number for the bare modes, [𝝋^j,𝒏^k]=iδjk\left[\hat{\bm{\varphi}}_{j},\hat{\bm{n}}_{k}\right]=i\delta_{jk} for the mode indices j,k=a,b,j,k=a,b, or cc, and we set =1\hbar=1. The Josephson energies are denoted EJaE_{\text{J}a}, EJbE_{\text{J}b} for the transmon modes, whereas βEJc\beta E_{\text{J}c} is the Josephson energy of one of NN array junctions in the coupler, and α\alpha is a factor parametrizing the anisotropy between the two branches. The parameter β\beta is a renormalization of the superinductance due to disorder in the junction array and finite zero-point fluctuations (see Appendix B). Moreover, the parameter α\alpha accounts for a renormalization of the small junction energy due to hybridization with the modes in the junction array. Furthermore, ECaE_{\text{C}a}, ECbE_{\text{C}b} and ECcE_{\text{C}c} are charging energies. In the transmon regime, EJa/ECaE_{\text{J}a}/E_{\text{C}a} and EJb/ECb50E_{\text{J}b}/E_{\text{C}b}\gtrsim 50 Koch et al. (2007).

The coupler loop is threaded by an external flux φext\varphi_{\text{ext}} which can be modulated in time with a modulation amplitude δφ\delta\varphi, taken to be small compared to the flux quantum

φext(t)=φ¯ext+δφsin(ωdt).\displaystyle\varphi_{\text{ext}}(t)=\overline{\varphi}_{\text{ext}}+\delta\varphi\sin(\omega_{\text{d}}t). (3)

As discussed by You et al. (2019), quantization of the coupler loop under time-dependent flux imposes that the external flux be included in both branches of the potential energy in H^c(t)\hat{H}_{c}(t), with weighting factors μα,β\mu_{\alpha,\beta} determined by the capacitive energies of the two branches (see Appendix B for a detailed derivation). This subtlety is important, as the details of the flux modulation determine the parametric interactions between the two qubit modes.

Finally, the three bare modes interact through linear terms induced by the capacitive coupling

H^g=4ECab𝒏^a𝒏^b+4ECbc𝒏^b𝒏^c+4ECca𝒏^c𝒏^a.\displaystyle\hat{H}_{g}=4E_{\text{C}ab}\hat{\bm{n}}_{a}\hat{\bm{n}}_{b}+4E_{\text{C}bc}\hat{\bm{n}}_{b}\hat{\bm{n}}_{c}+4E_{\text{C}ca}\hat{\bm{n}}_{c}\hat{\bm{n}}_{a}. (4)

The introduction of normal modes will eliminate this linear coupling Hamiltonian.

II.2 Toy model for circuit Hamiltonian

In this subsection, we introduce a simple model which captures the essential qualitative features of the Hamiltonian of Eq. 1. Our toy model consists of three linearly coupled Kerr-nonlinear oscillators and has the form given in Eq. 1 now with

H^a=𝝎a𝒂^𝒂^+𝜶a2𝒂^2𝒂^2,H^b=𝝎b𝒃^𝒃^+𝜶b2𝒃^2𝒃^2,H^c(t)=𝝎c(t)𝒄^𝒄^+𝜶c2𝒄^2𝒄^2.H^g=𝒈ab𝒂^𝒃^𝒈bc𝒃^𝒄^𝒈ca𝒄^𝒂^+H.c.\displaystyle\begin{split}\hat{H}_{a}&=\bm{\omega}_{a}\hat{\bm{a}}^{\dagger}\hat{\bm{a}}+\frac{\bm{\alpha}_{a}}{2}\hat{\bm{a}}^{\dagger 2}\hat{\bm{a}}^{2},\\ \hat{H}_{b}&=\bm{\omega}_{b}\hat{\bm{b}}^{\dagger}\hat{\bm{b}}+\frac{\bm{\alpha}_{b}}{2}\hat{\bm{b}}^{\dagger 2}\hat{\bm{b}}^{2},\\ \hat{H}_{c}(t)&=\bm{\omega}_{c}(t)\hat{\bm{c}}^{\dagger}\hat{\bm{c}}+\frac{\bm{\alpha}_{c}}{2}\hat{\bm{c}}^{\dagger 2}\hat{\bm{c}}^{2}.\\ \hat{H}_{g}&=-\bm{g}_{ab}\hat{\bm{a}}^{\dagger}\hat{\bm{b}}-\bm{g}_{bc}\hat{\bm{b}}^{\dagger}\hat{\bm{c}}-\bm{g}_{ca}\hat{\bm{c}}^{\dagger}\hat{\bm{a}}+\text{H.c.}\end{split} (5)

Comparing to the full-circuit model, note that 𝝎a(b)8EJa(b)ECa(b)ECa(b)\bm{\omega}_{a(b)}\approx\sqrt{8E_{\text{J}a(b)}E_{\text{C}a(b)}}-E_{\text{C}a(b)} whereas the anharmonicities of the transmon qubits are negative and amount to 𝜶a(b)ECa(b)\bm{\alpha}_{a(b)}\approx-E_{\text{C}a(b)}. In an experimental implementation, the parameters defining the coupler—the anharmonicity 𝜶c\bm{\alpha}_{c} and the frequency 𝝎c(t)\bm{\omega}_{c}(t)—can be varied by applying a time-dependent external flux to activate a chosen gate.

The parametric drive resulting from the flux modulation of Eq. 3 is modeled by a modulation of the coupler frequency at a frequency ωd\omega_{\text{d}}

𝝎c(t)=𝝎c+δsin(ωdt).\displaystyle\bm{\omega}_{c}(t)=\bm{\omega}_{c}+\delta\sin(\omega_{\text{d}}t). (6)

In a more detailed analysis of the coupler (see Sec. IV), we take into account the time dependence of the anharmonicity 𝜶c\bm{\alpha}_{c}, but we choose to neglect it in this toy model.

Note that we have reduced the complexity of the problem in a few ways: We have truncated the Josephson expansion to include only quartic terms. All photon number nonconserving terms have been dropped. Higher harmonics of the drive of Eq. 6 are neglected, and we have not considered the ac-Stark shifts of the various coupling constants. All of these refinements are taken into account in the analysis of the full circuit Hamiltonian in Sec. IV. Thus the toy model is significantly simpler than the full circuit theory, but nonetheless still contains the necessary ingredients that allow us to illustrate the general method introduced in this paper.

III Perturbative expansion

In this section, we introduce a perturbative expansion to obtain successive corrections to the effective Hamiltonian in the rotating-wave approximation. To simplify the discussion, we focus on the toy model and come back to the full circuit Hamiltonian in the next section. Our approach relies on a sequence of unitary transformations amounting to a time-dependent Schrieffer-Wolff treatment of the Floquet Hamiltonian in the normal-mode representation, an approach used before to derive corrections to the lifetime of driven transmon qubits Petrescu et al. (2020); Malekakhlagh et al. (2020b). Time-dependent extensions of Schrieffer-Wolff transformations have been shown to be necessary to capture effects of drives in the dispersive regime of circuit QED Theis and Wilhelm (2017), with quantitative agreement with experiment in the analysis of the cross-resonance gate Malekakhlagh et al. (2020a). A notable difference from prior work on microwave-activated two-qubit gates is that, in performing a normal-mode transformation, we are able to obtain good agreement with exact numerics already at second order in perturbation theory. For example, the calculation in Ref. Malekakhlagh et al. (2020a) relies on an expansion in capacitive couplings and drive power, which would require us, in the setup presented here, to go to higher order (fourth) in the calculation to obtain results comparable to the normal-mode approach.

III.1 Formalism

As usual, our starting point is a decomposition of the system Hamiltonian into an unperturbed, exactly solvable part, and a perturbation:

H^=H^(0)(t)+λH^(1)(t).\displaystyle\hat{H}=\hat{H}^{(0)}(t)+\lambda\hat{H}^{(1)}(t). (7)

Here, we have introduced the dimensionless power-counting parameter λ\lambda to keep track of the order in perturbation theory, to be set at the end of the calculation to unity, λ1\lambda\to 1. Now we move to the interaction picture with respect to H^(0)\hat{H}^{(0)}. Letting U^0(t)=𝒯ei0t𝑑tH^(0)(t)\hat{U}_{0}(t)=\mathcal{T}e^{-i\int_{0}^{t}dt^{\prime}\hat{H}^{(0)}(t^{\prime})}, where 𝒯\mathcal{T} is the time-ordering operator, we find for the interaction-picture Floquet Hamiltonian

λH^I(1)(t)it=U^0(t)[H^(0)+λH^(1)(t)it]U^0(t)=U^0(t)λH^(1)(t)U^0(t)it.\displaystyle\begin{split}\lambda\hat{H}_{I}^{(1)}(t)-i\partial_{t}&=\hat{U}_{0}^{\dagger}(t)\left[\hat{H}^{(0)}+\lambda\hat{H}^{(1)}(t)-i\partial_{t}\right]\hat{U}_{0}(t)\\ &=\hat{U}_{0}^{\dagger}(t)\;\lambda\hat{H}^{(1)}(t)\;\hat{U}_{0}(t)-i\partial_{t}.\end{split} (8)

In the above we only assume that the unperturbed time-evolution operator U^0(t)\hat{U}_{0}(t) is known. Equation 8 can be seen as a unitary transformation between two Floquet Hamiltonians Sambe (1973). Thus, the Floquet quasienergies corresponding to λH^I(1)(t)it\lambda\hat{H}_{I}^{(1)}(t)-i\partial_{t} must be identical to those of H^(0)+λH^(1)(t)it\hat{H}^{(0)}+\lambda\hat{H}^{(1)}(t)-i\partial_{t}, while the eigenstates are related by U^0(t)\hat{U}_{0}(t).

In an iterative Schrieffer-Wolff approach, we treat the operator λH^(1)(t)\lambda\hat{H}^{(1)}{(t)} as a small perturbation from which we derive corrections to the known Floquet quasienergies of H^(0)\hat{H}^{(0)} Malekakhlagh et al. (2020b); Petrescu et al. (2020). To this end, we consider a unitary transformation on the interaction-picture Floquet Hamiltonian H^I(t)itλH^I(1)(t)it\hat{H}_{I}(t)-i\partial_{t}\equiv\lambda\hat{H}_{I}^{(1)}(t)-i\partial_{t}, and the corresponding Baker-Campbell-Hausdorff (BCH) expansion in powers of the generator of this unitary, that is

H^I,effiteG^I(t)[H^I(t)it]eG^I(t)=H^IiG^˙I+[H^I,G^I]i2[G^˙I,G^I]it+\displaystyle\begin{split}\hat{H}_{I,\textit{eff}}-i\partial_{t}&\equiv e^{-\hat{G}_{\text{I}}(t)}[\hat{H}_{I}(t)-i\partial_{t}]e^{\hat{G}_{\text{I}}(t)}\\ &=\hat{H}_{I}-i\dot{\hat{G}}_{\text{I}}+[\hat{H}_{I},\hat{G}_{\text{I}}]-\frac{i}{2}[\dot{\hat{G}}_{\text{I}},\hat{G}_{\text{I}}]-i\partial_{t}+...\end{split} (9)

This equation defines the effective Hamiltonian, whose spectrum is equal (up to a desired precision in λ\lambda) to that of the original driven theory. The generator G^I(t)\hat{G}_{\text{I}}(t) can be solved for iteratively in powers of λ\lambda (see Appendix A), which allows us to perform the rotating-wave approximation order by order

H^I,eff=λH^¯I(1)+λ2H^¯I(2)+\displaystyle\hat{H}_{I,\textit{eff}}=\lambda\overline{\hat{H}}_{I}^{(1)}+\lambda^{2}\overline{\hat{H}}_{I}^{(2)}+\ldots (10)

where the terms on the right-hand side are defined below.

To obtain a lowest-order term of the effective Hamiltonian, λH^¯I(1)\lambda\overline{\hat{H}}_{I}^{(1)}, we separate the interaction picture Hamiltonian into oscillatory and non-oscillatory terms with the notation

λH^I(1)(t)λH^¯I(1)+λH^~I(1)(t),\displaystyle\lambda\hat{H}_{I}^{(1)}(t)\equiv\lambda\overline{\hat{H}}_{I}^{(1)}+\lambda\widetilde{\hat{H}}_{I}^{(1)}(t), (11)

where we define the constant part of a time-dependent operator O^(t)\hat{O}(t) by Mirrahimi and Rouchon (2015)

O^¯limT1T0T𝑑tO^(t),\displaystyle\overline{\hat{O}}\equiv\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}dt\hat{O}(t), (12)

whereas the oscillatory part of the operator is

O^~(t)O^(t)O^¯.\displaystyle\widetilde{\hat{O}}(t)\equiv\hat{O}(t)-\overline{\hat{O}}. (13)

Since the time-averaging operation removes all terms that are oscillatory in time, H^¯I(1)\overline{\hat{H}}_{I}^{(1)} is the first-order RWA Hamiltonian Mirrahimi and Rouchon (2015), whereas H^~I(1)\widetilde{\hat{H}}_{I}^{(1)} is canceled by an appropriate choice of the corresponding term at order λ\lambda in the generator.

One can iterate this procedure at every order, collecting terms that are oscillatory and then canceling them. The second-order RWA Hamiltonian (for a derivation, see Appendix A) reads

λ2H^¯I(2)=1i[H^¯I(1),0tλH^~I(1)(t)𝑑t]¯+12i[λH^~I(1)(t),0tλH^~I(1)(t)𝑑t]¯.\displaystyle\begin{split}\lambda^{2}\overline{\hat{H}}_{I}^{(2)}=&\frac{1}{i}\overline{\left[\overline{\hat{H}}_{I}^{(1)},\int_{0}^{t}\lambda\widetilde{\hat{H}}_{I}^{(1)}(t^{\prime})dt^{\prime}\right]}\\ &+\frac{1}{2i}\overline{\left[\lambda\widetilde{\hat{H}}_{I}^{(1)}(t),\int_{0}^{t}\lambda\widetilde{\hat{H}}_{I}^{(1)}(t^{\prime})dt^{\prime}\right]}.\end{split} (14)

This form becomes analogous to the second term in the Magnus expansion Goldman and Dalibard (2014); Mirrahimi and Rouchon (2015) when the perturbation has a vanishing mean, i.e. H^¯I(1)=0\overline{\hat{H}}_{I}^{(1)}=0.

III.2 Black-box quantization approach to the toy model

Expressing the toy-model Hamiltonian as the sum of static quadratic terms, H^(0)\hat{H}^{(0)}, and of time-dependent and Kerr terms, H^(1)(t)\hat{H}^{(1)}(t), the first step in deriving parametrically activated interactions between the transmon modes is to diagonalize the former, which we write as

H^(0)=(𝒂^𝒃^𝒄^)(𝝎a𝒈ab𝒈ca𝒈ab𝝎b𝒈bc𝒈ca𝒈bc𝝎c)(𝒂^𝒃^𝒄^).\displaystyle\hat{H}^{(0)}=\left(\begin{array}[]{ccc}\hat{\bm{a}}^{\dagger}&\hat{\bm{b}}^{\dagger}&\hat{\bm{c}}^{\dagger}\end{array}\right)\left(\begin{array}[]{ccc}\bm{\omega}_{a}&\bm{g}_{ab}&\bm{g}_{ca}\\ \bm{g}_{ab}&\bm{\omega}_{b}&\bm{g}_{bc}\\ \bm{g}_{ca}&\bm{g}_{bc}&\bm{\omega}_{c}\end{array}\right)\left(\begin{array}[]{c}\hat{\bm{a}}\\ \hat{\bm{b}}\\ \hat{\bm{c}}\end{array}\right). (22)

This diagonalization is achieved with an orthonormal transformation 𝜶^=β=a,b,cuαββ^\hat{\bm{\alpha}}=\sum_{\beta=a,b,c}u_{\alpha\beta}\hat{\beta}, for α=a,b,c\alpha=a,b,c, and which is chosen such that H^(0)\hat{H}^{(0)} takes the form

H^(0)=ωaa^a^+ωbb^b^+ωcc^c^,\displaystyle\hat{H}^{(0)}=\omega_{a}\hat{a}^{\dagger}\hat{a}+\omega_{b}\hat{b}^{\dagger}\hat{b}+\omega_{c}\hat{c}^{\dagger}\hat{c}, (23)

where a^\hat{a}, b^\hat{b} and c^\hat{c} are the normal modes and ωa,b,c\omega_{a,b,c} the corresponding mode frequencies. The uαβu_{\alpha\beta} are hybridization coefficients encoding the connectivity of the three modes through the capacitive couplings H^g\hat{H}_{g} entering in H^(0)\hat{H}^{(0)}. In this normal-mode basis, the remainder of the Hamiltonian reads

λH^(1)(t)=j=a,b,c𝜶j2(ujaa^+ujbb^+ujcc^)2(ujaa^+ujbb^+ujcc^)2+δsin(ωdt)(ucaa^+ucbb^+uccc^)(ucaa^+ucbb^+uccc^).\displaystyle\begin{split}&\lambda\hat{H}^{(1)}(t)=\\ &\sum_{j=a,b,c}\frac{\bm{\alpha}_{j}}{2}(u_{ja}\hat{a}+u_{jb}\hat{b}+u_{jc}\hat{c})^{\dagger 2}(u_{ja}\hat{a}+u_{jb}\hat{b}+u_{jc}\hat{c})^{2}\\ &+\delta\sin(\omega_{\text{d}}t)(u_{ca}\hat{a}+u_{cb}\hat{b}+u_{cc}\hat{c})^{\dagger}(u_{ca}\hat{a}+u_{cb}\hat{b}+u_{cc}\hat{c}).\end{split} (24)

The expression above illustrates that coupling between the normal modes arises from the nonlinearity and the parametric drive.

Refer to caption
Figure 2: a) Static cross-Kerr interaction χab(𝝎c)\chi_{ab}(\bm{\omega}_{c}), from first (black) and second-order RWA (blue), and from the full diagonalization of Sec. V) (light blue points) for: 𝝎a/2π=4.0\bm{\omega}_{a}/2\pi=4.0, 𝝎b/2π=5.5\bm{\omega}_{b}/2\pi=5.5, 𝜶a/2π=0.3\bm{\alpha}_{a}/2\pi=-0.3, 𝜶b/2π=0.2\bm{\alpha}_{b}/2\pi=-0.2, 𝜶c/2π=0.25\bm{\alpha}_{c}/2\pi=0.25, 𝒈ab/2π=0.12\bm{g}_{ab}/2\pi=0.12, 𝒈bc/2π=0.12\bm{g}_{bc}/2\pi=-0.12, all in GHz, and 𝒈ca/2π=0\bm{g}_{ca}/2\pi=0. b) Analogue of a) for dynamical cross-Kerr interaction at δ/2π=0.3GHz\delta/2\pi=0.3\,\text{GHz}. c) Same as b) for the gate interaction rate Jab(𝝎c)J_{ab}(\bm{\omega}_{c}). Inset: Jab(δ)J_{ab}(\delta) at 𝝎c/2π=4.25\bm{\omega}_{c}/2\pi=4.25 GHz.

