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

Modular tunable coupler for superconducting circuits

Daniel L. Campbell [email protected] Air Force Research Laboratory, Information Directorate, Rome NY 13441 USA    Archana Kamal Department of Physics and Applied Physics, University of Massachusetts, Lowell, MA 01854 USA    Leonardo Ranzani Raytheon BBN Technologies, Cambridge MA 02138, USA    Michael Senatore Air Force Research Laboratory, Information Directorate, Rome NY 13441 USA Department of Physics, Syracuse University, Syracuse NY 13244-1130 USA    Matthew LaHaye [email protected] Air Force Research Laboratory, Information Directorate, Rome NY 13441 USA
Abstract

The development of modular and versatile quantum interconnect hardware is a key next step in the scaling of quantum information platforms to larger size and greater functionality. For superconducting quantum systems, fast and well-controlled tunable circuit couplers will be paramount for achieving high fidelity and resource efficient connectivity, whether for performing two-qubit gate operations, encoding or decoding a quantum data bus, or interfacing across modalities. Here we propose a versatile and internally-tunable double-transmon coupler (DTC) architecture that implements tunable coupling via flux-controlled interference in a three-junction dcSQUID. Crucially, the DTC possesses an internally defined zero-coupling state that is independent of the coupled data qubits or circuit resonators. This makes it particular attractive as a modular and versatile design element for realizing fast and robust linear coupling in several applications such as high-fidelity two-qubit gate operations, qubit readout, and quantum bus interfacing.

I Introduction

Recently, several demonstrations of noisy intermediate scale quantum NISQ processors involving complex systems comprising tens of coupled qubits, have cemented superconducting circuits as a leading platform for large scale quantum processing [1, 2, 3, 4, 5]. Further scaling of quantum processors and the development of distributed quantum architectures will likely benefit from reconfigurable qubit connectivity, tunable dispersive readout, bus interfacing, and transducer interconnects. For all these functionalities the use of a standalone circuit interconnect, a coupler, helps to preserve the coherence of the interconnected circuits while enabling rapid, high fidelity entangling operations between them.

Tunable couplers use an external control parameter to turn on and off an effective coupling geffg_{\textrm{eff}}. In superconducting circuits threading flux through a superconducting quantum interference device (SQuID) and applying microwave driving fields are examples of external control parameters. We highlight two broad approaches to tunable coupling: (i) couplers that use current-divider circuit elements [6, 7, 8, 9, 10, 11, 12, 13, 14, 15] (a recent well-known example is the gmon coupler [11, 12]), and (ii) couplers that interfere direct and virtual interaction pathways [16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 20, 26], such as the MIT-style single transmon coupler [16, 17, 18, 19, 20, 21, 22]. For the latter case, mutual inductance or capacitance between two qubits constitutes a direct interaction pathway while interactions mediated by coupler circuitry constitute a virtual interaction pathway. Between these two coupling approaches, current divider couplers respond more linearly with flux, lending themselves to parametric coupling applications [27, 15, 13]. On the other hand, inductively connecting the qubits introduces extra flux degrees of freedom which, by extension, lead to additional noise and decoherence channels, and crosstalk between local flux bias lines. A survey of coupler modalities cited in this paper, therefore, suggests that capacitive interactions lead to cleaner and potentially easier-to-control circuitry. Moreover, for circuit-QED architectures involving transmon qubits, capacitive interactions are compatible with the fixed-frequency designs routinely employed in high-coherence circuits. Therefore, on balance, capacitive rather than inductive coupling seems to provide a significant advantage for the interference- over current divider-style coupler modality especially for multi-qubit superconducting systems.

It is worthwhile to note that in the highlighted coupler approaches, the frequencies of the qubits and the magnitude of their interaction with the coupler determine the decoupling external flux bias Φe(0)\Phi_{e}^{(0)} that zeroes the qubit-qubit effective coupling geff=0g_{\textrm{eff}}=0; this necessitates circuit remodeling and optimization each time data qubits or architecture is even slightly modified. In addition, the quantum level structure of an interference coupler makes the realization of parametrically-driven exchange interactions challenging, an increasingly critical functionality explored in several demonstrations [28, 29, 30]. Motivated by these considerations several workarounds for parametric coupling, which do not involve modulating the transition frequency of the coupler itself, are gaining traction, such as modulation of the data qubit frequency itself in the presence of a fixed coupling [31, 32, 33, 34] and techniques that do not use flux driving [7, 35, 36].