Our choice of unperturbed Hamiltonian H^(0)\hat{H}^{(0)} and perturbation λH^(1)\lambda\hat{H}^{(1)} in Eqs. 23 and LABEL:Eq:H1Toy, respectively, is guided by black-box quantization Nigg et al. (2012): the unperturbed Hamiltonian is linear and diagonal in the normal-mode basis, whereas the perturbation consists of Kerr-nonlinear terms, on the one hand, and quadratics appearing from the normal-mode expansion of the parametric drive, on the other hand. As we show below, while better choices are possible (see Sec. III.3), this choice leads to a simple and intuitive form for the effective Hamiltonian.

As an example of the many common types of interactions that can be activated by a parametric drive Blais et al. (2021), an iiSWAP interaction between the transmon modes arises if we set the modulation to be at the frequency difference between the two transmon modes

ωd=ωbωa.\displaystyle\omega_{\text{d}}=\omega_{b}-\omega_{a}. (25)

This choice yields the first-order RWA Hamiltonian

λH^¯I(1)=Jab(1)(ia^b^+H.c.)+αa(1)2a^2a^2+αb(1)2b^2b^2+αc(1)2c^2c^2+χab(1)a^a^b^b^+χbc(1)b^b^c^c^+χca(1)c^c^a^a^.\displaystyle\begin{split}\lambda\overline{\hat{H}}_{I}^{(1)}&=J^{(1)}_{ab}\left(-i\hat{a}^{\dagger}\hat{b}+\text{H.c.}\right)\\ &+\frac{\alpha_{a}^{(1)}}{2}\hat{a}^{\dagger 2}\hat{a}^{2}+\frac{\alpha_{b}^{(1)}}{2}\hat{b}^{\dagger 2}\hat{b}^{2}+\frac{\alpha_{c}^{(1)}}{2}\hat{c}^{\dagger 2}\hat{c}^{2}\\ &+\chi_{ab}^{(1)}\hat{a}^{\dagger}\hat{a}\hat{b}^{\dagger}\hat{b}+\chi_{bc}^{(1)}\hat{b}^{\dagger}\hat{b}\hat{c}^{\dagger}\hat{c}+\chi_{ca}^{(1)}\hat{c}^{\dagger}\hat{c}\hat{a}^{\dagger}\hat{a}.\end{split} (26)

The first row of this equation contains the iiSWAP interaction of amplitude Jab(1)J^{(1)}_{ab}. The second row contains the mode anharmonicities, and the third row contains cross-Kerr interactions, the first of which is the ZZZZ term.

The coupling constants in the above effective Hamiltonian result from the normal-mode transformation of the quadratic part of the toy model and take the form

Jab(1)=ucaucbδ2,αj(1)=i=a,b,cuij4𝜶i,χjk(1)=i=a,b,c2uij2uik2𝜶i,\displaystyle\begin{split}J_{ab}^{(1)}=&u_{ca}u_{cb}\frac{\delta}{2},\quad\alpha_{j}^{(1)}=\sum_{i=a,b,c}u_{ij}^{4}\bm{\alpha}_{i},\\ &\chi^{(1)}_{jk}=\sum_{i=a,b,c}2u_{ij}^{2}u_{ik}^{2}\bm{\alpha}_{i},\end{split} (27)

for all j,k=a,b,cj,k=a,b,c, and jkj\neq k. In practice, one wants to maximize Jab(1)J_{ab}^{(1)} to obtain a fast gate, while minimizing the cross-Kerr interactions χjk(1)\chi^{(1)}_{jk} to avoid the accumulation of coherent errors. Cross-Kerr interactions are a source of infidelity for a iiSWAP-type gate, as well as in other gate implementations Ku et al. (2020); Sung et al. (2020); Noguchi et al. (2020); Kandala et al. (2020); Zhao et al. (2020a); Krinner et al. (2020b). In the first-order RWA Hamiltonian, to cancel the cross-Kerr interaction between the two transmons, we use a coupler with a positive anharmonicity Zhao et al. (2020b) 𝜶c>0\bm{\alpha}_{c}>0, together with 𝜶a,𝜶b<0\bm{\alpha}_{a},\bm{\alpha}_{b}<0, which is distinct from using qubits of opposite anharmonicities Ku et al. (2020); Xu and Ansari (2020), or couplers with negative anharmonicity Zhao et al. (2020a); Sete et al. (2021). Equation 27 forms the basis for the optimization of the gate parameters. Before pursuing this further, we first derive important corrections to the gate Hamiltonian from the oscillatory part of the Hamiltonian, H^~I(1)(t)\widetilde{\hat{H}}_{I}^{(1)}(t). Finally, note that the first-order term χjk(1)\chi^{(1)}_{jk} is only a static, i.e. δ\delta-independent, cross-Kerr interaction.

At second order in perturbation theory, there is no correction to the iiSWAP gate frequency Jab(2)=0J_{ab}^{(2)}=0. In the regimes of interest, where the coupler frequency is close enough to the qubit frequencies for the interaction between the coupler and the qubits to be non-negligible, the dominant contribution to the second-order RWA correction to the cross-Kerr interaction χab(2)\chi_{ab}^{(2)} is

χab(2)2(j=a,b,cuajubjucj2𝜶j)2ωa+ωb2ωc+δuacubc[uaa3uba𝜶aubb3uab𝜶b]ωaωb.\displaystyle\begin{split}\chi_{ab}^{(2)}&\approx 2\frac{\left(\sum_{j=a,b,c}u_{aj}u_{bj}u_{cj}^{2}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}+\omega_{b}-2\omega_{c}}\\ &+\delta\frac{u_{ac}u_{bc}\left[u_{aa}^{3}u_{ba}\bm{\alpha}_{a}-u_{bb}^{3}u_{ab}\bm{\alpha}_{b}\right]}{\omega_{a}-\omega_{b}}.\end{split} (28)

The full expression for χab(2)\chi_{ab}^{(2)} can be found in Appendix C. Inspecting the hybridization coefficients uαβu_{\alpha\beta} and the denominators in Eq. 28, we deduce that the second-order correction to the static cross-Kerr interaction, corresponding to the first term in Eq. 28, arises from a virtual two-photon excitation of the coupler (generated by the commutator [a^b^c^2,a^b^c^2][\hat{a}\hat{b}\hat{c}^{\dagger 2},\hat{a}^{\dagger}\hat{b}^{\dagger}\hat{c}^{2}]). This correction would not be present in a two-level approximation McKay et al. (2016). On the other hand, the second term in Eq. 28 is the lowest-order contribution to the dynamical cross-Kerr interaction.

III.3 Improving the starting point of the perturbation theory

As mentioned in the previous subsection, other choices for H^(0)\hat{H}^{(0)} and λH^(1)\lambda\hat{H}^{(1)} are possible which give better accuracy in comparisons with exact Floquet numerics. In this subsection, we take the unperturbed Hamiltonian H^(0)(t)\hat{H}^{(0)}(t) to consist of the Fock-space diagonal part of H^(t)\hat{H}(t), namely

H^(0)(t)=ωaa^a^+ωbb^b^+ωcc^c^+δsin(ωdt)[uca2a^a^+ucb2b^b^+ucc2c^c^]+αa(0)2a^2a^2+αb(0)2b^2b^2+αc(0)2c^2c^2+χab(0)a^a^b^b^+χac(0)a^a^c^c^+χbc(0)b^b^c^c^,\displaystyle\begin{split}\hat{H}^{(0)}(t)&=\omega_{a}\hat{a}^{\dagger}\hat{a}+\omega_{b}\hat{b}^{\dagger}\hat{b}+\omega_{c}\hat{c}^{\dagger}\hat{c}\\ &+\delta\sin(\omega_{\text{d}}t)\left[u_{ca}^{2}\hat{a}^{\dagger}\hat{a}+u_{cb}^{2}\hat{b}^{\dagger}\hat{b}+u_{cc}^{2}\hat{c}^{\dagger}\hat{c}\right]\\ &+\frac{\alpha_{a}^{(0)}}{2}\hat{a}^{\dagger 2}\hat{a}^{2}+\frac{\alpha_{b}^{(0)}}{2}\hat{b}^{\dagger 2}\hat{b}^{2}+\frac{\alpha_{c}^{(0)}}{2}\hat{c}^{\dagger 2}\hat{c}^{2}\\ &+\chi_{ab}^{(0)}\hat{a}^{\dagger}\hat{a}\hat{b}^{\dagger}\hat{b}+\chi_{ac}^{(0)}\hat{a}^{\dagger}\hat{a}\hat{c}^{\dagger}\hat{c}+\chi_{bc}^{(0)}\hat{b}^{\dagger}\hat{b}\hat{c}^{\dagger}\hat{c},\end{split} (29)

where the quartic couplings in the last two rows are exactly those defined in Eq. 27, with the superscript now changed from 11 to 0 to reflect their presence in the unperturbed Hamiltonian.

We expect this starting point, Eq. 29, to lead to more precise results, because of two reasons. Firstly, the perturbation H^(1)\hat{H}^{(1)} is now off-diagonal in Fock-space. The effects of the anharmonicities of the modes are now included at the level of H^(0)\hat{H}^{(0)}, and, in particular, we expect a dressing of contributions corresponding to two-photon excitations, such as Eq. 28. Secondly, due to the second row of Eq. 29, we can derive the effect of harmonics of the drive frequency through a Fourier expansion of the time-evolution operator U^0(t)\hat{U}_{0}(t), defined below.

Following the steps of the previous subsections, we evaluate the time evolution operator with respect to the unperturbed Hamiltonian H^(0)(t)\hat{H}^{(0)}(t), that is

U^0(t)=ei0t𝑑tH^(0)(t),\displaystyle\hat{U}_{0}(t)=e^{-i\int_{0}^{t}dt^{\prime}\hat{H}^{(0)}(t^{\prime})}, (30)

with [H^(0)(t),H^(0)(t)]=0[\hat{H}^{(0)}(t),\hat{H}^{(0)}(t^{\prime})]=0. The time-dependence in the exponent is handled via the Jacobi-Anger identity Abramowitz and Stegun (1964)

eiO^δωdcosωdt=J0(O^δωd)+2n=1inJn(O^δωd)cosnωdt,\displaystyle e^{i\hat{O}\frac{\delta}{\omega_{d}}\cos\omega_{\text{d}}t}=J_{0}\left(\frac{\hat{O}\delta}{\omega_{\text{d}}}\right)+2\sum_{n=1}^{\infty}i^{n}J_{n}\left(\frac{\hat{O}\delta}{\omega_{\text{d}}}\right)\cos n\omega_{\text{d}}t, (31)

for any operator O^\hat{O}, where Jn(z)J_{n}(z) is the nthn^{\textit{th}} Bessel function of the first kind Abramowitz and Stegun (1964). This expansion allows us to keep track of all harmonics of the drive. In practice, since the modulation amplitude is small, δ/ωd1\delta/\omega_{\text{d}}\ll 1, only a few terms will be necessary.

We obtain to first order (and truncating after the first Bessel function)

Jab(1)=δ2ucaucb[J0(δuca2ωd)J0(δucb2ωd)+3J1(δuca2ωd)J1(δucb2ωd)],\displaystyle\begin{split}J_{ab}^{(1)}=\frac{\delta}{2}u_{ca}u_{cb}\Bigg{[}&J_{0}\left(\frac{\delta u_{ca}^{2}}{\omega_{\text{d}}}\right)J_{0}\left(\frac{\delta u_{cb}^{2}}{\omega_{\text{d}}}\right)\\ &+3J_{1}\left(\frac{\delta u_{ca}^{2}}{\omega_{\text{d}}}\right)J_{1}\left(\frac{\delta u_{cb}^{2}}{\omega_{\text{d}}}\right)\Bigg{]},\end{split} (32)

which agrees, up to linear terms in δ\delta, with the expression in Eq. 27. On the other hand, the cross-Kerr interaction at this order is vanishing χab(1)=0\chi_{ab}^{(1)}=0.

To second order in perturbation theory, the dominant contributions to Jab(2)J_{ab}^{(2)} are

Jab(2)=iδ22ucaucbucc2(1ωaωc1ωbωc)×J1(δucc2ωd)j=a,b,cJ0(δucj2ωd)+.\displaystyle\begin{split}J_{ab}^{(2)}=\frac{i\delta^{2}}{2}&u_{ca}u_{cb}u_{cc}^{2}\left(\frac{1}{\omega_{a}-\omega_{c}}-\frac{1}{\omega_{b}-\omega_{c}}\right)\\ &\times J_{1}\left(\frac{\delta u_{cc}^{2}}{\omega_{\text{d}}}\right)\prod_{j=a,b,c}J_{0}\left(\frac{\delta u_{cj}^{2}}{\omega_{\text{d}}}\right)+\ldots.\end{split} (33)

We do not reproduce here the full form containing 20 terms. The second-order contribution χab(2)\chi_{ab}^{(2)} is non-vanishing, and contains approximately 450 terms in expanded form. Despite the complexity of these full expressions, they are easy to derive and manipulate with symbolic computation tools Žitko (2011). Focusing on the static cross-Kerr interaction, i.e. in the δ0\delta\to 0 limit, the dominant correction resulting from the above changes amounts to our previous Eq. 28, but replacing the denominator of the first term of that expression by a form that faithfully includes the contribution from the anharmonicities, as expected in the case of a virtual two-photon excitation of the coupler mode. That is, approximately

ωa+ωb2ωcωa+ωb(2ωc+ucc4𝜶c).\displaystyle\omega_{a}+\omega_{b}-2\omega_{c}\to\omega_{a}+\omega_{b}-(2\omega_{c}+u_{cc}^{4}\bm{\alpha}_{c}). (34)

The full expression for the corrected static cross-Kerr interaction can be found in Appendix C.

In Fig. 2 we compare analytical results to numerical results obtained from exact diagonalization (at δ=0\delta=0), or a solution of the Floquet eigenspectrum at δ0\delta\neq 0 (see Sec. V). We find that the agreement between numerics and analytics is excellent for the gate interaction strength, as well as for the static δ=0\delta=0 cross-Kerr interaction. However, we find that second-order perturbation theory is insufficient to reproduce the effects of the drive on the cross-Kerr interaction, even for modest drive amplitudes. We expect that higher-order perturbation theory should correctly capture the drive-amplitude dependence of the anharmonicities, but these contributions have been inaccessible in our study due to the large memory demands of the computer algebra manipulations.

IV Full circuit Hamiltonian

Building on the previous results, we now turn to deriving an effective Hamiltonian for the full circuit Hamiltonian of Eq. 1 and Eq. (2). The full circuit model goes beyond the toy model in that it systematically includes the effects of the parametric drive on all of the coupling constants. Although the simplicity of the toy model is useful in developing an intuitive understanding of the effect of parametric drives on the system, the full circuit model can lead to more accurate comparisons with experimental data.

The full circuit theory is constructed with the following steps: We first introduce creation and annihilation operators for the bare circuit modes starting from the first-order RWA driven circuit Hamiltonian in Sec. IV.1. Because the drive is taken into account at that level, the frequencies and zero-point fluctuations of these bare modes will be explicitly corrected by the drive. In Sec. IV.2, we perform a normal-mode transformation amounting to a driven black-box quantization approach. We then show in Sec. IV.3 how a variety of quantum gates can be addressed by appropriate choices of the parametric drive frequency. Lastly, we find corrections to the desired gate Hamiltonian using a time-dependent Schrieffer-Wolff perturbation theory in Sec. IV.3.

IV.1 Bare-mode Hamiltonian

To define the bare modes, we begin with the full circuit model Hamiltonian of Eqs. 1 and 2. We normal-order expand the Josephson cosine potentials in this Hamiltonian over a set of creation and annihilation operators, which we define as follows

𝝋^a=ηa2(𝒂^+𝒂^),𝒏^a=i12ηa(𝒂^𝒂^),\displaystyle\begin{split}\hat{\bm{\varphi}}_{a}&=\sqrt{\frac{\eta_{a}}{2}}(\hat{\bm{a}}+\hat{\bm{a}}^{\dagger}),\\ \hat{\bm{n}}_{a}&=-i\sqrt{\frac{1}{2\eta_{a}}}(\hat{\bm{a}}-\hat{\bm{a}}^{\dagger}),\end{split} (35)

with analogous equations for modes 𝒃^\hat{\bm{b}} and 𝒄^\hat{\bm{c}}. The coefficients ηa,b,c\eta_{a,b,c} are chosen such that terms proportional to 𝒂^2,𝒃^2\hat{\bm{a}}^{2},\hat{\bm{b}}^{2}, and 𝒄^2\hat{\bm{c}}^{2} vanish in the time-averaged Hamiltonian. This amounts to three transcendental equations:

(ηj)ηj2=8ECj/EJj,\displaystyle\mathcal{F}(\eta_{j})\eta_{j}^{2}=8E_{\text{C}j}/E_{\text{J}j}, (36)

for j=a,b,cj=a,b,c, where we have defined the form factors

(ηa(b))eηa(b)4,(ηc)αeηc4J0(δ)cos(φ¯ext)+βeηc4N2/N.\displaystyle\begin{split}\mathcal{F}(\eta_{a(b)})&\equiv e^{-\frac{\eta_{a(b)}}{4}},\\ \mathcal{F}(\eta_{c})&\equiv\alpha e^{-\frac{\eta_{c}}{4}}J_{0}(\delta)\cos\left(\overline{\varphi}_{\text{ext}}\right)+\beta e^{-\frac{\eta_{c}}{4N^{2}}}/N.\end{split} (37)

Two remarks are in order: First, in the transmon limit (ηa(b))1\mathcal{F}(\eta_{a(b)})\approx 1 where we recover the usual expression ηa(b)8ECa/b/EJa/b\eta_{a(b)}\approx\sqrt{8E_{\text{C}a/b}/E_{\text{J}a/b}}. Second, the parameter ηc\eta_{c} depends on the parametric drive amplitude δ\delta, which indicates that the mode cc impedance is drive-dependent. This has important consequences for the precision of the calculation of coupling constants dressed by the parametric drives. In particular it allows us to capture the ac-Stark shift of the coupler mode at the lowest order in perturbation theory. In what follows, sine and cosine functions of the phase are normal-order expanded according to Eq. 83 of Appendix D. In turn, trigonometric functions of the flux modulation are expanded in Jacobi-Anger series over the harmonics of the frequency of the drive.

Using the above definitions, the transmon Hamiltonian H^a\hat{H}_{a} takes the familiar form

H^a=𝝎a𝒂^𝒂^EJa(cos𝝋^a+eηa4𝝋^a22).\displaystyle\hat{H}_{a}=\bm{\omega}_{a}\hat{\bm{a}}^{\dagger}\hat{\bm{a}}-E_{\text{J}a}\left(\cos\hat{\bm{\varphi}}_{a}+e^{-\frac{\eta_{a}}{4}}\frac{\hat{\bm{\varphi}}_{a}^{2}}{2}\right). (38)

The second term on the right-hand side contains the nonlinear part of the Josephson potential, i.e. the inductive part is subtracted. Up to quartic order, H^a\hat{H}_{a} takes the form

H^a=𝝎a𝒂^𝒂^+𝜶a2𝒂^2𝒂^2+𝜶a12(𝒂^4+𝒂^4)+𝜶a3(𝒂^𝒂^3+𝒂^3𝒂^)+\displaystyle\begin{split}\hat{H}_{a}&=\bm{\omega}_{a}\hat{\bm{a}}^{\dagger}\hat{\bm{a}}+\frac{\bm{\alpha}_{a}}{2}\hat{\bm{a}}^{\dagger 2}\hat{\bm{a}}^{2}\\ &+\frac{\bm{\alpha}_{a}}{12}\left(\hat{\bm{a}}^{4}+\hat{\bm{a}}^{\dagger 4}\right)+\frac{\bm{\alpha}_{a}}{3}\left(\hat{\bm{a}}^{\dagger}\hat{\bm{a}}^{3}+\hat{\bm{a}}^{\dagger 3}\hat{\bm{a}}\right)+\cdots\end{split} (39)

The first row of this expression is a Kerr oscillator Hamiltonian as in the toy model of Sec. II, whereas the second row contains corrections from quartic counter-rotating terms. Here, we have introduced the mode frequency and anharmonicity which take the forms Koch et al. (2007)

𝝎a=4ECaηa+12(ηa)ηaEJa8ECaEJaECa,𝜶a=ECa.\displaystyle\begin{split}\bm{\omega}_{a}&=\frac{4E_{\text{C}a}}{\eta_{\mathit{a}}}+\frac{1}{2}\mathcal{F}(\eta_{a})\eta_{a}E_{\text{J}a}\approx\sqrt{8E_{\text{C}a}E_{\text{J}a}}-E_{\text{C}a},\\ \bm{\alpha}_{a}&=-E_{\text{C}a}.\end{split} (40)

Note that for the approximate equality in the first row we have used a Taylor expansion of Eq. 36 for ηa\eta_{a}. The equations for mode 𝒃^\hat{\bm{b}} are identical from the above with a change of subscripts and operators aba\to b.

The coupler Hamiltonian differs from that of the transmon modes in two fundamental ways: It breaks parity symmetry due to the external flux, and it is time-dependent. Following Eqs. 13 and 12, we write this time-dependent Hamiltonian as

H^c(t)=H^¯c(t)+H^~c(t).\displaystyle\hat{H}_{c}(t)=\overline{\hat{H}}_{c}(t)+\widetilde{\hat{H}}_{c}(t). (41)

The creation and annihilation operators of the coupler mode can then be defined by extracting the quadratic part of the time-averaged coupler Hamiltonian. Using Eq. 35 where aca\rightarrow c together with

cos[𝝋^c+μαφext(t)]¯=cos(μαφ¯ext)J0(μαδ)cos(𝝋^c)\displaystyle\overline{\cos\left[\hat{\bm{\varphi}}_{c}+\mu_{\alpha}\varphi_{\text{ext}}(t)\right]}=\cos(\mu_{\alpha}\overline{\varphi}_{\text{ext}})J_{0}(\mu_{\alpha}\delta)\cos(\hat{\bm{\varphi}}_{c}) (42)

and a similar relation for the second branch of the coupler [see Eq. 2], we find in analogy to Eq. 38 for the Hamiltonian of the transmon mode

H^¯c=𝝎c𝒄^𝒄^αEJcJ0(μαδ)cos(μαφ¯ext)(cos𝝋^c+eηc4𝝋^c22)βNEJcJ0(μβδ)cos(μβφ¯ext)(cos𝝋^cN+eηc4N2𝝋^c22N2)+αEJcJ0(μαδ)sin(μαφ¯ext)sin𝝋^c+βNEJcJ0(μβδ)sin(μβφ¯ext)sin𝝋^cN.\displaystyle\begin{split}&\overline{\hat{H}}_{c}=\bm{\omega}_{c}\hat{\bm{c}}^{\dagger}\hat{\bm{c}}\\ &-\alpha E_{\text{J}c}J_{0}(\mu_{\alpha}\delta)\cos(\mu_{\alpha}\overline{\varphi}_{\text{ext}})\left(\cos\hat{\bm{\varphi}}_{c}+e^{-\frac{\eta_{c}}{4}}\frac{\hat{\bm{\varphi}}_{c}^{2}}{2}\right)\\ &-\beta NE_{\text{J}c}J_{0}(\mu_{\beta}\delta)\cos(\mu_{\beta}\overline{\varphi}_{\text{ext}})\left(\cos\frac{\hat{\bm{\varphi}}_{c}}{N}+e^{-\frac{\eta_{c}}{4N^{2}}}\frac{\hat{\bm{\varphi}}_{c}^{2}}{2N^{2}}\right)\\ &+\alpha E_{\text{J}c}J_{0}(\mu_{\alpha}\delta)\sin(\mu_{\alpha}\overline{\varphi}_{\text{ext}})\sin\hat{\bm{\varphi}}_{c}\\ &+\beta NE_{\text{J}c}J_{0}(\mu_{\beta}\delta)\sin(\mu_{\beta}\overline{\varphi}_{\text{ext}})\sin\frac{\hat{\bm{\varphi}}_{c}}{N}.\end{split} (43)

Crucially, in this first-order rotating-wave approximation of the parametric drive, the Josephson energy is renormalized by the factor J0(μα,βδ)J_{0}(\mu_{\alpha,\beta}\delta), see also Verney et al. (2019). We interpret this as an effective reduction of the Josephson potential barrier, and consequently an increase of phase fluctuations, in the presence of drives. Moreover, the presence of the non-zero external flux results in the parity breaking sine terms in Eq. 43.

The second term of H^c(t)\hat{H}_{c}(t) in Eq. 41, the oscillatory part, take the form

H^~c(t)=αEJccos[𝝋^c+μαφext(t)]~βNEJccos[𝝋^cN+μβφext(t)]~,\displaystyle\begin{split}\widetilde{\hat{H}}_{c}(t)=&-\alpha E_{\text{J}c}\widetilde{\cos\left[\hat{\bm{\varphi}}_{c}+\mu_{\alpha}\varphi_{\text{ext}}(t)\right]}\\ &-\beta NE_{\text{J}c}\widetilde{\cos\left[\frac{\hat{\bm{\varphi}}_{c}}{N}+\mu_{\beta}\varphi_{\text{ext}}(t)\right]},\end{split} (44)

which can be expanded in a Jacobi-Anger series in harmonics oscillating at the frequency nωdn\omega_{\text{d}}, where nn is an integer.

As above, the next step is to expand the coupler Hamiltonian up to quartic terms in the creation and annihilation operators. In contrast to the transmon Hamiltonian of Eq. 39, parity breaking leads to the appearance of monomials of odd order. The non-oscillatory part is

H^¯c=𝝎c𝒄^𝒄^+𝜶c2𝒄^2𝒄^2+𝜶c12(𝒄^4+𝒄^4)+𝜶c3(𝒄^𝒄^3+𝒄^3𝒄^)+𝒈c,3(𝒄^3+𝒄^3+3𝒄^𝒄^2+3𝒄^2𝒄^)+𝒈c,1(𝒄^+𝒄^)+.\displaystyle\begin{split}\overline{\hat{H}}_{c}&=\bm{\omega}_{c}\hat{\bm{c}}^{\dagger}\hat{\bm{c}}+\frac{\bm{\alpha}_{c}}{2}\hat{\bm{c}}^{\dagger 2}\hat{\bm{c}}^{2}\\ &+\frac{\bm{\alpha}_{c}}{12}\left(\hat{\bm{c}}^{4}+\hat{\bm{c}}^{\dagger 4}\right)+\frac{\bm{\alpha}_{c}}{3}\left(\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{3}+\hat{\bm{c}}^{\dagger 3}\hat{\bm{c}}\right)\\ &+\bm{g}_{c,3}\left(\hat{\bm{c}}^{3}+\hat{\bm{c}}^{\dagger 3}+3\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{2}+3\hat{\bm{c}}^{\dagger 2}\hat{\bm{c}}\right)\\ &+\bm{g}_{c,1}\left(\hat{\bm{c}}+\hat{\bm{c}}^{\dagger}\right)+\cdots.\end{split} (45)

The first row of the above expression takes the form of the coupler Hamiltonian in the approximation of the toy model of Sec. II, while the remaining rows contain counter-rotating terms to quartic order. Here, the parametric drive-dependent mode frequency and anharmonicity read

𝝎c=4ECcηc+12(ηc)ηcEJc,𝜶c=ECc,\displaystyle\begin{split}\bm{\omega}_{c}&=\frac{4E_{\text{C}c}}{\eta_{c}}+\frac{1}{2}\mathcal{F}(\eta_{c})\eta_{c}E_{\text{J}c},\\ \bm{\alpha}_{c}&=-E_{\text{C}c},\\ \end{split} (46)

while the prefactors of the counter-rotating terms are

𝒈c,3=αϵeηc4ηc3/2J0(μαδ)sin(μαφ¯ext)EJc/(122)βN2eηc4N2ηc3/2J0(μβδ)sin(μβφ¯ext)EJc/(122),𝒈c,1=αϵeηc4ηc1/2J0(μαδ)sin(μαφ¯ext)EJc/2+βeηc4N2ηc1/2J0(μβδ)sin(μβφ¯ext)EJc/2.\displaystyle\begin{split}\bm{g}_{c,3}&=-\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{3/2}J_{0}(\mu_{\alpha}\delta)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)E_{\text{J}c}/(12\sqrt{2})\\ &-\frac{\beta}{N^{2}}e^{-\frac{\eta_{c}}{4N^{2}}}\eta_{c}^{3/2}J_{0}(\mu_{\beta}\delta)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)E_{\text{J}c}/(12\sqrt{2}),\\ \bm{g}_{c,1}&=\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{1/2}J_{0}(\mu_{\alpha}\delta)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)E_{\text{J}c}/\sqrt{2}\\ &+\beta e^{-\frac{\eta_{c}}{4N^{2}}}\eta_{c}^{1/2}J_{0}(\mu_{\beta}\delta)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)E_{\text{J}c}/\sqrt{2}.\end{split} (47)

The contribution from the oscillatory part H^~c(t)\widetilde{\hat{H}}_{c}(t) is too lengthy to be reproduced here, and is given up to the second harmonic of the parametric modulation frequency ωd\omega_{\text{d}} in Table 2 of Appendix D.

Finally, the last term of the full circuit Hamiltonian to consider is the linear interaction H^g\hat{H}_{g} induced by the capacitive coupling. Using Eq. 35, this Hamiltonian takes the form

H^g=2ECabηaηb(𝒂^𝒂^)(𝒃^𝒃^)+\displaystyle\hat{H}_{g}=-\frac{2E_{\text{C}ab}}{\sqrt{\eta_{a}\eta_{b}}}(\hat{\bm{a}}-\hat{\bm{a}}^{\dagger})(\hat{\bm{b}}-\hat{\bm{b}}^{\dagger})+\ldots (48)

where the ellipsis represents two more terms corresponding to the cyclic permutations of the mode indices.

The Hamiltonian specified by Eqs. (39) and its equivalent for the bb transmon mode, Eq. 45, the terms summarized in Table 2 of Appendix D, and Eq. 48, form the basis of both the normal-mode analysis, and of the full-circuit numerical simulation performed in Sec. V.

IV.2 Driven black-box quantization approach for parametrically activated interactions

Building upon the above results, we now follow the procedure developed with the toy model in Sec. III.2 to obtain effective gate Hamiltonians under parametric modulations. To do so, we first collect under H^(0)\hat{H}^{(0)} the time-independent quadratic terms derived above. We then eliminate the linear coupling H^g\hat{H}_{g} of Eq. 48 from H^(0)\hat{H}^{(0)} through a normal-mode transformation

H^(0)=𝝎a𝒂^𝒂^+𝝎b𝒃^𝒃^+𝝎c𝒄^𝒄^+H^gωaa^a^+ωbb^b^+ωcc^c^.\displaystyle\begin{split}\hat{H}^{(0)}&=\bm{\omega}_{a}\hat{\bm{a}}^{\dagger}\hat{\bm{a}}+\bm{\omega}_{b}\hat{\bm{b}}^{\dagger}\hat{\bm{b}}+\bm{\omega}_{c}\hat{\bm{c}}^{\dagger}\hat{\bm{c}}+\hat{H}_{g}\\ &\equiv\omega_{a}\hat{a}^{\dagger}\hat{a}+\omega_{b}\hat{b}^{\dagger}\hat{b}+\omega_{c}\hat{c}^{\dagger}\hat{c}.\end{split} (49)

The linear transformation is determined by a set of 18 hybridization coefficients that relate bare mode coordinates to normal mode coordinates (see Appendix E for the procedure to compute these coefficients)

𝝋^α=β=a,b,cuαβ2(β^+β^),𝒏^α=β=a,b,cvαβi2(β^β^),\displaystyle\begin{split}\hat{\bm{\varphi}}_{\alpha}&=\sum_{\beta=a,b,c}\frac{u_{\alpha\beta}}{\sqrt{2}}(\hat{\beta}+\hat{\beta}^{\dagger}),\\ \hat{\bm{n}}_{\alpha}&=\sum_{\beta=a,b,c}\frac{v_{\alpha\beta}}{i\sqrt{2}}(\hat{\beta}-\hat{\beta}^{\dagger}),\end{split} (50)

for α=a,b,c\alpha=a,b,c. We stress that the hybridization coefficients uαβu_{\alpha\beta}, vαβv_{\alpha\beta} depend on the amplitude of the parametric drive. As a result, drive effects such as the ac-Stark shift of the nonlinear oscillators are accounted for already at the level of the normal-mode decomposition of the circuit Hamiltonian.

In analogy with our treatment in Sec. II of the toy model, we have taken H^(0)\hat{H}^{(0)} to be the unperturbed Hamiltonian with respect to which the interaction picture is defined. Primarily in order to keep the expressions more concise, we opt to neglect the corrections analyzed in Sec. III.3. With H^(0)\hat{H}^{(0)} the unperturbed Hamiltonian, the remaining interaction terms are λH^(1)(t)H^H^(0)\lambda\hat{H}^{(1)}(t)\equiv\hat{H}-\hat{H}^{(0)}. Expressing these in the interaction picture with respect to U^0=eiH^(0)t\hat{U}_{0}=e^{-i\hat{H}^{(0)}t} as in Eq. 8, we find

λH^I(1)(t)=U^0(t)[H^(t)H^(0)]U^0(t),\displaystyle\lambda\hat{H}_{I}^{(1)}(t)=\hat{U}_{0}^{\dagger}(t)\left[\hat{H}(t)-\hat{H}^{(0)}\right]\hat{U}_{0}(t), (51)

which, as before, is decomposed into oscillatory and non-oscillatory parts.

As a first example, to realize a beam-splitter interaction, the first-order RWA Hamiltonian is obtained in the form of Eq. 26 for a modulation frequency that satisfies

ωd=ωbωa.\displaystyle\omega_{\text{d}}=\omega_{b}-\omega_{a}. (52)

Importantly, as already mentioned, the right-hand side of the above definition depends implicitly on the drive frequency ωd\omega_{\text{d}}, since it is defined in terms of ac-Stark shifted normal-mode frequencies. In Sec. V we present a numerical procedure to obtain the parametric drive frequency.

With this choice of modulation frequency, the effective Hamiltonian takes the form

λH^¯I(1)=Jab(1)(ia^b^+H.c.)+αa(1)2a^2a^2+αb(1)2b^2b^2+αc(1)2c^2c^2+χab(1)a^a^b^b^+χbc(1)b^b^c^c^+χca(1)c^c^a^a^+Jab;a(1)(ia^a^a^b^+H.c.)+Jab;b(1)(ib^b^a^b^+H.c.)+Jab;c(1)(ic^c^a^b^+H.c.)+Kab(1)(a^2b^2+H.c.).\displaystyle\begin{split}\lambda\overline{\hat{H}}_{I}^{(1)}&=J^{(1)}_{ab}(-i\hat{a}^{\dagger}\hat{b}+\text{H.c.})\\ &+\frac{\alpha_{a}^{(1)}}{2}\hat{a}^{\dagger 2}\hat{a}^{2}+\frac{\alpha_{b}^{(1)}}{2}\hat{b}^{\dagger 2}\hat{b}^{2}+\frac{\alpha_{c}^{(1)}}{2}\hat{c}^{\dagger 2}\hat{c}^{2}\\ &+\chi_{ab}^{(1)}\hat{a}^{\dagger}\hat{a}\hat{b}^{\dagger}\hat{b}+\chi_{bc}^{(1)}\hat{b}^{\dagger}\hat{b}\hat{c}^{\dagger}\hat{c}+\chi_{ca}^{(1)}\hat{c}^{\dagger}\hat{c}\hat{a}^{\dagger}\hat{a}\\ &+J^{(1)}_{ab;a}(-i\hat{a}^{\dagger}\hat{a}\hat{a}^{\dagger}\hat{b}+\text{H.c.})\\ &+J^{(1)}_{ab;b}(-i\hat{b}^{\dagger}\hat{b}\hat{a}^{\dagger}\hat{b}+\text{H.c.})\\ &+J^{(1)}_{ab;c}(-i\hat{c}^{\dagger}\hat{c}\hat{a}^{\dagger}\hat{b}+\text{H.c.})\\ &+K^{(1)}_{ab}(\hat{a}^{\dagger 2}\hat{b}^{2}+\text{H.c.}).\end{split} (53)

In contrast to the effective gate Hamiltonian Eq. 26 obtained for the toy model, there are additional terms in the last four rows, namely photon-number-conditioned beam-splitter terms and a photon-pair beam-splitter term. The prefactors of the above Hamiltonian are