In this paper, we describe and analyze a novel double transmon coupler (DTC) design, also explored in Refs. [37, 38], that provides the linearity to implement two-qubit parametric gates efficiently and possesses an internally defined zero-coupling state that is independent of the coupled data qubits or circuit resonators. Furthermore we introduce novel computationally friendly numerics and analytical expressions for the effective coupling of this circuit, as well as provide theoretical predictions for parametric and waveguide coupling use cases that are unique to this work.

Refer to caption
Figure 1: a) Prospective device layout. The coupler consists of a pair of transmon circuits (green) coupled by variably strong capacitance gCg_{C} and strong flux tunable inductance gL(Φ)g_{L}(\Phi). Each coupler transmon capacitively couples, in turn, to separate qubits (red and blue). {ωa,ω1,ω2,ωb}/2π\{\omega_{a},\omega_{1},\omega_{2},\omega_{b}\}/2\pi label transmon plasma frequencies in order of appearance, left to right. b) The equivalent circuit representation used for the analysis in this work [appendix A]. c) A flux-controlled interaction gCgLg_{C}-g_{L} turns on a hybridization between the coupler transmons. Solid lines represent hybridized coupler eigenenergies ω±\omega_{\pm}: the dark and light green lines show ω±\omega_{\pm} for capacitive coupling gC/2π=0g_{C}/2\pi=0 MHz and 123123 MHz, respectively. Average coupler energy, ω¯=(ω1+ω2)/2\bar{\omega}=(\omega_{1}+\omega_{2})/2, is represented with the dashed black line. The detuning of each qubit from the average energy is denoted with the detuning Δj=ωjω¯\Delta_{j}=\omega_{j}-\bar{\omega} for j{a,b}j\in\{a,b\}. d) Different coupling configurations realized for three distinct flux biases. For simplicity of presentation, we have assumed ω1=ω2\omega_{1}=\omega_{2} (i.e. δ=0\delta=0) in (c) and (d), for which the magnitude of the interaction between coupler transmons can be directly read off from the splitting between the |+|+\rangle and ||-\rangle eigenstates.

The DTC uses two inductively coupled transmon qubits to combine the linearity of current divider couplers with the capacitive coupler-qubit interactions of interference-style couplers at the cost of added level structure. From this combination of attributes, we predict fast, linear, and low noise tunable coupling for the DTC that is compatible with both fixed-frequency transmon architectures and parametric coupling use cases. The design paradigm works on a simple principle: each qubit capacitively interacts with a different transmon belonging to the coupler, as illustrated in Figs. 1a-b. These two coupler transmons in turn hybridize with an interaction that can be flux-tuned from negative to positive values. The qubits thus virtually interact via simultaneous coupling to the same hybridized states. Moreover, this net virtual interaction between the qubits can be turned off by flux-tuning the coupler transmon interaction and hybridization to zero.

Crucially, the addition of a second interposing transmon in the DTC compared to the interference approach suppresses the effective capacitance coupling between data qubits. The suppression geometrically scales with each interposing coupler transmon in the dispersive limit. By controllably hybridizing the DTC’s transmons, effective coupling comparable to that of a single interposing transmon is achievable, while preserving the increased data qubit isolation. Moreover, the junction and capacitor parameters of the DTC set the functional dependence of the DTC’s flux-tuning almost exclusively, making the decoupling flux bias Φe(0)\Phi_{e}^{(0)} depend only weakly on the transition frequencies of the qubits or qubit-DTC interaction strength (if at all). In summary, this hybridization style coupler isolates qubits in its ‘off’ configuration and its internal operations are insensitive to the frequencies and use case of the qubits to which it’s coupled: a desirable feature of modular design. We will also see that the additional transmon degree of freedom in the DTC can be leveraged to realize parametric coupling that mitigates deleterious nonlinear effects such as rectification in the presence of fast flux modulation.