Jab(1)=ucaucb2αϵJ1(μαδ)sin(μαφ¯ext)EJc(α)ucaucb2βNJ1(μβδ)sin(μβφ¯ext)EJc(β),αj(1)=18i=a,b,cuij4EJ,i,χjk(1)=14i=a,b,cuij2uik2EJ,i,Jab;j(1)=ucj24Jab(1), for j=a,b,c,Kab(1)=uca2ucb216αϵJ2(μαδ)cos(μαφ¯ext)EJc(α)uca2ucb216βN3J2(μβδ)cos(μβφ¯ext)EJc(β),\displaystyle\begin{split}J_{ab}^{(1)}&=-\frac{u_{ca}u_{cb}}{2}\alpha\epsilon J_{1}(\mu_{\alpha}\delta)\sin(\mu_{\alpha}\overline{\varphi}_{\text{ext}})E_{\text{J}c}^{(\alpha)}\\ &-\frac{u_{ca}u_{cb}}{2}\frac{\beta}{N}J_{1}(\mu_{\beta}\delta)\sin(\mu_{\beta}\overline{\varphi}_{\text{ext}})E_{\text{J}c}^{(\beta)},\\ \alpha_{j}^{(1)}&=-\frac{1}{8}\sum_{i=a,b,c}u_{ij}^{4}E^{\prime}_{\text{J},i},\\ \chi_{jk}^{(1)}&=-\frac{1}{4}\sum_{i=a,b,c}u_{ij}^{2}u_{ik}^{2}E_{\text{J},i}^{\prime},\\ J_{ab;j}^{(1)}&=-\frac{u_{cj}^{2}}{4}J_{ab}^{(1)},\text{ for }j=a,b,c,\\ K^{(1)}_{ab}&=-\frac{u_{ca}^{2}u_{cb}^{2}}{16}\alpha\epsilon J_{2}(\mu_{\alpha}\delta)\cos(\mu_{\alpha}\overline{\varphi}_{\text{ext}})E_{\text{J}c}^{(\alpha)}\\ &-\frac{u_{ca}^{2}u_{cb}^{2}}{16}\frac{\beta}{N^{3}}J_{2}(\mu_{\beta}\delta)\cos(\mu_{\beta}\overline{\varphi}_{\text{ext}})E_{\text{J}c}^{(\beta)},\end{split} (54)

where we have used

EJaeuaa24uab24uac24EJa,EJbeuba24ubb24ubc24EJb,EJcαϵJ0(μαδ)cos(μαφ¯ext)EJc(α)+(β/N3)J0(μβδ)cos(μβφ¯ext)EJc(β),EJc(α)euca24ucb24ucc24EJc,EJc(β)euca24N2ucb24N2ucc24N2EJc.\displaystyle\begin{split}E_{\text{J}a}^{\prime}&\equiv e^{-\frac{u_{aa}^{2}}{4}-\frac{u_{ab}^{2}}{4}-\frac{u_{ac}^{2}}{4}}E_{\text{J}a},\\ E_{\text{J}b}^{\prime}&\equiv e^{-\frac{u_{ba}^{2}}{4}-\frac{u_{bb}^{2}}{4}-\frac{u_{bc}^{2}}{4}}E_{\text{J}b},\\ E_{\text{J}c}^{\prime}&\equiv\alpha\epsilon J_{0}(\mu_{\alpha}\delta)\cos(\mu_{\alpha}\overline{\varphi}_{\text{ext}})E_{\text{J}c}^{(\alpha)}\\ &+(\beta/N^{3})J_{0}(\mu_{\beta}\delta)\cos(\mu_{\beta}\overline{\varphi}_{\text{ext}})E_{\text{J}c}^{(\beta)},\\ E_{\text{J}c}^{(\alpha)}&\equiv e^{-\frac{u_{ca}^{2}}{4}-\frac{u_{cb}^{2}}{4}-\frac{u_{cc}^{2}}{4}}E_{\text{J}c},\\ E_{\text{J}c}^{(\beta)}&\equiv e^{-\frac{u_{ca}^{2}}{4N^{2}}-\frac{u_{cb}^{2}}{4N^{2}}-\frac{u_{cc}^{2}}{4N^{2}}}E_{\text{J}c}.\end{split} (55)

The above expressions depend on the drive both explicitly, through the Bessel functions, and implicitly, through the hybridization coefficients ujku_{jk}. The first three lines of Eq. 54 are similar in form to those obtained for the toy model in Eqs. (27). Of the two additional classes of terms possible in the full circuit model at this order in perturbation theory, the photon-pair beam-splitter term, in the last row of Eq. 53, is generated by the second harmonic of the drive. However, since this term is fourth order in the hybridization coefficients, it can only become comparable to the beam-splitter interaction at vanishing external flux φ¯ext0\overline{\varphi}_{\text{ext}}\approx 0, or if δ\delta is set to cancel Jab(1)J_{ab}^{(1)}.

As in the case of the toy model, going to second order in perturbation theory using Eq. 14 we find corrections to the coupling constants derived above to first order. To increase the accuracy of these calculations, we first displace the coupler phase variable by a canonical transformation 𝝋^c𝝋^c+𝝋c,cls\hat{\bm{\varphi}}_{c}\to\hat{\bm{\varphi}}_{c}+\bm{\varphi}_{c,\mathrm{cls}}, where the time-independent real number 𝝋c,cls\bm{\varphi}_{c,\mathrm{cls}} denotes the minimum of the static coupler potential in Eq. 2 resulting from

αJ0(μαδϕ)sin[𝝋c,cls+μαφ¯ext]+βJ0(μβδϕ)sin[𝝋c,clsN+μβφ¯ext]=0.\displaystyle\begin{split}&\alpha J_{0}(\mu_{\alpha}\delta\phi)\sin\left[\bm{\varphi}_{c,\mathrm{cls}}+\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right]\\ &+\beta J_{0}(\mu_{\beta}\delta\phi)\sin\left[\frac{\bm{\varphi}_{c,\mathrm{cls}}}{N}+\mu_{\beta}\overline{\varphi}_{\text{ext}}\right]=0.\end{split} (56)

This displacement allows us to remove all static terms which are linear in the coupler quadratures {𝝋^c,𝒏^c}\{\hat{\bm{\varphi}}_{c},\hat{\bm{n}}_{c}\}, before performing the normal-mode transformation. As such, these terms are accounted for exactly, and not as part of the perturbative expansion. Secondly, noting that parity-breaking terms significantly dress the coupler 010\to 1 transition frequency, we absorb this renormalization into a reparametrization of the external flux φ¯extφ¯ext(φ¯ext)\overline{\varphi}_{\text{ext}}\to\overline{\varphi}_{\text{ext}}^{\prime}(\overline{\varphi}_{\text{ext}}) such that ωc(φ¯ext)=ωc(2)(φ¯ext)\omega_{c}(\overline{\varphi}_{\text{ext}}^{\prime})=\omega_{c}^{(2)}(\overline{\varphi}_{\text{ext}}), i.e.  we absorb the corrections to the coupler pole at second order in perturbation theory into a redefinition of the coupler normal mode, in a self-consistent approach that can be further validated with exact numerics.

In Fig. 3 we show a comparison between exact Floquet numerics (see Sec. V) and second-order perturbation theory for the full circuit model. We find that the analytics reproduce with good accuracy the numerical results for the gate interaction rate JabJ_{ab} in the region where the coupler 010\to 1 frequency lies between the two transmons: ωa<ωc<ωa\omega_{a}<\omega_{c}<\omega_{a}. There are poles in the numerical gate rate JabJ_{ab} for ωc<ωa\omega_{c}<\omega_{a} or for ωb<ωc\omega_{b}<\omega_{c} that we expect to capture only at third order in perturbation theory. The numerical cross-Kerr interaction, as in the case of the toy model, only agrees well with analytics in the static case δφ=0\delta\varphi=0. Focusing our attention on the curves obtained from Floquet numerics, we see that with a typical set of parameters gate rates as large as Jab/2π20J_{ab}/2\pi\sim 20 MHz (equivalent to a 25 ns iSWAP\sqrt{i\mathrm{SWAP}} gate) can be achieved while maintaining a vanishing dynamical cross-Kerr interaction. The tools presented in this paper feed into a larger scale optimization of the circuit parameters, which forms the subject of a future study.

Refer to caption
Figure 3: Coupling constants in the effective Hamiltonian for the full circuit as a function of external DC flux φ¯ext\overline{\varphi}_{\text{ext}}. Dots (lines) represent Floquet two-tone spectroscopy data with Hilbert space dimension 10 per mode (second-order RWA calculations). Color (see legend) encodes parametric drive amplitude δφ/2π\delta\varphi/2\pi. Parameter choices: Ca=134.205C_{a}=134.205 fF, Cb=134.218C_{b}=134.218 fF, Cc=75.987C_{c}=75.987 fF, Cac=11.11C_{ac}=11.11 fF, Cbc=11.22C_{bc}=11.22 fF, Cab=0C_{ab}=0, EJa/2π=37E_{\text{J}a}/2\pi=37 GHz, EJb/2π=27E_{\text{J}b}/2\pi=27 GHz, EJc/2π=50E_{\text{J}c}/2\pi=50 GHz, α=0.258,β=1\alpha=0.258,\;\beta=1, and N=3N=3. We attribute large discontinuities in the numerical curves to state tracking errors near avoided crossings (see Sec. V).

IV.3 Other parametric gates

Gate Bosonic operator Drive frequency Dominant unwanted interaction Equation
iiSWAP/beam-splitter ia^b^+ib^a^-i\hat{a}^{\dagger}\hat{b}+i\hat{b}^{\dagger}\hat{a} ωaωb\omega_{a}-\omega_{b} a^a^b^b^\hat{a}^{\dagger}\hat{a}\hat{b}^{\dagger}\hat{b} Eq. 53
Two-mode squeezing ia^b^+ib^a^-i\hat{a}^{\dagger}\hat{b}^{\dagger}+i\hat{b}\hat{a} ωa+ωb\omega_{a}+\omega_{b} a^a^b^b^\hat{a}^{\dagger}\hat{a}\hat{b}^{\dagger}\hat{b} Eqs. (53) and (58)
CZ/Ising-ZZ a^a^b^b^\hat{a}^{\dagger}\hat{a}\hat{b}^{\dagger}\hat{b} no drive Eq. 53
CNOT i(a^a^)b^b^-i(\hat{a}-\hat{a}^{\dagger})\hat{b}^{\dagger}\hat{b} ωa\omega_{a} i(a^a^)a^a^-i(\hat{a}-\hat{a}^{\dagger})\hat{a}^{\dagger}\hat{a} Eq. 59
CSWAP ic^c^(a^b^b^a^)-i\hat{c}^{\dagger}\hat{c}(\hat{a}^{\dagger}\hat{b}-\hat{b}^{\dagger}\hat{a}) ωaωb\omega_{a}-\omega_{b} ia^b^+ib^a^-i\hat{a}^{\dagger}\hat{b}+i\hat{b}^{\dagger}\hat{a} Eq. 53
Table 1: List of the most accessible gate Hamiltonians realizable with a parametric drive in the analyzed architecture.

The space of parametric gates is not limited to beam splitter-type, or red sideband, terms. Indeed, different interactions can be activated by appropriate choices of the frequency of the parametric drive  Bertet et al. (2006); Niskanen et al. (2006, 2007); Beaudoin et al. (2012); McKay et al. (2016); Reagor et al. (2018); Caldwell et al. (2018); Didier et al. (2018). For example, if instead the modulation frequency targets the blue sideband,

ωd=ωa+ωb,\displaystyle\omega_{\text{d}}=\omega_{a}+\omega_{b}, (57)

then the resulting interaction is a two-mode squeezing term. The effective gate Hamiltonian is formally the same as Eq. 53 with the simple modification

a^b^a^b^,\displaystyle\hat{a}^{\dagger}\hat{b}\to\hat{a}^{\dagger}\hat{b}^{\dagger}, (58)

in the first line and in the last four lines of Eq. 53. The coupling constants remain as in Eq. 54.

It is also possible to obtain a CNOT interaction induced by a parametric drive at ωd=ωa\omega_{\text{d}}=\omega_{a}, which makes the aa transmon mode into the target mode of a cross-resonance protocol Rigetti and Devoret (2010); Chow et al. (2011). Following the same procedure as in the preceding subsection, with this choice of modulation frequency we arrive at the effective gate Hamiltonian

λH^¯I(1)=iΩa;b(a^a^)b^b^iΩa;c(a^a^)c^c^iΩa(a^a^)iΩa;a(a^a^a^a^a^a^)\displaystyle\begin{split}\lambda\overline{\hat{H}}_{I}^{(1)}&=-i\Omega_{a;b}(\hat{a}-\hat{a}^{\dagger})\hat{b}^{\dagger}\hat{b}-i\Omega_{a;c}(\hat{a}-\hat{a}^{\dagger})\hat{c}^{\dagger}\hat{c}\\ &-i\Omega_{a}(\hat{a}-\hat{a}^{\dagger})-i\Omega_{a;a}(\hat{a}^{\dagger}\hat{a}\hat{a}-\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a})\end{split} (59)

The first term of above expression generates the cross-resonance gate, while the second term is a coupler-state conditional drive on mode aa which is negligible for c^c^0\langle\hat{c}^{\dagger}\hat{c}\rangle\approx 0. On the other hand, the second row contains local operations on qubit aa.

The coupling constants in Eq. 59 take the form

Ωa=ucaEJ,c′′,Ωa;a=uca3EJ,c′′′/2Ωa;b=ucaucb2EJ,c′′′,Ωa;c=ucaucc2EJ,c′′′,\displaystyle\begin{split}\Omega_{a}&=u_{ca}E_{\text{J},c}^{\prime\prime},\Omega_{a;a}=u_{ca}^{3}E_{\text{J},c}^{\prime\prime\prime}/2\\ \Omega_{a;b}&=u_{ca}u_{cb}^{2}E_{\text{J},c}^{\prime\prime\prime},\Omega_{a;c}=u_{ca}u_{cc}^{2}E_{\text{J},c}^{\prime\prime\prime},\end{split} (60)

where we have defined

EJ,c′′=αϵJ1(μαδ)cos(μαφ¯ext)EJ,c(α)/2+βJ1(μβδ)cos(μβφ¯ext)EJ,c(β)/2,EJ,c′′′=αϵJ1(μαδ)cos(μαφ¯ext)EJ,c(α)/2βJ1(μβδ)cos(μβφ¯ext)EJ,c(β)/(2N2).\displaystyle\begin{split}E_{\text{J},c}^{\prime\prime}&=\alpha\epsilon J_{1}\left(\mu_{\alpha}\delta\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)E_{\text{J},c}^{(\alpha)}/\sqrt{2}\\ &+\beta J_{1}\left(\mu_{\beta}\delta\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)E_{\text{J},c}^{(\beta)}/\sqrt{2},\\ E_{\text{J},c}^{\prime\prime\prime}&=-\alpha\epsilon J_{1}\left(\mu_{\alpha}\delta\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)E_{\text{J},c}^{(\alpha)}/\sqrt{2}\\ &-\beta J_{1}\left(\mu_{\beta}\delta\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)E_{\text{J},c}^{(\beta)}/(\sqrt{2}N^{2}).\end{split} (61)

While in the standard cross-resonance gate protocol the gate is activated by a microwave tone on one of the qubits Rigetti and Devoret (2010); Chow et al. (2011), here it is the coupler mode cc that is parametrically driven. This protocol to achieve a CNOT gate is advantageous if the coupler mode is much more strongly coupled to the transmon modes aa and bb than their direct capacitive coupling. As in the standard cross-resonance protocol Chow et al. (2011); Sheldon et al. (2016), the CNOT gate rate Ωa;b\Omega_{a;b} saturates as a function of the amplitude of the parametric drive, in this model due to the Bessel function dependence J1(μα,βδ)J_{1}\left(\mu_{\alpha,\beta}\delta\right). Table 1 summarizes the different interactions that can be obtained for different choices of modulation frequencies.

V Floquet numerics

In this section we use exact numerical Floquet methods to extract the effective gate Hamiltonian from quasienergy spectra. Floquet theory validates the results obtained using perturbation theory in Secs. III and IV.3. On the other hand, this numerically exact method is applicable beyond the regime of validity of perturbation theory. In this section, we first briefly introduce the method and the notation in Sec. V.1 and, as an example application, return to our toy model to extract the cross-Kerr interaction χab\chi_{ab} and the iSWAP\sqrt{i\mathrm{SWAP}} gate amplitude JabJ_{ab}. Using these results, we show how to adjust the system parameters such as to cancel the dynamical cross-Kerr interaction during an iSWAP\sqrt{i\mathrm{SWAP}} gate. Then, in Sec. V.2, we apply the method to the full circuit Hamiltonian. In particular, we perform a numerical experiment analogous to two-tone spectroscopy for the parametrically driven circuit. For completenes, an introduction to Floquet theory is presented in Appendix F.

V.1 Effective Hamiltonian from Floquet spectra

Our analysis starts from the observation that the effective Hamiltonian is unitarily equivalent to the Floquet Hamiltonian according to Eqs. 8 and 9, and therefore their quasienergy spectra (see Appendix F) are identical. In the laboratory frame, we can write

H^effit=eG^(t)[H^(t)it]eG^(t).\displaystyle\hat{H}_{\textit{eff}}-i\partial_{t}=e^{-\hat{G}(t)}\left[\hat{H}(t)-i\partial_{t}\right]e^{\hat{G}(t)}. (62)

The perturbative expansion for eG^(t)e^{-\hat{G}(t)}, and consequently that for H^eff\hat{H}_{\textit{eff}}, is therefore an iterative approach to finding the Floquet spectrum.

In this section we compute the Floquet spectrum exactly and show how the parameters of the effective Hamiltonian can be extracted from it. Ac-Stark shifted normal mode frequencies, self- and cross-Kerr interactions, and gate amplitudes are formulated as linear combinations of appropriately identified eigenvalues of the Floquet Hamiltonian. For illustration, in this subsection we will confine our attention to the Floquet analysis of the toy model of Eq. 5.

To identify states in the Floquet quasienergy spectrum, we find eigenvectors that have a maximum overlap with a set of known, unperturbed states. We let the state |iaibic\left|i_{a}i_{b}i_{c}\right\rangle be the eigenstate of the time-independent Schrödinger equation for the undriven Hamiltonian, that has maximum overlap with the Fock state |ia|ib|ic\left|i_{a}\right\rangle\left|i_{b}\right\rangle\left|i_{c}\right\rangle, and denote its eigenenergy by EiaibicE_{i_{a}i_{b}i_{c}}. Finally, we define |iaibicF\left|i_{a}i_{b}i_{c}\right\rangle_{F} as the Floquet eigenmode having maximum overlap with |iaibic\left|i_{a}i_{b}i_{c}\right\rangle, and we denote its quasienergy with ϵiaibic\epsilon_{i_{a}i_{b}i_{c}}. In what follows, we label kets by three integers as above, in the order abca-b-c.

Refer to caption
Figure 4: Quasienergies of the Floquet modes with maximum overlap with eigenstates |0a1b0c|0_{a}1_{b}0_{c}\rangle and |1a0b0c|1_{a}0_{b}0_{c}\rangle, for the toy model. The light and dark blue dashed lines correspond to the eigenenergies of the uncoupled system. The inset shows the population of the Floquet states, Pnanbnc(t)=|nanbnc|ψ(t)F|2P_{n_{a}n_{b}n_{c}}(t)=|{}_{F}\langle n_{a}n_{b}n_{c}|\psi(t)\rangle|^{2}, compared to the state populations of a two-level system (dots), driven resonantly with Rabi rate JabJ_{ab}, where JabJ_{ab} is the gate amplitude obtained from the avoided crossing in the Floquet spectrum.

With these definitions, the gate amplitude JabJ_{ab} has a natural interpretation in the Floquet formalism. As shown above, the iSWAP\sqrt{i\mathrm{SWAP}} interaction arises in the toy model if

ωd=ωbωaE100E010.\displaystyle\omega_{\text{d}}=\omega_{b}-\omega_{a}\equiv E_{100}-E_{010}. (63)

Since the parametric drive enters via a term proportional to 𝒄^𝒄^\hat{\bm{c}}^{\dagger}\hat{\bm{c}}, which couples the undriven eigenstates in the two-state manifold {|100,|010}\{\left|100\right\rangle,\left|010\right\rangle\}, there is an avoided crossing between the Floquet modes |100;k+1F\left|100;k+1\right\rangle_{F} and |010;kF\left|010;k\right\rangle_{F}, as shown in Fig. 4. Because the gate operation is analogous to Rabi oscillations in the two-state manifold {|100,|010}\{\left|100\right\rangle,\left|010\right\rangle\}, the size of the avoided crossing is twice the effective gate amplitude, 2Jab2J_{ab}. For example, if an excitation is originally prepared in the transmon bb, then population dynamics would obey P010(t)|010|ψ(t)F|2=sin2(Jabt)P_{010}(t)\equiv|{}_{F}\langle 010|\psi(t)\rangle|^{2}=\sin^{2}\left(J_{ab}t\right) and P100=1P010P_{100}=1-P_{010}, in full agreement with exact numerics (inset of Fig. 4). Away from the avoided crossing, the difference between the dressed states and the undriven states corresponds to the ac-Stark shift of the transmon normal modes due to the off-resonant drive.

Refer to caption
Figure 5: a) Static χab\chi_{ab} interaction at δ=0\delta=0 for the toy model versus the coupler frequency for different values of the coupler anharmonicities, with remaining parameters 𝝎a/2π=4.0\bm{\omega}_{a}/2\pi=4.0, 𝝎b/2π=5.75\bm{\omega}_{b}/2\pi=5.75, 𝜶a/2π=𝜶b/2π=0.2\bm{\alpha}_{a}/2\pi=\bm{\alpha}_{b}/2\pi=-0.2, and 𝒈ac/2π=𝒈bc/2π=0.05\bm{g}_{ac}/2\pi=-\bm{g}_{bc}/2\pi=0.05 GHz. b) Dynamical χab\chi_{ab} versus bare coupler frequency for the parameters above and 𝜶c/2π=0.12\bm{\alpha}_{c}/2\pi=0.12 GHz, for different values of the drive amplitude. c) Gate amplitude JabJ_{ab} for the parameters in b). The Hilbert space dimension for each mode is 5.

Note that, in practice, the two-state manifold {|100,|010}\{\left|100\right\rangle,\left|010\right\rangle\} is coupled by the drive to other levels. The resonant drive frequency ωd\omega_{\text{d}} is then slightly shifted from Eq. 63 due to the ac-Stark effect induced by these additional couplings, and the exact value can be determined numerically by minimizing the size of the anticrossing.

The dynamical cross-Kerr interaction χab\chi_{ab} is written in terms of a Walsh transform Berke et al. (2020) of the quasienergies

χab(δ)=ϵ110ϵ100ϵ010+ϵ000,\displaystyle\chi_{ab}(\delta)=\epsilon_{110}-\epsilon_{100}-\epsilon_{010}+\epsilon_{000}, (64)

and reduces to the static cross-Kerr when the parametric drive is turned off:

χab(0)=E110E100E010+E000.\displaystyle\chi_{ab}(0)=E_{110}-E_{100}-E_{010}+E_{000}. (65)

Along with JabJ_{ab} and χab\chi_{ab}, any ac-Stark-shifted quantity pertaining to the effective Hamiltonian can, in principle, be obtained by taking appropriate linear combinations of the quasienergies in the Floquet spectrum.

Since the Floquet quasienergy spectrum can be obtained from the propagator U^(2π/ωd,0)\hat{U}(2\pi/\omega_{\text{d}},0) over one period of the drive (Appendix F), the Floquet method is numerically efficient as compared to the simulation of the dynamics over the complete gate time. The period of the drive is on the order of 1ns1\,\text{ns}, which is between two and three orders of magnitude shorter than the gate times studied here. Due to its relatively small computational footprint, the Floquet method allows us to efficiently search for optimal gate parameters, e.g. a maximal JabJ_{ab} with a minimal residual cross-Kerr interaction, χab\chi_{ab}. As an example, in  Fig. 5 we study the behavior of JabJ_{ab} and χab\chi_{ab} as a function of the bare coupler frequency 𝝎c\bm{\omega}_{c} for different choices of drive amplitude, δ\delta, and bare coupler anharmonicity, 𝜶c\bm{\alpha}_{c}.

From these studies we can, for example, find parameters for which the cross-Kerr interaction χab\chi_{ab} vanishes. As already mentioned, this is important to obtain high-fidelity two-qubit gates and relies on choosing a positive coupler anharmonicity, 𝜶c\bm{\alpha}_{c}. In Fig. 5 we find that, while varying 𝜶c\bm{\alpha}_{c} does not affect JabJ_{ab} to lowest order in perturbation theory, it has a considerable impact on χab\chi_{ab}. Indeed, Fig. 5a) shows the static χab\chi_{ab} for multiple values of 𝜶c\bm{\alpha}_{c}, and illustrates that it is possible to fine-tune 𝜶c\bm{\alpha}_{c} to cancel χab\chi_{ab}. We observe empirically that whenever the bare anharmonicities obey 𝜶a1+𝜶b1+𝜶c1=0\bm{\alpha}_{a}^{-1}+\bm{\alpha}_{b}^{-1}+\bm{\alpha}_{c}^{-1}=0, one can find 𝝎c\bm{\omega}_{c} for which χab=0\chi_{ab}=0.

For the dynamical cross-Kerr interaction χab\chi_{ab} one observes complex variations with δ\delta. The main resonances appear for 𝝎c𝝎a\bm{\omega}_{c}\approx\bm{\omega}_{a}, 𝝎b\bm{\omega}_{b}, and (𝝎a+𝝎b)/2(\bm{\omega}_{a}+\bm{\omega}_{b})/2 but the slopes and the sign of χab\chi_{ab} change and additional resonances appear away from the qubit frequencies, especially when the drive amplitude and coupling strengths gab,acg_{ab,ac} are sufficiently large. As illustrated in Fig. 5b), by tuning the drive amplitude it is possible to find a bare coupler frequency, 𝝎c\bm{\omega}_{c}, for which the effective χab(δ)=0\chi_{ab}(\delta)=0. Moreover, the cancellation of χab(δ)\chi_{ab}(\delta) can occur in the form of a ‘sweet spot’ where χab/δ0\partial\chi_{ab}/\partial\delta\approx 0. On the other hand, as seen in Fig. 5c), the gate rate increases with δ\delta without qualitative changes of its dependence on 𝝎c\bm{\omega}_{c}. Therefore, as the gate is turned on or off by varying δ\delta, one can adjust the bare coupler frequency 𝝎c\bm{\omega}_{c} to maintain the instantaneous χab(δ)=0\chi_{ab}(\delta)=0. This defines a cross-Kerr-free curve in the parameter space (𝝎c,δ)(\bm{\omega}_{c},\delta) connecting the ‘off’ point δ=0,χab(0)=0,Jab(0)=0\delta=0,\chi_{ab}(0)=0,J_{ab}(0)=0 to the ‘on’ point δ0,χab(δ)=0,Jab(δ)0\delta\neq 0,\chi_{ab}(\delta)=0,J_{ab}(\delta)\neq 0. We study this in detail on the realistic full circuit model in Sec. V.2.

Refer to caption
Figure 6: Same parameter choices as Fig. 3. Gate amplitude JabJ_{ab} (a) and cross-Kerr χab\chi_{ab} (b) for different parameteric drive amplitudes δφ\delta\varphi (encoded in curve color) and as a function of the static flux φ¯ext\overline{\varphi}_{\text{ext}} from Floquet simulations. Data has been excluded where deficient state-tracking in the vicinity of avoided crossings led to unphysical discontinuities in the quantities. c) for regions I,III,II, and IIIIII identified in panels a) and b), we eliminate the common parameter φ¯ext\overline{\varphi}_{\text{ext}} and plot Jab(χab)J_{ab}(\chi_{ab}). This allows us to identify those regimes in which the iSWAP\sqrt{i\mathrm{SWAP}} gate interaction can be turned on, while maintaining a vanishing dynamical χab\chi_{ab}.

V.2 Full circuit simulation

In this section, we apply the Floquet numerical method to the full circuit Hamiltonian of Sec. IV.1. We study the dependence of the coupling constants in the effective gate Hamiltonian versus DC flux and as a function of the drive amplitude.

Figure 6 shows the analogues of the plots in Fig. 5 for the gate amplitude Jab(φ¯ext)J_{ab}(\overline{\varphi}_{\text{ext}}) and of the cross-Kerr χab(φ¯ext)\chi_{ab}(\overline{\varphi}_{\text{ext}}) now for the full circuit Hamiltonian. State tracking is performed as described in the previous subsection. However, in the vicinity of avoided crossings, it is impossible to identify with certainty the states generated by the relatively large capacitive couplings considered here. We therefore introduce exclusion regions where state tracking is unreliable. Even though the tracking is expected to be complicated by the presence of counter-rotating terms coupling states with different photon numbers in the full device Hamiltonian of Sec. IV.1, we find that this is not a significant source of tracking error, as compared to errors due to large hybridization.

In Fig. 6b), we represent χab\chi_{ab} versus the DC-flux φ¯ext\overline{\varphi}_{\text{ext}} for different values of the flux drive amplitude. Unlike the toy model, χab\chi_{ab} does not go to zero away from the qubit-coupler resonances. This is because the qubit-coupler detuning saturates as a function of φ¯ext\overline{\varphi}_{\text{ext}}, as opposed to the toy model where the detuning could be increased arbitrarily. In the undriven case (black), we see that for this set of device parameters there does not exist a flux value for which χab\chi_{ab} vanishes. However, increasing the drive amplitude allows for an active cancellation of the dynamical χab\chi_{ab} at some flux value. The corresponding behavior of JabJ_{ab} is shown in Fig. 6a). In Fig. 6c), we synthesize the numerical results into three favorable regions of operation for the parametric gate, denoted II, IIII, and IIIIII, respectively [see panel a)]. For these regions, we eliminate the external flux and plot directly the gate amplitude JabJ_{ab} against the dynamical cross-Kerr interaction χab\chi_{ab}. This allows us to determine regimes of optimal iSWAP\sqrt{i\mathrm{SWAP}} gate operation. We conclude that gate amplitudes as high as 4040 MHz, corresponding to a gate time of 12.512.5 ns, can be achieved with vanishing cross-Kerr interaction for these parameter choices.

For both JabJ_{ab} and χab\chi_{ab} there exist peaks away from the qubit-coupler resonances, situated at φ¯ext/2π0.13,0.42\bar{\varphi}_{ext}/2\pi\approx 0.13,0.42. These correspond to avoided crossings appearing in the driven Floquet spectrum, corresponding to the hybridization of Floquet levels involving distinct numbers of drive photons. For example, the Floquet level |100,kF|100,k\rangle_{F} can couple to the Floquet level |001,k1F|001,k-1\rangle_{F}. This can be seen by unfolding the Floquet spectrum in spectroscopy simulations (see Fig. 7).

Refer to caption
Figure 7: Two-tone spectroscopy data from Floquet numerics. Each point corresponds to a possible transition and its size is weighed by the matrix element of the charge operator of bare qubit aa, X^=𝒏^a\hat{X}=\hat{\bm{n}}_{a}. Parameters chosen as in Fig. 3 with δφ/2π=\delta\varphi/2\pi= 0.0 (black dots) and 0.03 (crosses). The subscripts α,β\alpha,\beta sweep over the subset of Floquet modes {|000F,|100F,|010F,|001F}\{\left|000\right\rangle_{F},\left|100\right\rangle_{F},\left|010\right\rangle_{F},\left|001\right\rangle_{F}\}, whereas the drive photon number kk takes integer values between 15-15 and 1515.

To exemplify the full extent of the Floquet analysis, we generate two-tone spectroscopy data from our simulations according to Eqs. 97 and 98 in Appendix F, by focusing on the experimentally relevant situation where the parametric drive is on, while the (second) probe tone acts on the bare charge operator 𝒏^a\hat{\bm{n}}_{a}. In Fig. 7 we represent the numerically computed spectrum close to the two qubit transition frequencies. The size of each point is proportional to the absolute value of the corresponding matrix element. The black dots correspond to transition frequencies in the undriven spectrum. As expected, the dot sizes are larger for the transitions involving the probed qubit aa. The large avoided crossings around φ¯ext/2π{0.32,0.38}\overline{\varphi}_{\text{ext}}/2\pi\approx\{0.32,0.38\} result from the capacitive couplings between the coupler and the qubits. Secondary avoided crossings appear between the coupler mode and the transmons near φ¯ext/2π{0.13,0.42}\bar{\varphi}_{ext}/2\pi\approx\{0.13,0.42\} in the driven spectrum, and are responsible for the secondary poles mentioned in the discussion of the coupling constants of the effective Hamiltonian, Fig. 6. Furthermore, as we detail in Appendix G, counterrotating terms induce important corrections when attempting an accurate comparison with spectroscopic data from experiments.

VI Conclusion

In summary, we have presented two complementary methods for the analysis of parametrically activated two-qubit gates, one based on analytical time-dependent Schrieffer-Wolff perturbation theory, and one based on numerical Floquet methods. Although we have mostly focused on coupler-mediated parametric iSWAP\sqrt{i\mathrm{SWAP}} gates, a larger collection of gates can be generated in the same model Hamiltonian. The methods presented here allow one to efficiently evaluate the terms present in the effective gate Hamiltonian.

For the iSWAP\sqrt{i\mathrm{SWAP}} interaction, we have shown that with experimentally accessible parameters, a gate frequency of 40\sim 40 MHz corresponding to a gate time as short as 12.512.5 ns can be obtained with vanishing dynamical cross-Kerr interaction. This fast gate is achieved by working with large capacitive couplings between the qubits and the coupler, while cancelling the cross-Kerr interactions by setting the coupler anharmonicity to positive values, and choosing the right modulation amplitude. Optimization of realistic device parameters based on close agreements between the Floquet simulations and the experimental data will be published elsewhere Pranav S. Mundada, Sara Sussman et al. (2021).

We have argued that the analytical method introduced here and which is based on a drive-dependent normal-mode expansion is a computationally efficient strategy to organize the perturbation theory as compared to an energy eigenbasis calculation, for it allows to obtain the parameters of the effective Hamiltonian at lower orders in perturbation theory. Moreover, this strategy is suitable in the regime of comparatively large linear couplings, where the dispersive approximation breaks down. Nonetheless, we have shown that higher orders in analytical perturbation theory are needed for full agreement with exact numerical results, especially for higher-order interactions, such as the dynamical cross-Kerr. Generating higher-order contributions efficiently using computer algebra techniques is the subject of future studies. On the other hand, this work indicates that Floquet numerical methods, as compared to full time-dynamics simulations, is a numerically efficient and exact method for minute optimization studies of parametric gates.

Acknowledgments

We thank Joachim Cohen, Moein Malekakhlagh, and Baptiste Royer for useful discussions. We are grateful to Ross Shillito for help with optimizing numerical simulations. This work was undertaken thanks to funding from NSERC, the Canada First Research Excellence Fund, and the U.S. Army Research Office Grant No. W911NF-18-1-0411.

Appendix A Time-dependent Schrieffer-Wolff transformation

To obtain equations for G^I(t)\hat{G}_{\text{I}}(t), we assume that the generator can be expanded as a series in λ\lambda, that is

G^I(t)=λG^I(1)(t)+λ2G^I(2)(t)+,\displaystyle\hat{G}_{\text{I}}(t)=\lambda\hat{G}_{\text{I}}^{(1)}(t)+\lambda^{2}\hat{G}_{\text{I}}^{(2)}(t)+\cdots, (66)

and collect powers of λ\lambda in the BCH expansion of Eq. 9

eG^I(H^Iit)eG^I=λH^I(1)iλG^˙I(1)+[λH^I(1),λG^I(1)]i2[λG^˙I(1),λG^I(1)]iλ2G^˙I(2)it+O(λ3).\displaystyle\begin{split}&e^{-\hat{G}_{\text{I}}}(\hat{H}_{\text{I}}-i\partial_{t})e^{\hat{G}_{\text{I}}}=\lambda\hat{H}_{\text{I}}^{(1)}-i\lambda\dot{\hat{G}}^{(1)}_{\text{I}}\\ &+[\lambda\hat{H}_{\text{I}}^{(1)},\lambda\hat{G}_{\text{I}}^{(1)}]-\frac{i}{2}[\lambda\dot{\hat{G}}^{(1)}_{\text{I}},\lambda\hat{G}_{\text{I}}^{(1)}]-i\lambda^{2}\dot{\hat{G}}^{(2)}_{\text{I}}\\ &-i\partial_{t}+O(\lambda^{3}).\end{split} (67)

The above expansion can be expressed compactly

eG^I(H^Iit)eG^I=λkk=1[H^I(k)(t)iG^˙I(k)]it.\displaystyle e^{-\hat{G}_{\text{I}}}(\hat{H}_{\text{I}}-i\partial_{t})e^{\hat{G}_{\text{I}}}=\lambda^{k}\sum_{k=1}^{\infty}\left[\hat{H}_{\text{I}}^{(k)}(t)-i\dot{\hat{G}}^{(k)}_{\text{I}}\right]-i\partial_{t}. (68)

Provided a prescription for λkG^I(k)(t)\lambda^{k}\hat{G}_{\text{I}}^{(k)}(t), we have a recursive way of determining higher-order corrections to the interaction Hamiltonian: knowledge of λH^I(1)(t)\lambda\hat{H}_{I}^{(1)}(t) allows one to determine λ2H^I(2)\lambda^{2}\hat{H}_{I}^{(2)}, then λ3H^I(3)\lambda^{3}\hat{H}_{I}^{(3)} etc.