II Double Transmon Coupler Model and Operation

The DTC circuit consists of two superconducting transmon qubits [39] sharing a common ground. The Josephson and charging energies for each transmon are denoted by EjE_{j} and ECjE_{Cj} for j{1,2}j\in\{1,2\}. A relatively high inductance junction E120.1×E1,2E_{12}\sim 0.1\times E_{1,2} connects the two transmons, forming a three junction dc SQuID, see Fig. 1b. A local bias line with an externally sourced current threads a tunable magnetic field through the SQuID, which tunes the inductive interaction between the transmons,

gL=4E12EC1EC22ω1ω2,g_{L}=\frac{4E_{12}^{\prime}\sqrt{E_{C1}E_{C2}}}{\hbar^{2}\sqrt{\omega_{1}\omega_{2}}}, (1)

where E12E12cos(Φe/φ0)E_{12}^{\prime}\sim E_{12}\cos{(\Phi_{e}/\varphi_{0})} is the external flux-dependent Josephson energy of the center junction [appendix A] and φ0=Φ0/2π\varphi_{0}=\Phi_{0}/2\pi is the reduced flux quantum. The plasma frequencies, ωj=(8(Ej+E12)ECj)1/2\omega_{j}=(8(E_{j}^{\prime}+E_{12}^{\prime})E_{Cj})^{1/2} also tune modestly with flux through flux-dependence of both E12E_{12}^{\prime} and EjEjE122sin2(Φe/φ0)/2EjE_{j}^{\prime}\sim E_{j}-E_{12}^{2}\sin^{2}{(\Phi_{e}/\varphi_{0})}/2E_{j}. In contrast, the capacitive interaction between the coupler transmons,

gCC12ω1ω22(C1+C12)(C2+C12),g_{C}\approx\frac{C_{12}\sqrt{\omega_{1}\omega_{2}}}{2\sqrt{(C_{1}+C_{12})(C_{2}+C_{12})}}, (2)

remains relatively constant as a function of flux. These inductive and capacitive contributions compete to define the net exchange interaction gCgL(Φe)g_{C}-g_{L}(\Phi_{e}), as shown in Fig. 1c [also see Eq. (13)].

To model and simulate this coupling approach, we now consider a chain of four transmon circuits. The outer transmons operate as data qubits with plasma frequencies {ωa,ωb}\{\omega_{a},\omega_{b}\}, while the inner transmons comprise the coupler with plasma frequencies {ω1,ω2}\{\omega_{1},\omega_{2}\}. Each qubit capacitively couples to a different coupler transmon, with strength gag_{a} and gbg_{b} respectively. The effective coupling between the data qubits may then be understood as a competition between virtual interactions mediated through the |+|+\rangle and ||-\rangle coupler eigenstates, as shown in Fig. 1(d). This allows tuning geffg_{\textrm{eff}} from positive to negative values as a nearly linear function of gCgL(Φe)g_{C}-g_{L}(\Phi_{e}), and guarantees the existence of an external flux Φe(0)\Phi_{e}^{(0)} such that geff=0g_{\textrm{eff}}=0. Anharmonicity in the DTC renormalizes geffg_{\textrm{eff}} at the 20%20\% level but otherwise the form for geffg_{\textrm{eff}} and Φe(0)\Phi_{e}^{(0)} is similar to that of coupled harmonic oscillators with the same connectivity. The equivalent harmonic oscillator Hamiltonian of circuit can then be represented as [see appendix B],

H/\displaystyle H/\hbar =j=a,b,1,2ωjajaj+(gCgL(Φe))(a1a2+a1a2)\displaystyle=\sum_{j=a,b,1,2}\omega_{j}a^{\dagger}_{j}a_{j}+(g_{C}-g_{L}(\Phi_{e}))(a_{1}^{\dagger}a_{2}+a_{1}a_{2}^{\dagger})
+ga(aaa1+aaa1)+gb(aba2+aba2),\displaystyle\quad+g_{a}(a_{a}^{\dagger}a_{1}+a_{a}a^{\dagger}_{1})+g_{b}(a_{b}^{\dagger}a_{2}+a_{b}a^{\dagger}_{2}), (3)

where aja^{\dagger}_{j} and aja_{j} are the raising and lowering operators, while the index jj identifies the respective qubit j={a,b}j=\{a,b\} or DTC j={1,2}j=\{1,2\} transmon modes.

To compute geffg_{\textrm{eff}} we first diagonalize the coupler part of the circuit, yielding ω±=ω¯±(δ2+(gCgL)2)1/2\omega_{\pm}=\bar{\omega}\pm(\delta^{2}+(g_{C}-g_{L})^{2})^{1/2} where 2ω¯=ω1+ω22\bar{\omega}=\omega_{1}+\omega_{2} and 2δ=ω1ω22\delta=\omega_{1}-\omega_{2} are the coupler transmon sum and difference plasma frequencies, respectively. Since {ω1,ω2}\{\omega_{1},\,\omega_{2}\} move together with flux, the common-mode frequency ω¯\bar{\omega} also tunes with flux, whereas δ\delta may tune only weakly, if at all. With δ=0\delta=0, the modulation of both gLg_{L} and ω¯\bar{\omega} are nearly equal so that one of the two coupler eigenmode plasma frequencies tunes only weakly with flux, as shown in Fig. 1c. In the other regime, when δ|gCgL|\delta\gg|g_{C}-g_{L}|, both coupler eigenmode plasma frequencies modulate similarly with flux, driven by flux tuning of ω¯\bar{\omega}.

II.1 Derivation of the effective coupling

We now consider the data qubit interaction with the coupler eigenmodes. To gain insight into the nature of the coupler operation, we will treat this under the dispersive approximation, i.e. gj/|ωjω±|<0.1g_{j}/|\omega_{j}-\omega_{\pm}|<0.1, where approximate analytical expression can be derived. To this end, the first step involves perturbatively decoupling the coupler eigenmodes from the qubit modes using Schrieffer-Wolff transformation of the form USW=exp[±j=a,bgjωjω±(aja±aja±)]U_{SW}=\text{exp}{\left[\sum_{\pm}\sum_{j=a,b}\frac{g_{j}}{\omega_{j}-\omega_{\pm}}(a^{\dagger}_{j}a_{\pm}-a_{j}a^{\dagger}_{\pm})\right]}. Retaining terms to second order in gj/(ωjω±)g_{j}/(\omega_{j}-\omega_{\pm}), and continuing to omit qubit and DTC anharmonicity, leads to the following effective two-qubit Hamiltonian,

H~/j=a,b(ωj+ωacj)ajaj+geff(aaab+aaab),\displaystyle\tilde{H}/\hbar\approxeq\sum_{j=a,b}(\omega_{j}+\omega_{\textrm{ac}}^{j})a_{j}^{\dagger}a_{j}+g_{\textrm{eff}}\left(a^{\dagger}_{a}a_{b}+a_{a}a^{\dagger}_{b}\right), (4)

with a simplified form of the effective qubit-qubit coupling and dispersive shifts on the qubits given by [see appendix C],

geffj=a,bgagb[(gCgL)+(gC+gL)(Δj/ω¯)]2Dj2,\displaystyle g_{\textrm{eff}}\approx\sum_{j=a,b}\frac{g_{a}g_{b}[(g_{C}-g_{L})+(g_{C}+g_{L})(\Delta_{j}/\bar{\omega})]}{2D_{j}^{2}}, (5)
ωacaga2[ωaω12Dj2],ωacbgb2[ωbω22Dj2],\displaystyle\omega_{\textrm{ac}}^{a}\approx g_{a}^{2}\Big{[}\frac{\omega_{a}-\omega_{1}}{2D_{j}^{2}}\Big{]},\quad\omega_{\textrm{ac}}^{b}\approx g_{b}^{2}\Big{[}\frac{\omega_{b}-\omega_{2}}{2D_{j}^{2}}\Big{]}, (6)