The kthk^{\textit{th}} order term in the generator, λkG^I(k)(t)\lambda^{k}\hat{G}_{\text{I}}^{(k)}(t), is determined by the condition that the Hamiltonian be free of oscillatory terms of order λk\lambda^{k} or less. This condition can be formulated explicitly if we write, as in Eq. 11,

λkH^I(k)(t)=λkH^¯I(k)+λkH^~I(k)(t).\displaystyle\lambda^{k}\hat{H}_{I}^{(k)}(t)=\lambda^{k}\overline{\hat{H}}_{I}^{(k)}+\lambda^{k}\widetilde{\hat{H}}_{I}^{(k)}(t). (69)

Then oscillatory terms λkH^~I(k)\lambda^{k}\widetilde{\hat{H}}_{I}^{(k)} are canceled for every kk if

λkG^I(k)(t)=1i0tλkH^~I(k)(t).\displaystyle\lambda^{k}\hat{G}_{\text{I}}^{(k)}(t)=\frac{1}{i}\int_{0}^{t}\lambda^{k}\widetilde{\hat{H}}_{I}^{(k)}(t). (70)

Note that, in the above expression, we have imposed the boundary condition G^I(k)(0)=0\hat{G}_{\text{I}}^{(k)}(0)=0 by specifying the lower limit of the integration. Noting that Eq. 70 implies

λkG^I~(k)(t)=1itλkH^~I(k)(t),λkG^I¯(k)(t)=1i[tλkH^~I(k)(t)]t=0.\displaystyle\begin{split}\lambda^{k}\widetilde{\hat{G}_{\text{I}}}^{(k)}(t)=&\frac{1}{i}\int^{t}\lambda^{k}\widetilde{\hat{H}}_{I}^{(k)}(t),\\ \lambda^{k}\overline{\hat{G}_{\text{I}}}^{(k)}(t)=&\frac{1}{i}\left[\int^{t}\lambda^{k}\widetilde{\hat{H}}_{I}^{(k)}(t)\right]_{t=0}.\end{split} (71)

The DC part of the generator, λkG^I¯(k)\lambda^{k}\overline{\hat{G}_{\text{I}}}^{(k)}, is non-vanishing here as a result of the boundary condition in Eq. 70, as opposed to the zero time-average property of kick operators, to which the generator studied here is related Goldman and Dalibard (2014).

With the above formalism in place, we are now ready to compute perturbative corrections. From Eq. 9 we identify the λ2\lambda^{2} correction to the interaction-picture Hamiltonian

λ2H^I(2)(t)=[λH^I(1),λG^I(1)]i2[λG^˙I(1),λG^I(1)].\displaystyle\lambda^{2}\hat{H}_{\text{I}}^{(2)}(t)=[\lambda\hat{H}_{\text{I}}^{(1)},\lambda\hat{G}_{\text{I}}^{(1)}]-\frac{i}{2}[\lambda\dot{\hat{G}}^{(1)}_{\text{I}},\lambda\hat{G}_{\text{I}}^{(1)}]. (72)

Going ahead and solving the RWA condition in Eq. 70 at order λ1\lambda^{1}, we find the order-λ2\lambda^{2} RWA Hamiltonian

λ2H^¯I(2)=1i[H^¯I(1),0tλH^~I(1)(t)𝑑t]¯+12i[λH^~I(1)(t),0tλH^~I(1)(t)𝑑t]¯.\displaystyle\begin{split}\lambda^{2}\overline{\hat{H}}_{I}^{(2)}=&\frac{1}{i}\overline{\left[\overline{\hat{H}}_{I}^{(1)},\int_{0}^{t}\lambda\widetilde{\hat{H}}_{I}^{(1)}(t^{\prime})dt^{\prime}\right]}\\ &+\frac{1}{2i}\overline{\left[\lambda\widetilde{\hat{H}}_{I}^{(1)}(t),\int_{0}^{t}\lambda\widetilde{\hat{H}}_{I}^{(1)}(t^{\prime})dt^{\prime}\right]}.\end{split} (73)

This procedure can be iterated to higher orders, with increasing complexity due to the proliferation of terms from nested commutators in the BCH expansion.

Appendix B Circuit quantization

In this appendix we derive the model Hamiltonian of Eq. 1 from the circuit Lagrangian corresponding to Fig. 8. Assuming the individual modes of the junction array have small impedance, guaranteed by sufficiently large Josephson energy, the junction array can be described by an effective one-dimensional Lagrangian where the total phase difference across the array is spread evenly through the junctions. The effective one-dimensional Lagrangian associated with the bare coupler mode is

c=k=α,βCk2ϕ˙k2+αEJccos[𝝋α]+βNEJccos[𝝋βN],\displaystyle\mathcal{L}_{c}=\sum_{k=\alpha,\beta}\frac{C_{k}}{2}\dot{\bm{\phi}}_{k}^{2}+\alpha E_{\text{J}c}\cos\left[\bm{\varphi}_{\alpha}\right]+\beta NE_{\text{J}c}\cos\left[\frac{\bm{\varphi}_{\beta}}{N}\right], (74)

where ϕα\bm{\phi}_{\alpha} is the branch flux across the small junction and the shunt capacitor with total capacitance CαC_{\alpha}, ϕβ\bm{\phi}_{\beta} is the branch flux across the junction array with effective capacitance CβC_{\beta}, and 𝝋k=2πϕk/Φ0\bm{\varphi}_{k}=2\pi\bm{\phi}_{k}/\Phi_{0} are the associated reduced phase variables, and Φ0\Phi_{0} is the superconducting flux quantum. The phases 𝝋α\bm{\varphi}_{\alpha} and 𝝋β\bm{\varphi}_{\beta} are constrained by the fluxoid quantization, 𝝋α+𝝋β=φext\bm{\varphi}_{\alpha}+\bm{\varphi}_{\beta}=\varphi_{\rm ext}. We define the new coordinates

𝝋α=𝝋c+μαφext,𝝋β=𝝋cNμβφext,\displaystyle\begin{split}\bm{\varphi}_{\alpha}&=\bm{\varphi}_{c}+\mu_{\alpha}\varphi_{\rm ext},\\ \bm{\varphi}_{\beta}&=-\bm{\varphi}_{c}-N\mu_{\beta}\varphi_{\rm ext},\end{split} (75)

with μαNμβ=1\mu_{\alpha}-N\mu_{\beta}=1, such that the capacitive energy in the Lagrangian is now purely quadratic in ϕ˙c\dot{\bm{\phi}}_{c}. We thus require Cαμα+CβNμβ=0C_{\alpha}\mu_{\alpha}+C_{\beta}N\mu_{\beta}=0. We obtain

μα=CβCα+Cα,μβ=1NCαCα+Cβ.\displaystyle\begin{split}\mu_{\alpha}&=\frac{C_{\beta}}{C_{\alpha}+C_{\alpha}},\\ \mu_{\beta}&=-\frac{1}{N}\frac{C_{\alpha}}{C_{\alpha}+C_{\beta}}.\end{split} (76)

Up to time-dependent scalar terms, we obtain the form

c=Cc2ϕ˙c2+αEJccos[𝝋c+μαφext]+βNEJccos[𝝋cN+μβφext].\displaystyle\begin{split}\mathcal{L}_{c}=\frac{C_{c}}{2}\dot{\bm{\phi}}_{c}^{2}+\alpha E_{\text{J}c}\cos\left[\bm{\varphi}_{c}+\mu_{\alpha}\varphi_{\rm ext}\right]\\ +\beta NE_{\text{J}c}\cos\left[\frac{\bm{\varphi}_{c}}{N}+\mu_{\beta}\varphi_{\rm ext}\right].\end{split} (77)

Moreover, for the two bare transmon modes j=a,bj=a,b the Lagrangian reads

j=Cj2ϕ˙j2+EJjcos𝝋j.\displaystyle\mathcal{L}_{j}=\frac{C_{j}}{2}\dot{\bm{\phi}}_{j}^{2}+E_{\text{J}j}\cos\bm{\varphi}_{j}. (78)

The total Lagrangian of the system then takes the form

=a+b+c+g,\displaystyle\mathcal{L}=\mathcal{L}_{a}+\mathcal{L}_{b}+\mathcal{L}_{c}+\mathcal{L}_{g}, (79)

where we have introduced the capacitive coupling between the three bare modes

g=Cab2ϕ˙aϕ˙b+Cbc2ϕ˙bϕ˙c+Cca2ϕ˙cϕ˙a.\displaystyle\mathcal{L}_{g}=\frac{C_{ab}}{2}\dot{\bm{\phi}}_{a}\dot{\bm{\phi}}_{b}+\frac{C_{bc}}{2}\dot{\bm{\phi}}_{b}\dot{\bm{\phi}}_{c}+\frac{C_{ca}}{2}\dot{\bm{\phi}}_{c}\dot{\bm{\phi}}_{a}. (80)

Appendix C Perturbation theory for the toy model

In this section, we reproduce expressions for the cross-Kerr interaction obtained to second-order in perturbation theory for the toy model. The full expression of the second-order RWA correction to the cross-Kerr interaction in Sec. III.2 reads

χab,Sec. III.2(2)=4(j=a,b,cuaj2ubjucj𝜶j)2ωbωc+4(j=a,b,cuajubj2ucj𝜶j)2ωaωc+2(j=a,b,cuajubjucj2𝜶j)2ωa+ωb2ωc2(j=a,b,cuaj3ubj𝜶j)2ωaωb+2(j=a,b,cuajubj3𝜶j)2ωaωb+uacubcj=a,b,cuajubj(uaj2ubj2)ωaωbδ.\displaystyle\begin{split}\chi_{ab,\text{\lx@cref{creftype~refnum}{Subsec:EffGateToy}}}^{(2)}&=\frac{4\left(\sum_{j=a,b,c}u_{aj}^{2}u_{bj}u_{cj}\bm{\alpha}_{j}\right)^{2}}{\omega_{b}-\omega_{c}}\\ &+\frac{4\left(\sum_{j=a,b,c}u_{aj}u_{bj}^{2}u_{cj}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}-\omega_{c}}\\ &+\frac{2\left(\sum_{j=a,b,c}u_{aj}u_{bj}u_{cj}^{2}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}+\omega_{b}-2\omega_{c}}\\ &-\frac{2\left(\sum_{j=a,b,c}u_{aj}^{3}u_{bj}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}-\omega_{b}}\\ &+\frac{2\left(\sum_{j=a,b,c}u_{aj}u_{bj}^{3}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}-\omega_{b}}\\ &+\frac{u_{ac}u_{bc}\sum_{j=a,b,c}u_{aj}u_{bj}\left(u_{aj}^{2}-u_{bj}^{2}\right)}{\omega_{a}-\omega_{b}}\delta.\end{split} (81)

The second-order correction to the static cross-Kerr interaction as calculated in Sec. III.3 is:

χab,Sec. III.3(2)=4(j=a,b,cuaj2ubjucj𝜶j)2ωbωc+j=a,b,c2uaj2(ubj2ucj2)𝜶j+4(j=a,b,cuajubj2ucj𝜶j)2ωaωc+j=a,b,c2ubj2(uaj2ucj2)𝜶j+2(j=a,b,cuajubjucj2𝜶j)2ωa+ωb2ωc+(2uaj2ubj2ucj4)𝜶j2(j=a,b,cuaj3ubj𝜶j)2ωaωb+j=a,b,c(uaj42uaj2ubj2)𝜶j+2(j=a,b,cuajubj3𝜶j)2ωaωb+j=a,b,c(2uaj2ubj2ubj4)𝜶a.\displaystyle\begin{split}\chi_{ab,\text{\lx@cref{creftype~refnum}{Sec:AnhCorrections}}}^{(2)}&=\frac{4\left(\sum_{j=a,b,c}u_{aj}^{2}u_{bj}u_{cj}\bm{\alpha}_{j}\right)^{2}}{\omega_{b}-\omega_{c}+\sum_{j=a,b,c}2u_{aj}^{2}\left(u_{bj}^{2}-u_{cj}^{2}\right)\bm{\alpha}_{j}}\\ &+\frac{4\left(\sum_{j=a,b,c}u_{aj}u_{bj}^{2}u_{cj}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}-\omega_{c}+\sum_{j=a,b,c}2u_{bj}^{2}\left(u_{aj}^{2}-u_{cj}^{2}\right)\bm{\alpha}_{j}}\\ &+\frac{2\left(\sum_{j=a,b,c}u_{aj}u_{bj}u_{cj}^{2}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}+\omega_{b}-2\omega_{c}+\left(2u_{aj}^{2}u_{bj}^{2}-u_{cj}^{4}\right)\bm{\alpha}_{j}}\\ &-\frac{2\left(\sum_{j=a,b,c}u_{aj}^{3}u_{bj}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}-\omega_{b}+\sum_{j=a,b,c}\left(u_{aj}^{4}-2u_{aj}^{2}u_{bj}^{2}\right)\bm{\alpha}_{j}}\\ &+\frac{2\left(\sum_{j=a,b,c}u_{aj}u_{bj}^{3}\bm{\alpha}_{j}\right)^{2}}{\omega_{a}-\omega_{b}+\sum_{j=a,b,c}\left(2u_{aj}^{2}u_{bj}^{2}-u_{bj}^{4}\right)\bm{\alpha}_{a}}.\end{split} (82)

The expression for the dynamical cross-Kerr interaction, χab,Sec. III.3(2)\chi_{ab,\text{\lx@cref{creftype~refnum}{Sec:AnhCorrections}}}^{(2)} at δ0\delta\neq 0, is available from the formalism, but it is too lengthy to be reproduced here. In the main text, an evaluation of this expression is used in making direct comparisons to exact numerics.

Refer to caption
Figure 8: Circuit schematic and notations used in the derivation of the circuit Lagrangian in Appendix B. The coupler consists of two branches of total capacitances CαC_{\alpha} and CβC_{\beta} (not indicated in the figure). The α\alpha branch consists of a single Josephson junction, while the ‘β\beta’ branch NN junctions in series. The bare coupler and transmon modes are connected capacitively through coupling capacitances Cab,bc,caC_{ab,bc,ca}.

Appendix D Details for full circuit Hamiltonian

In this Appendix, we record a number of results used in Sec. IV, in particular the formulae used for normal-ordered expansions in Sec. D.1, and the time-dependent terms in the coupler Hamiltonian in Sec. D.2.

D.1 Normal-ordered expansions of trigonometric functions. Jacobi-Anger expansions

Sine and cosine are expanded in normal order using the following two expressions Marcos et al. (2013) [recall that φ^a=ηa/2(a^+a^)\hat{\varphi}_{a}=\sqrt{\eta_{a}/2}(\hat{a}+\hat{a}^{\dagger})]

cosφ^a=eηa4m,n0m+n= even(ηa2)m+n2a^ma^nm!n!,sinφ^a=eηa4ηa2m,n0m+n= odd(ηa2)m+n12a^ma^nm!n!,\displaystyle\begin{split}\cos\hat{\varphi}_{a}&=e^{-\frac{\eta_{a}}{4}}\sum\limits_{\begin{subarray}{c}m,n\geq 0\\ m+n=\text{ even}\end{subarray}}\frac{\left(-\frac{\eta_{a}}{2}\right)^{\frac{m+n}{2}}\hat{a}^{\dagger m}\hat{a}^{n}}{m!n!},\\ \sin\hat{\varphi}_{a}&=e^{-\frac{\eta_{a}}{4}}\sqrt{\frac{\eta_{a}}{2}}\sum\limits_{\begin{subarray}{c}m,n\geq 0\\ m+n=\text{ odd}\end{subarray}}\frac{\left(-\frac{\eta_{a}}{2}\right)^{\frac{m+n-1}{2}}\hat{a}^{\dagger m}\hat{a}^{n}}{m!n!},\end{split} (83)

with analogous expressions for the operators b^\hat{b} and c^\hat{c}.

D.2 Time-dependent terms in the coupler Hamiltonian

Terms corresponding to the Jacobi-Anger expansion up to the second harmonic of the drive in the bare coupler Hamiltonian H^~c(t)\widetilde{\hat{H}}_{c}(t) in Sec. IV are listed in Table 2. The operator monomial at the beginning of each row is to be multiplied by the sum of the two following columns, and then results from all rows are to be summed. The coefficients of the missing monomials 𝒄^,𝒄^2,𝒄^3,𝒄^𝒄^2,𝒄^4,𝒄^𝒄^3\hat{\bm{c}},\hat{\bm{c}}^{2},\hat{\bm{c}}^{3},\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{2},\hat{\bm{c}}^{4},\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{3} are obtained by Hermitian conjugation.

Monomial J1(δ)J_{1}(\delta) J2(δ)J_{2}(\delta)
𝒄^\hat{\bm{c}}^{\dagger} 2αϵeηc4ηcEJcJ1(δμα)sin(tωd)cos(μαφ¯ext)+\sqrt{2}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\sqrt{\eta_{c}}E_{\text{J}c}J_{1}\left(\delta\mu_{\alpha}\right)\sin\left(t\omega_{d}\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)+ 2αϵeηc4ηcEJcJ2(δμα)cos(2tωd)sin(μαφ¯ext)+\sqrt{2}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\sqrt{\eta_{c}}E_{\text{J}c}J_{2}\left(\delta\mu_{\alpha}\right)\cos\left(2t\omega_{d}\right)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)+
2βNeηc4N2ηcN2EJcJ1(δμβ)sin(tωd)cos(μβφ¯ext)\sqrt{2}\beta Ne^{-\frac{\eta_{c}}{4N^{2}}}\sqrt{\frac{\eta_{c}}{N^{2}}}E_{\text{J}c}J_{1}\left(\delta\mu_{\beta}\right)\sin\left(t\omega_{d}\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right) 2βNeηc4N2ηcN2EJcJ2(δμβ)cos(2tωd)sin(μβφ¯ext)\sqrt{2}\beta Ne^{-\frac{\eta_{c}}{4N^{2}}}\sqrt{\frac{\eta_{c}}{N^{2}}}E_{\text{J}c}J_{2}\left(\delta\mu_{\beta}\right)\cos\left(2t\omega_{d}\right)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)
𝒄^𝒄^\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger} 12αϵeηc4ηcEJcJ1(δμα)sin(tωd)sin(μαφ¯ext)-\frac{1}{2}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}E_{\text{J}c}J_{1}\left(\delta\mu_{\alpha}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right) 12αϵeηc4ηcEJcJ2(δμα)cos(2tωd)cos(μαφ¯ext)\frac{1}{2}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}E_{\text{J}c}J_{2}\left(\delta\mu_{\alpha}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)
βηceηc4N2EJcJ1(δμβ)sin(tωd)sin(μβφ¯ext)2N-\frac{\beta\eta_{c}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{1}\left(\delta\mu_{\beta}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{2N} +βηceηc4N2EJcJ2(δμβ)cos(2tωd)cos(μβφ¯ext)2N+\frac{\beta\eta_{c}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{2}\left(\delta\mu_{\beta}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{2N}
𝒄^𝒄^\hat{\bm{c}}^{\dagger}\hat{\bm{c}} αϵeηc4ηcEJcJ1(δμα)sin(tωd)sin(μαφ¯ext)-\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}E_{\text{J}c}J_{1}\left(\delta\mu_{\alpha}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right) +αϵeηc4ηcEJcJ2(δμα)cos(2tωd)cos(μαφ¯ext)+\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}E_{\text{J}c}J_{2}\left(\delta\mu_{\alpha}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)
βηceηc4N2EJcJ1(δμβ)sin(tωd)sin(μβφ¯ext)N-\frac{\beta\eta_{c}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{1}\left(\delta\mu_{\beta}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{N} +βηceηc4N2EJcJ2(δμβ)cos(2tωd)cos(μβφ¯ext)N+\frac{\beta\eta_{c}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{2}\left(\delta\mu_{\beta}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{N}
𝒄^𝒄^𝒄^\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger} αϵeηc4ηc3/2EJcJ1(δμα)sin(tωd)cos(μαφ¯ext)62-\frac{\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{3/2}E_{\text{J}c}J_{1}\left(\delta\mu_{\alpha}\right)\sin\left(t\omega_{d}\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)}{6\sqrt{2}} αϵeηc4ηc3/2EJcJ2(δμα)cos(2tωd)sin(μαφ¯ext)62-\frac{\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{3/2}E_{\text{J}c}J_{2}\left(\delta\mu_{\alpha}\right)\cos\left(2t\omega_{d}\right)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)}{6\sqrt{2}}
βηceηc4N2ηcN2EJcJ1(δμβ)sin(tωd)cos(μβφ¯ext)62N-\frac{\beta\eta_{c}e^{-\frac{\eta_{c}}{4N^{2}}}\sqrt{\frac{\eta_{c}}{N^{2}}}E_{\text{J}c}J_{1}\left(\delta\mu_{\beta}\right)\sin\left(t\omega_{d}\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{6\sqrt{2}N} βηceηc4N2ηcN2EJcJ2(δμβ)cos(2tωd)sin(μβφ¯ext)62N-\frac{\beta\eta_{c}e^{-\frac{\eta_{c}}{4N^{2}}}\sqrt{\frac{\eta_{c}}{N^{2}}}E_{\text{J}c}J_{2}\left(\delta\mu_{\beta}\right)\cos\left(2t\omega_{d}\right)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{6\sqrt{2}N}
𝒄^𝒄^𝒄^\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger}\hat{\bm{c}} αϵeηc4ηc3/2EJcJ1(δμα)sin(tωd)cos(μαφ¯ext)22-\frac{\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{3/2}E_{\text{J}c}J_{1}\left(\delta\mu_{\alpha}\right)\sin\left(t\omega_{d}\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)}{2\sqrt{2}} αϵeηc4ηc3/2EJcJ2(δμα)cos(2tωd)sin(μαφ¯ext)22-\frac{\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{3/2}E_{\text{J}c}J_{2}\left(\delta\mu_{\alpha}\right)\cos\left(2t\omega_{d}\right)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)}{2\sqrt{2}}
βηceηc4N2ηcN2EJcJ1(δμβ)sin(tωd)cos(μβφ¯ext)22N-\frac{\beta\eta_{c}e^{-\frac{\eta_{c}}{4N^{2}}}\sqrt{\frac{\eta_{c}}{N^{2}}}E_{\text{J}c}J_{1}\left(\delta\mu_{\beta}\right)\sin\left(t\omega_{d}\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{2\sqrt{2}N} βηceηc4N2ηcN2EJcJ2(δμβ)cos(2tωd)sin(μβφ¯ext)22N-\frac{\beta\eta_{c}e^{-\frac{\eta_{c}}{4N^{2}}}\sqrt{\frac{\eta_{c}}{N^{2}}}E_{\text{J}c}J_{2}\left(\delta\mu_{\beta}\right)\cos\left(2t\omega_{d}\right)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{2\sqrt{2}N}
𝒄^𝒄^𝒄^𝒄^\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger} 148αϵeηc4ηc2EJcJ1(δμα)sin(tωd)sin(μαφ¯ext)\frac{1}{48}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{2}E_{\text{J}c}J_{1}\left(\delta\mu_{\alpha}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right) 148αϵeηc4ηc2EJcJ2(δμα)cos(2tωd)cos(μαφ¯ext)-\frac{1}{48}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{2}E_{\text{J}c}J_{2}\left(\delta\mu_{\alpha}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)
+βηc2eηc4N2EJcJ1(δμβ)sin(tωd)sin(μβφ¯ext)48N3+\frac{\beta\eta_{c}^{2}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{1}\left(\delta\mu_{\beta}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{48N^{3}} βηc2eηc4N2EJcJ2(δμβ)cos(2tωd)cos(μβφ¯ext)48N3-\frac{\beta\eta_{c}^{2}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{2}\left(\delta\mu_{\beta}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{48N^{3}}
𝒄^𝒄^𝒄^𝒄^\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger}\hat{\bm{c}} 112αϵeηc4ηc2EJcJ1(δμα)sin(tωd)sin(μαφ¯ext)\frac{1}{12}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{2}E_{\text{J}c}J_{1}\left(\delta\mu_{\alpha}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right) 112αϵeηc4ηc2EJcJ2(δμα)cos(2tωd)cos(μαφ¯ext)-\frac{1}{12}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{2}E_{\text{J}c}J_{2}\left(\delta\mu_{\alpha}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)
+βηc2eηc4N2EJcJ1(δμβ)sin(tωd)sin(μβφ¯ext)12N3+\frac{\beta\eta_{c}^{2}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{1}\left(\delta\mu_{\beta}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{12N^{3}} βηc2eηc4N2EJcJ2(δμβ)cos(2tωd)cos(μβφ¯ext)12N3-\frac{\beta\eta_{c}^{2}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{2}\left(\delta\mu_{\beta}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{12N^{3}}
𝒄^𝒄^𝒄^𝒄^\hat{\bm{c}}^{\dagger}\hat{\bm{c}}^{\dagger}\hat{\bm{c}}\hat{\bm{c}} +18αϵeηc4ηc2EJcJ1(δμα)sin(tωd)sin(μαφ¯ext)+\frac{1}{8}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{2}E_{\text{J}c}J_{1}\left(\delta\mu_{\alpha}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right) 18αϵeηc4ηc2EJcJ2(δμα)cos(2tωd)cos(μαφ¯ext)-\frac{1}{8}\alpha\epsilon e^{-\frac{\eta_{c}}{4}}\eta_{c}^{2}E_{\text{J}c}J_{2}\left(\delta\mu_{\alpha}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\alpha}\overline{\varphi}_{\text{ext}}\right)
+βηc2eηc4N2EJcJ1(δμβ)sin(tωd)sin(μβφ¯ext)8N3+\frac{\beta\eta_{c}^{2}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{1}\left(\delta\mu_{\beta}\right)\sin\left(t\omega_{d}\right)\sin\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{8N^{3}} βηc2eηc4N2EJcJ2(δμβ)cos(2tωd)cos(μβφ¯ext)8N3-\frac{\beta\eta_{c}^{2}e^{-\frac{\eta_{c}}{4N^{2}}}E_{\text{J}c}J_{2}\left(\delta\mu_{\beta}\right)\cos\left(2t\omega_{d}\right)\cos\left(\mu_{\beta}\overline{\varphi}_{\text{ext}}\right)}{8N^{3}}
Table 2: Time-dependent terms, up to quartics, in the bare coupler Hamiltonian.

Appendix E Normal-mode transformation

In Sec. IV, we made use of a normal mode transformation that eliminates the off-diagonal capacitive coupling terms from the time-independent quadratic Hamiltonian. In this section we provide the steps to obtain the normal mode coefficients.

Consider the quadratic form (repeated indices are summed over):

H^=Aαβn¯^αn¯^β+Bαβφ¯^αφ¯^β.\displaystyle\hat{H}=A_{\alpha\beta}\hat{\bar{n}}_{\alpha}\hat{\bar{n}}_{\beta}+B_{\alpha\beta}\hat{\bar{\varphi}}_{\alpha}\hat{\bar{\varphi}}_{\beta}. (84)

We make a simplification by assuming that there are no off-diagonal inductive terms, BαβδαβB_{\alpha\beta}\propto\delta_{\alpha\beta}, which is valid for the circuit studied here. The diagonalization involves three steps:

Step 1. Rescale the variables so that the diagonal part of the Hamiltonian, the inductive part, contains terms with the same inductive energy. For this, let us define the square root of the product of the inductive energies B=(αBαα)1/2B=\left(\prod_{\alpha}B_{\alpha\alpha}\right)^{1/2} and the dimensionless coefficients fα=B/Bααf_{\alpha}=\sqrt{B/B_{\alpha\alpha}}. Then we introduce new canonically-conjugate coordinates:

φ^α=fα1φ¯^α,n^α=fαn¯^α.\displaystyle\hat{\varphi}_{\alpha}^{\prime}=f_{\alpha}^{-1}\hat{\bar{\varphi}}_{\alpha},\qquad\hat{n}_{\alpha}^{\prime}=f_{\alpha}\hat{\bar{n}}_{\alpha}. (85)

In terms of the new coordinates, and letting Aαβ=Aαβ/(fαfβ)A_{\alpha\beta}^{\prime}=A_{\alpha\beta}/(f_{\alpha}f_{\beta}) (no implicit summation), we have

H^=Aαβn^αn^β+Bδαβφ^αφ^β.\displaystyle\hat{H}=A^{\prime}_{\alpha\beta}\hat{n}_{\alpha}^{\prime}\hat{n}_{\beta}^{\prime}+B\delta_{\alpha\beta}\hat{\varphi}_{\alpha}^{\prime}\hat{\varphi}_{\beta}^{\prime}. (86)

Step 2. Diagonalize the capacitive coupling matrix AA^{\prime}. We assume here that this is possible and is achieved by an orthonormal matrix SS, such that

Aαβ=(ST)αμDμνSνβ=SμαDμνSνβ,\displaystyle A^{\prime}_{\alpha\beta}=(S^{T})_{\alpha\mu}D_{\mu\nu}S_{\nu\beta}=S_{\mu\alpha}D_{\mu\nu}S_{\nu\beta}, (87)

with DμνD_{\mu\nu} a diagonal matrix. Rewriting the above as Aαβ(ST)βγ=(ST)αμDμνSνβ(ST)βγ=(ST)αμDμγA^{\prime}_{\alpha\beta}(S^{T})_{\beta\gamma}=(S^{T})_{\alpha\mu}D_{\mu\nu}S_{\nu\beta}(S^{T})_{\beta\gamma}=(S^{T})_{\alpha\mu}D_{\mu\gamma}, or 𝐀𝐒T=𝐒T𝐃\mathbf{A}^{\prime}\cdot\mathbf{S}^{T}=\mathbf{S}^{T}\cdot\mathbf{D}, then the matrix 𝐒\mathbf{S} contains the eigenvectors of 𝐀\mathbf{A}^{\prime} on its rows. This diagonalization leads to

H^=SμαDμνSνβn^αn^β+Bδαβφ^αφ^β.\displaystyle\hat{H}=S_{\mu\alpha}D_{\mu\nu}S_{\nu\beta}\hat{n}_{\alpha}^{\prime}\hat{n}_{\beta}^{\prime}+B\delta_{\alpha\beta}\hat{\varphi}_{\alpha}^{\prime}\hat{\varphi}_{\beta}^{\prime}. (88)

Inspecting the first term, we again define new coordinates

n^μ′′=Sμαn^α,φ^μ′′=Sμαφ^α.\displaystyle\hat{n}_{\mu}^{\prime\prime}=S_{\mu\alpha}\hat{n}_{\alpha}^{\prime},\qquad\hat{\varphi}_{\mu}^{\prime\prime}=S_{\mu\alpha}\hat{\varphi}_{\alpha}^{\prime}. (89)

One can verify that the new double-primed coordinates are canonically conjugate because the transformation is orthonormal: [n^μ′′,φ^ν′′]=SμαSνβ[n^α,φ^β]=iSμαSνβδαβ=iSμαSνα=iδμν[\hat{n}_{\mu}^{\prime\prime},\hat{\varphi}_{\nu}^{\prime\prime}]=S_{\mu\alpha}S_{\nu\beta}[\hat{n}_{\alpha}^{\prime},\hat{\varphi}_{\beta}^{\prime}]=iS_{\mu\alpha}S_{\nu\beta}\delta_{\alpha\beta}=iS_{\mu\alpha}S_{\nu\alpha}=i\delta_{\mu\nu}. With this, we obtain a diagonal form for the Hamiltonian

H^=n^α′′Dαβn^β′′+Bδαβφ^α′′φ^β′′,\displaystyle\hat{H}=\hat{n}_{\alpha}^{\prime\prime}D_{\alpha\beta}\hat{n}_{\beta}^{\prime\prime}+B\delta_{\alpha\beta}\hat{\varphi}_{\alpha}^{\prime\prime}\hat{\varphi}_{\beta}^{\prime\prime}, (90)

where in the second term we used the fact that the orthogonal transformation preserves the inner product.

Step 3. Finally, we need to undo the rescaling transformation of Step 1. That is, introduce a third and last pair of canonically conjugate coordinates, the normal-mode coordinates

φ^α=fαφ^α′′,n^α=fα1n^α′′.\displaystyle\hat{\varphi}_{\alpha}=f_{\alpha}\hat{\varphi}_{\alpha}^{\prime\prime},\qquad\hat{n}_{\alpha}=f_{\alpha}^{-1}\hat{n}_{\alpha}^{\prime\prime}. (91)

At last the quadratic Hamiltonian reads

H^=n^αfαfβDαβn^β+Bδαβfαfβφ^αφ^β=n^αfαfβDαβn^β+φ^αBαβφ^β.\displaystyle\begin{split}\hat{H}&=\hat{n}_{\alpha}f_{\alpha}f_{\beta}D_{\alpha\beta}\hat{n}_{\beta}+\frac{B\delta_{\alpha\beta}}{f_{\alpha}f_{\beta}}\hat{\varphi}_{\alpha}\hat{\varphi}_{\beta}\\ &=\hat{n}_{\alpha}f_{\alpha}f_{\beta}D_{\alpha\beta}\hat{n}_{\beta}+\hat{\varphi}_{\alpha}B_{\alpha\beta}\hat{\varphi}_{\beta}.\end{split} (92)

This is the final normal-mode Hamiltonian.

Hybridization coefficients. It is helpful to summarize the normal mode transformation by skipping over the intermediate variables (primed, and double-primed). For this we have to invert the definitions of the intermediate coordinates to obtain

φ¯^α=βfαSβαfβ1φ^ββUαβφ^β,n¯^α=βfα1Sβαfβn^ββVαβn^β,\displaystyle\begin{split}\hat{\bar{\varphi}}_{\alpha}&=\sum_{\beta}f_{\alpha}S_{\beta\alpha}f_{\beta}^{-1}\hat{\varphi}_{\beta}\equiv\sum_{\beta}U_{\alpha\beta}\hat{\varphi}_{\beta},\\ \hat{\bar{n}}_{\alpha}&=\sum_{\beta}f_{\alpha}^{-1}S_{\beta\alpha}f_{\beta}\hat{n}_{\beta}\equiv\sum_{\beta}V_{\alpha\beta}\hat{n}_{\beta},\end{split} (93)

where we have used φ^α=Sμαφ^μ′′\hat{\varphi}^{\prime}_{\alpha}=S_{\mu\alpha}\hat{\varphi}_{\mu}^{\prime\prime} and n^α=Sμαn^μ′′\hat{n}^{\prime}_{\alpha}=S_{\mu\alpha}\hat{n}_{\mu}^{\prime\prime}. Note that 𝐔𝐕T=1\mathbf{U}\cdot\mathbf{V}^{T}=1, i.e. the transformation from bare to normal modes is canonical.

Creation and annihilation operators. Lastly, we consider the creation and annihiliation operators. In order for squeezing terms to disappear in the Hamiltonian, we need:

φ¯^α=β=a,b,cuαβ2(β^+β^),n¯^α=β=a,b,cvαβi2(β^β^),\displaystyle\begin{split}\hat{\bar{\varphi}}_{\alpha}&=\sum_{\beta=a,b,c}\frac{u_{\alpha\beta}}{\sqrt{2}}(\hat{\beta}+\hat{\beta}^{\dagger}),\\ \hat{\bar{n}}_{\alpha}&=\sum_{\beta=a,b,c}\frac{v_{\alpha\beta}}{i\sqrt{2}}(\hat{\beta}-\hat{\beta}^{\dagger}),\end{split} (94)

where

uαβ=Uαβϵβ,vαβ=Vαβϵβ.\displaystyle u_{\alpha\beta}=U_{\alpha\beta}\sqrt{\epsilon_{\beta}},\;v_{\alpha\beta}=\frac{V_{\alpha\beta}}{\sqrt{\epsilon_{\beta}}}. (95)

Finally, we have obtained the hybridization coefficients entering Eq. 50 in the main text. The approach given in this Appendix generalizes to an arbitrary number of modes with off-diagonal coupling in either the capacitive matrix, or in the inductive matrix.

Appendix F Floquet theory

This appendix provides a practical summary of Floquet theory. The spectrum of a monochromatically driven system can be obtained from the Floquet formalism Grifoni and Hänggi (1998), according to which the time-dependent Schrödinger equation for a periodically driven Hamiltonian H^(t)=H^(t+2π/ωd)\hat{H}(t)=\hat{H}(t+2\pi/\omega_{\text{d}}) can be recast into a numerically solvable eigenproblem for the so-called Floquet Hamiltonian Sambe (1973)

[H^(t)it]|ϕα(t)=ϵα|ϕα(t).\displaystyle\left[\hat{H}(t)-i\partial_{t}\right]\left|\phi_{\alpha}(t)\right\rangle=\epsilon_{\alpha}\left|\phi_{\alpha}(t)\right\rangle. (96)

The eigenvalues are the quasienergies ϵα\epsilon_{\alpha}, and whose eigenvectors are the Floquet modes which are periodic functions of time with |ϕα(t)=|ϕα(t+2π/ωd)\left|\phi_{\alpha}(t)\right\rangle=\left|\phi_{\alpha}(t+2\pi/\omega_{\text{d}})\right\rangle. In terms of these, the solution to the time-dependent Schrödinger is |ψα(t)=eiϵαt|ϕα(t)\left|\psi_{\alpha}(t)\right\rangle=e^{-i\epsilon_{\alpha}t}\left|\phi_{\alpha}(t)\right\rangle. Importantly, the solutions to Eq. 96 are only defined up to an integer multiple kk of the drive frequency ωd\omega_{\text{d}}, for if {ϵα,|ϕα(t)}\{\epsilon_{\alpha},\left|\phi_{\alpha}(t)\right\rangle\} is a solution, then so is {ϵαkϵα+kωd,|ϕαk(t)=eiωdt|ϕα(t)}\{\epsilon_{\alpha k}\equiv\epsilon_{\alpha}+k\omega_{\text{d}},\left|\phi_{\alpha k}(t)\right\rangle=e^{-i\omega_{\text{d}}t}\left|\phi_{\alpha}(t)\right\rangle\}, which is a consequence of the periodicity of the Floquet modes.