where Dj2=Δj2δ2(gCgL)2D_{j}^{2}=\Delta_{j}^{2}-\delta^{2}-(g_{C}-g_{L})^{2}. The flux dependence of ω¯\bar{\omega}, and therefore Δj=ωjω¯{\Delta_{j}=\omega_{j}-\bar{\omega}}, accounts for the non-sinusoidal tuning with flux shown in Fig. 2. Operating in the regimes where |ωaω1|,|ωbω2||gCgL|{|\omega_{a}-\omega_{1}|,\,|\omega_{b}-\omega_{2}|\gg|g_{C}-g_{L}|} and to a lesser extent when δ|gCgL|\delta\gg|g_{C}-g_{L}| results in more sinusoidal tuning. For the parameters used in figures of this paper, change in ωacj\omega_{\textrm{ac}}^{j} with flux is less than/equal to geffg_{\textrm{eff}}.

Refer to caption
Figure 2: The 2geff2g_{\textrm{eff}} between degenerate qubits, mediated by the coupler as a function of flux, changes if the coupler transitions are set above verses below the qubit transitions, all else being equal. The shading indicates the standard deviation from the mean after applying 6% relative Gaussian noise on all the coupler junctions, obtained using numerical simulation [Eq. (28)]. The numerical simulations find 2geff/2π2g_{\textrm{eff}}/2\pi by taking the minimum difference between the eigenfrequencies that best correspond to the states |1000\mathinner{|1000\rangle} and |0001\mathinner{|0001\rangle} as ωb\omega_{b} is swept past ωa\omega_{a}. A comparison of panel (a) versus panel (b) shows that adding capacitance shifts geffg_{\textrm{eff}} to more positive values. For each scenario, the approximate analytic geffg_{\textrm{eff}} is plotted in black [Eq. (55)]. At Φe(0)\Phi_{e}^{(0)}, the coupling parameters are given by g/2π=0.25 GHzg/2\pi=0.25\text{ GHz}, |Δ|/2π=1.1 GHz|\Delta|/2\pi=1.1\text{ GHz}, δ0\delta\approx 0, and the qubit plasma frequencies are ω/2π=4.7 GHz\omega/2\pi=4.7\text{ GHz}. The average junction parameters for (a) are E12/h=1.46 GHzE_{12}/h=1.46\text{ GHz} and E1/h,E2/h=7.5 GHzE_{1}/h,\,E_{2}/h=7.5\text{ GHz}, whereas for (b) they are E12/h=2.56 GHzE_{12}/h=2.56\text{ GHz} and E1/h,E2/h=23 GHzE_{1}/h,\,E_{2}/h=23\text{ GHz}. For all calculations, we chose E12/hE_{12}/h to set the maximum inductive coupling to gL/2π=0.31 GHzg_{L}/2\pi=0.31\text{ GHz}.

The externally applied flux (Φe\Phi_{e}) preferentially drops across the junction E12E_{12} which has the highest impedance (E12E1,E2E_{12}\ll E_{1},\,E_{2}): E12E_{12} therefore sets coupler hybridization via gL(Φe)g_{L}(\Phi_{e}). The shunt junctions E1E_{1} and E2E_{2} protect the coupler hybridization from charge noise, interaction with qubits, and other external circuitry. As a consequence, the decoupling flux bias, Φe(0)\Phi_{e}^{(0)} s.t. gCgL(Φe(0))geff(Φe(0))=0g_{C}-g_{L}(\Phi_{e}^{(0)})\propto g_{\textrm{eff}}(\Phi_{e}^{(0)})=0, is internally defined by the choice of coupler junction parameters. Figure 2 shows 2geff2g_{\textrm{eff}} as a function of flux Φe\Phi_{e}, for two choices of coupler transmon transitions, above (orange) or below the qubit (blue) transitions. The shaded region shows the gap between the qubit eigenfrequencies, numerically calculated using Eq. (28), where qubit bb is swept into degeneracy with qubit aa. These calculations were repeated thirty times with 6% Gaussian random variation applied independently to the three DTC junction parameters: the width of the shading shows the standard deviation in 2geff2g_{\textrm{eff}} at each Φe\Phi_{e}. Note the insensitivity of Φe(0)\Phi_{e}^{(0)} to variation in the junction parameters. By contrast, increasing gCg_{C} shifts Φe(0)\Phi_{e}^{(0)} further away from Φe/Φ0=0.5\Phi_{e}/\Phi_{0}=0.5 as shown by comparing Fig. 2a-b. The analytical expression for geffg_{\textrm{eff}} [Eq. (55)] matches numerical simulations to within several percent in the dispersive limit. Equation (55) contains additional terms that correspond to an approximately 20% correction to Eq. (5).