Information about the monochromatically driven system can be obtained from the quasienergy spectra. For example, two-tone spectroscopy experiments where a weak tone is used to probe the spectra of the driven system can be modeled in the linear response regime Verney et al. (2019). In such experiments, probe-tone-induced transitions occur at frequency differences

Δαβk=ϵαϵβ+kωd,\displaystyle\Delta_{\alpha\beta k}=\epsilon_{\alpha}-\epsilon_{\beta}+k\omega_{d}, (97)

provided that the operator corresponding to the probe tone, denoted generically as X^\hat{X}, has a nonzero matrix element between the corresponding Floquet modes. With the above notation, the corresponding matrix elements read

Xαβk=1T0T𝑑tei(ϵαϵβ+kωd)tϕβ(t)|X^|ϕα(t),\displaystyle X_{\alpha\beta k}=\frac{1}{T}\int_{0}^{T}dte^{-i(\epsilon_{\alpha}-\epsilon_{\beta}+k\omega_{d})t}\langle\phi_{\beta}(t)|\hat{X}|\phi_{\alpha}(t)\rangle, (98)

where T=2π/ωdT=2\pi/\omega_{\text{d}} is the period of the drive. This takes the form of a Fourier series coefficient fk=1T0T𝑑teik2πTtf(t)f_{k}=\frac{1}{T}\int_{0}^{T}dt^{\prime}e^{-ik\frac{2\pi}{T}t^{\prime}}f(t^{\prime}) of the matrix element of the operator XX between the two Floquet modes |ϕα,β(t)\left|\phi_{\alpha,\beta}(t)\right\rangle.

Numerically, the Floquet spectrum is efficiently obtained from the time-evolution operator over one period of the drive, which has a compact expression in terms of the Floquet modes Grifoni and Hänggi (1998)

U^(t+T,t)=𝒯eitt+TH^(t)𝑑t=αeiϵαT|ϕα(t)ϕα(t)|,\displaystyle\begin{split}\hat{U}(t+T,t)&=\mathcal{T}e^{-i\int_{t}^{t+T}\hat{H}(t^{\prime})dt^{\prime}}\\ &=\sum_{\alpha}e^{-i\epsilon_{\alpha}T}\left|\phi_{\alpha}(t)\right\rangle\left\langle\phi_{\alpha}(t)\right|,\end{split} (99)

where 𝒯\mathcal{T} is the time-ordering operator. According to the above expression, the Floquet modes at time t=0t=0, |ϕα(0)\left|\phi_{\alpha}(0)\right\rangle, are the eigenvectors of U(T,0)U(T,0), whereas the quasienergies are obtained modulo an integer multiple of ωd\omega_{\text{d}} from the eigenvalues. The time-dependence over one period of the drive is obtained by propagating each mode |ϕα(0)\left|\phi_{\alpha}(0)\right\rangle with the time-evolution operator U^(t,0)\hat{U}(t,0) in the interval 0<tT0<t\leq T.

Refer to caption
Figure 9: Floquet eigenspectrum for a rotating-wave approximation in which all photon-number nonconserving terms have been removed from the full circuit Hamiltonian analyzed in Sec. V.2. This figure is to be compared to the analogous result for the full Hamiltonian in Fig. 7.

To summarize, the steady-state dynamics can be obtained from the propagator U^(t,0)\hat{U}(t,0) over a single period of the drive, which makes the Floquet method an efficient alternative to numerical simulation of the dynamics over the complete gate time. Indeed, the period of the drive, on the order of 1ns1\,\text{ns} is between two to three orders of magnitude shorter than the typical gate times. In this work we obtain the quantities above by using the QuTip implementation of the Floquet formalism Johansson et al. (2013), to which we have contributed clc (2020), amended by a numerically efficient evaluation of the time-evolution operator developed by Shillito et al. (2020).

Appendix G Non-RWA effects in Floquet simulations of the full device

In this appendix we briefly discuss the role of counterrotating terms in the Floquet simulations of the full device Hamiltonian. Counterrotating terms (among which the parity-breaking cubic terms play an important role) in the coupler Hamiltonian induce an important correction to the coupler frequency, as can be seen by comparing Fig. 9 to Fig. 7. This indicates, among other things, that a mere approximation of the coupler Hamiltonian as a Kerr nonlinear oscillator, as was done in the case of the toy model, would be insufficient for precise comparisons with experimental data. Moreover, the speedup obtained by using the Floquet method, together with the numerically efficient method for computing the time-evolution operator, enables us to study non-RWA effects efficiently as compared to full time-dynamics.

References

  • Kjaergaard et al. (2020) M. Kjaergaard, M. E. Schwartz, J. Braumüller, P. Krantz, J. I.-J. Wang, S. Gustavsson,  and W. D. Oliver, Annual Review of Condensed Matter Physics 11, 369 (2020).
  • Krantz et al. (2019) P. Krantz, M. Kjaergaard, F. Yan, T. P. Orlando, S. Gustavsson,  and W. D. Oliver, Applied Physics Reviews 6, 021318 (2019).
  • Blais et al. (2021) A. Blais, A. L. Grimsmo, S. M. Girvin,  and A. Wallraff, Rev. Mod. Phys. 93, 025005 (2021).
  • Versluis et al. (2017) R. Versluis, S. Poletto, N. Khammassi, B. Tarasinski, N. Haider, D. J. Michalak, A. Bruno, K. Bertels,  and L. DiCarlo, Phys. Rev. Applied 8, 034021 (2017).
  • Arute et al. (2019) F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, R. Biswas, S. Boixo, F. G. S. L. Brandao, D. A. Buell, B. Burkett, Y. Chen, Z. Chen, B. Chiaro, R. Collins, W. Courtney, A. Dunsworth, E. Farhi, B. Foxen, A. Fowler, C. Gidney, M. Giustina, R. Graff, K. Guerin, S. Habegger, M. P. Harrigan, M. J. Hartmann, A. Ho, M. Hoffmann, T. Huang, T. S. Humble, S. V. Isakov, E. Jeffrey, Z. Jiang, D. Kafri, K. Kechedzhi, J. Kelly, P. V. Klimov, S. Knysh, A. Korotkov, F. Kostritsa, D. Landhuis, M. Lindmark, E. Lucero, D. Lyakh, S. Mandrà, J. R. McClean, M. McEwen, A. Megrant, X. Mi, K. Michielsen, M. Mohseni, J. Mutus, O. Naaman, M. Neeley, C. Neill, M. Y. Niu, E. Ostby, A. Petukhov, J. C. Platt, C. Quintana, E. G. Rieffel, P. Roushan, N. C. Rubin, D. Sank, K. J. Satzinger, V. Smelyanskiy, K. J. Sung, M. D. Trevithick, A. Vainsencher, B. Villalonga, T. White, Z. J. Yao, P. Yeh, A. Zalcman, H. Neven,  and J. M. Martinis, Nature 574, 505 (2019).
  • Corcoles et al. (2020) A. D. Corcoles, A. Kandala, A. Javadi-Abhari, D. T. McClure, A. W. Cross, K. Temme, P. D. Nation, M. Steffen,  and J. M. Gambetta, Proceedings of the IEEE 108, 1338 (2020).
  • Gottesman (1997) D. Gottesman, arXiv preprint quant-ph/9705052  (1997).
  • Aharonov and Ben-Or (2008) D. Aharonov and M. Ben-Or, SIAM J. Comput. 38, 1207 (2008).
  • Mundada et al. (2019) P. Mundada, G. Zhang, T. Hazard,  and A. Houck, Phys. Rev. Applied 12, 054023 (2019).
  • Ku et al. (2020) J. Ku, X. Xu, M. Brink, D. C. McKay, J. B. Hertzberg, M. H. Ansari,  and B. L. T. Plourde, Phys. Rev. Lett. 125, 200504 (2020).
  • Sung et al. (2020) Y. Sung, L. Ding, J. Braumüller, A. Vepsäläinen, B. Kannan, M. Kjaergaard, A. Greene, G. O. Samach, C. McNally, D. Kim, A. Melville, B. M. Niedzielski, M. E. Schwartz, J. L. Yoder, T. P. Orlando, S. Gustavsson,  and W. D. Oliver, “Realization of high-fidelity cz and zz-free iswap gates with a tunable coupler,”  (2020), arXiv:2011.01261 [quant-ph] .
  • Noguchi et al. (2020) A. Noguchi, A. Osada, S. Masuda, S. Kono, K. Heya, S. P. Wolski, H. Takahashi, T. Sugiyama, D. Lachance-Quirion,  and Y. Nakamura, “Fast parametric two-qubit gates with suppressed residual interaction using a parity-violated superconducting qubit,”  (2020), arXiv:2005.02630 [quant-ph] .
  • Kandala et al. (2020) A. Kandala, K. X. Wei, S. Srinivasan, E. Magesan, S. Carnevale, G. A. Keefe, D. Klaus, O. Dial,  and D. C. McKay, “Demonstration of a high-fidelity cnot for fixed-frequency transmons with engineered zz suppression,”  (2020), arXiv:2011.07050 [quant-ph] .
  • Zhao et al. (2020a) P. Zhao, D. Lan, P. Xu, G. Xue, M. Blank, X. Tan, H. Yu,  and Y. Yu, “Suppression of static zz interaction in an all-transmon quantum processor,”  (2020a), arXiv:2011.03976 [quant-ph] .
  • Berke et al. (2020) C. Berke, E. Varvelis, S. Trebst, A. Altland,  and D. P. DiVincenzo, “Transmon platform for quantum computing challenged by chaotic fluctuations,”  (2020), arXiv:2012.05923 [quant-ph] .
  • McKay et al. (2016) D. C. McKay, S. Filipp, A. Mezzacapo, E. Magesan, J. M. Chow,  and J. M. Gambetta, Phys. Rev. Applied 6, 064007 (2016).
  • Collodo et al. (2020) M. C. Collodo, J. Herrmann, N. Lacroix, C. K. Andersen, A. Remm, S. Lazar, J.-C. Besse, T. Walter, A. Wallraff,  and C. Eichler, “Implementation of conditional-phase gates based on tunable zz-interactions,”  (2020), arXiv:2005.08863 [quant-ph] .
  • Shirley (1965) J. H. Shirley, Phys. Rev. 138, B979 (1965).
  • Sambe (1973) H. Sambe, Phys. Rev. A 7, 2203 (1973).
  • Grifoni and Hänggi (1998) M. Grifoni and P. Hänggi, Physics Reports 304, 229 (1998).
  • Nigg et al. (2012) S. E. Nigg, H. Paik, B. Vlastakis, G. Kirchmair, S. Shankar, L. Frunzio, M. H. Devoret, R. J. Schoelkopf,  and S. M. Girvin, Physical Review Letters 108 (2012), 10.1103/physrevlett.108.240502.
  • Minev et al. (2020) Z. K. Minev, Z. Leghtas, S. O. Mundhada, L. Christakis, I. M. Pop,  and M. H. Devoret, “Energy-participation quantization of josephson circuits,”  (2020), arXiv:2010.00620 [quant-ph] .
  • Verney et al. (2019) L. Verney, R. Lescanne, M. H. Devoret, Z. Leghtas,  and M. Mirrahimi, Phys. Rev. Applied 11, 024003 (2019).
  • (24) A. Petrescu, B. Royer,  and A. Blais, in preparation .
  • Theis and Wilhelm (2017) L. S. Theis and F. K. Wilhelm, Physical Review A 95 (2017), 10.1103/physreva.95.022314.
  • Petrescu et al. (2020) A. Petrescu, M. Malekakhlagh,  and H. E. Türeci, Physical Review B 101 (2020), 10.1103/physrevb.101.134510.
  • Malekakhlagh et al. (2020a) M. Malekakhlagh, E. Magesan,  and D. C. McKay, Physical Review A 102 (2020a), 10.1103/physreva.102.042605.
  • Malekakhlagh et al. (2020b) M. Malekakhlagh, A. Petrescu,  and H. E. Türeci, Physical Review B 101 (2020b), 10.1103/physrevb.101.134509.
  • Koch et al. (2007) J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin,  and R. J. Schoelkopf, Phys. Rev. A 76, 042319 (2007).
  • Rigetti and Devoret (2010) C. Rigetti and M. Devoret, Physical Review B 81, 134507 (2010).
  • Chow et al. (2011) J. M. Chow, A. D. Córcoles, J. M. Gambetta, C. Rigetti, B. R. Johnson, J. A. Smolin, J. R. Rozen, G. A. Keefe, M. B. Rothwell, M. B. Ketchen,  and M. Steffen, Phys. Rev. Lett. 107, 080502 (2011).
  • Sheldon et al. (2016) S. Sheldon, E. Magesan, J. M. Chow,  and J. M. Gambetta, Physical Review A 93 (2016), 10.1103/physreva.93.060302.
  • Zhang et al. (2019) Y. Zhang, B. J. Lester, Y. Y. Gao, L. Jiang, R. J. Schoelkopf,  and S. M. Girvin, Phys. Rev. A 99, 012314 (2019).
  • Krinner et al. (2020a) S. Krinner, P. Kurpiers, B. Royer, P. Magnard, I. Tsitsilin, J. C. Besse, A. Remm, A. Blais,  and A. Wallraff, “Demonstration of an all-microwave controlled-phase gate between far detuned qubits,”  (2020a), arXiv:2006.10639 [quant-ph] .
  • Mundada et al. (2020) P. S. Mundada, A. Gyenis, Z. Huang, J. Koch,  and A. A. Houck, Phys. Rev. Applied 14, 054033 (2020).
  • Huang et al. (2021) Z. Huang, P. S. Mundada, A. Gyenis, D. I. Schuster, A. A. Houck,  and J. Koch, Phys. Rev. Applied 15, 034065 (2021).
  • You et al. (2007) J. Q. You, X. Hu, S. Ashhab,  and F. Nori, Phys. Rev. B 75, 140515 (2007).
  • Steffen et al. (2010) M. Steffen, F. Brito, D. DiVincenzo, M. Farinelli, G. Keefe, M. Ketchen, S. Kumar, F. Milliken, M. B. Rothwell, J. Rozen,  and R. H. Koch, Journal of Physics: Condensed Matter 22, 053201 (2010).
  • Yan et al. (2016) F. Yan, S. Gustavsson, A. Kamal, J. Birenbaum, A. P. Sears, D. Hover, T. J. Gudmundsen, D. Rosenberg, G. Samach, S. Weber, J. L. Yoder, T. P. Orlando, J. Clarke, A. J. Kerman,  and W. D. Oliver, Nature Communications 7, 12964 (2016).
  • Xu and Ansari (2020) X. Xu and M. H. Ansari, “Zz freedom in two qubit gates,”  (2020), arXiv:2009.00485 [quant-ph] .
  • Zhao et al. (2020b) P. Zhao, P. Xu, D. Lan, X. Tan, H. Yu,  and Y. Yu, Phys. Rev. Applied 14, 064016 (2020b).
  • Reynaud et al. (1997) S. Reynaud, E. Giacobino,  and J. Zinn-Justin, eds., Fluctuations quantiques : Les Houches, Session LXIII, 27 juin-28 juillet 1995 (Elsevier Amsterdam, New York, https://www.worldcat.org/title/fluctuations-quantiques-les-houches-session-lxiii-27-juin-28-juillet-1995-quantum-fluctuations/oclc/36407840, 1997).
  • Vool and Devoret (2017) U. Vool and M. Devoret, International Journal of Circuit Theory and Applications 45, 897 (2017).
  • You et al. (2019) X. You, J. A. Sauls,  and J. Koch, Phys. Rev. B 99, 174512 (2019).
  • Mirrahimi and Rouchon (2015) M. Mirrahimi and P. Rouchon,   (2015).
  • Goldman and Dalibard (2014) N. Goldman and J. Dalibard, Phys. Rev. X 4, 031027 (2014).
  • Krinner et al. (2020b) S. Krinner, S. Lazar, A. Remm, C. Andersen, N. Lacroix, G. Norris, C. Hellings, M. Gabureac, C. Eichler,  and A. Wallraff, Phys. Rev. Applied 14, 024042 (2020b).
  • Sete et al. (2021) E. A. Sete, A. Q. Chen, R. Manenti, S. Kulshreshtha,  and S. Poletto, “Floating tunable coupler for scalable quantum computing architectures,”  (2021), arXiv:2103.07030 [quant-ph] .
  • Abramowitz and Stegun (1964) M. Abramowitz and I. A. Stegun, Handbook of mathematical functions: with formulas, graphs, and mathematical tables, Vol. 55 (Courier Corporation, 1964).
  • Žitko (2011) R. Žitko, Computer Physics Communications 182, 2259 (2011).
  • Bertet et al. (2006) P. Bertet, C. J. P. M. Harmans,  and J. E. Mooij, Phys. Rev. B 73, 064512 (2006).
  • Niskanen et al. (2006) A. O. Niskanen, Y. Nakamura,  and J.-S. Tsai, Phys. Rev. B 73, 094506 (2006).
  • Niskanen et al. (2007) A. Niskanen, K. Harrabi, F. Yoshihara, Y. Nakamura, S. Lloyd,  and J. S. Tsai, Science 316, 723 (2007).
  • Beaudoin et al. (2012) F. Beaudoin, M. P. da Silva, Z. Dutton,  and A. Blais, Phys. Rev. A 86, 022305 (2012).
  • Reagor et al. (2018) M. Reagor, C. B. Osborn, N. Tezak, A. Staley, G. Prawiroatmodjo, M. Scheer, N. Alidoust, E. A. Sete, N. Didier, M. P. da Silva,  and et al., Science Advances 4, eaao3603 (2018).
  • Caldwell et al. (2018) S. A. Caldwell, N. Didier, C. A. Ryan, E. A. Sete, A. Hudson, P. Karalekas, R. Manenti, M. P. da Silva, R. Sinclair, E. Acala, N. Alidoust, J. Angeles, A. Bestwick, M. Block, B. Bloom, A. Bradley, C. Bui, L. Capelluto, R. Chilcott, J. Cordova, G. Crossman, M. Curtis, S. Deshpande, T. E. Bouayadi, D. Girshovich, S. Hong, K. Kuang, M. Lenihan, T. Manning, A. Marchenkov, J. Marshall, R. Maydra, Y. Mohan, W. O’Brien, C. Osborn, J. Otterbach, A. Papageorge, J.-P. Paquette, M. Pelstring, A. Polloreno, G. Prawiroatmodjo, V. Rawat, M. Reagor, R. Renzas, N. Rubin, D. Russell, M. Rust, D. Scarabelli, M. Scheer, M. Selvanayagam, R. Smith, A. Staley, M. Suska, N. Tezak, D. C. Thompson, T.-W. To, M. Vahidpour, N. Vodrahalli, T. Whyland, K. Yadav, W. Zeng,  and C. Rigetti, Phys. Rev. Applied 10, 034050 (2018).
  • Didier et al. (2018) N. Didier, E. A. Sete, M. P. da Silva,  and C. Rigetti, Phys. Rev. A 97, 022330 (2018).
  • Pranav S. Mundada, Sara Sussman et al. (2021) Pranav S. Mundada, Sara Sussman et al.in preparation  (2021).
  • Marcos et al. (2013) D. Marcos, P. Rabl, E. Rico,  and P. Zoller, Phys. Rev. Lett. 111, 110504 (2013).
  • Johansson et al. (2013) J. Johansson, P. Nation,  and F. Nori, Computer Physics Communications 184, 1234 (2013).
  • clc (2020) https://github.com/qutip/qutip/pull/1248  (2020).
  • Shillito et al. (2020) R. Shillito, J. A. Gross, A. D. Paolo, É. Genois,  and A. Blais, “Fast and differentiable simulation of driven quantum systems,”  (2020), arXiv:2012.09282 [quant-ph] .