Refer to caption
Figure 3: Quantum state and coupling primitives for resonant, parametric, and dissipative use cases. a) A static coupling between nearly resonant qubits may be turned on with a static gLg_{L}. b) Parametric coupling between circuit elements with very different transition frequencies may be accomplished by modulating gLg_{L} at the frequency difference. The hybridization of coupler transmons modulates even when it’s transmon plasma frequencies differ. Coupling magnitude scales inversely with the relative detuning of each coupler transmon from its qubit. c) Purcell decay of a qubit through the coupler is turned off when gL=0g_{L}=0. This functionality allows waveform shaping of qubit emission into a waveguide.

II.2 Coupling nearly degenerate qubits

The independence of Φe(0)\Phi_{e}^{(0)} from the qubit plasma frequency allows for straightforward implementation of degenerate coupling strategies. As an example, consider two qubits, each dispersively coupled to a separate coupler transmon, as shown in Figs. 3a-b. To implement a controlled-Z operation, data qubit b is initialized at a frequency ωb\omega_{b} that is 500 MHz higher than ωa\omega_{a}; having some frequency detuning between the qubits helps to mitigate single-qubit microwave crosstalk. Meanwhile, the DTC flux bias is set to Φe=Φe(0)\Phi_{e}=\Phi_{e}^{(0)}. Next, by tuning the frequency of either qubit, the state |2000|2000\rangle is swept into degeneracy with |1001|1001\rangle (Fig. 3a), where the kets label transmon modes in Fig. 1b from left to right. When the states approach degeneracy, Φe\Phi_{e} is ramped away from Φe(0)\Phi_{e}^{(0)}, turning on a static coupling. The system then evolves for the requisite dwell time at the resulting avoided level crossing. Subsequently, the inverse of the previous flux ramps is implemented to complete the controlled-Z operation [11, 16, 17, 20, 19]. Unlike interference couplers [16], Φe(0)\Phi_{e}^{(0)} remains a constant throughout the procedure, enhancing ease of use. A procedure for using a DTC to implement controlled-Z between non-degenerate fixed-frequency qubits is also explored in Refs. [37, 38].

In a similar manner, it is also possible to implement iSWAP operations via standard approaches [40] utilizing the DTC SQuID. It should be noted, though, that even for evolution under the effective Hamiltonian in Eq. (4) – as is the case for the DTC – turning on geffg_{\textrm{eff}} to perform iSWAP between |1000|1000\rangle and |0001|0001\rangle likewise turns on parasitic σzσz\sigma_{z}\sigma_{z} phase evolution from interactions in the second excitation manifold |0002|1001|0002\rangle\leftrightarrow|1001\rangle, which must be separately cancelled [17, 41, 42, 43].

The DTC shares features with other three island transmon circuits [44] as well as the capacitively shunted flux (CSFQ) qubit [45]. Like the CSFQ, the anharmonicity of DTC transitions change as Φe\Phi_{e} is swept from zero flux to 0.5Φ00.5\,\Phi_{0}. In the E12E1,E2E_{12}\ll E_{1},\,E_{2} regime, the anharmonicity of the symmetric eigenstate transition of two strongly hybridized transmons ω1ω2\omega_{1}\approxeq\omega_{2} changes from weakly negative (transmon regime) to near zero. In Fig. 1c, the eigenenergy of the symmetric eigenstate can be seen to vary with Φe\Phi_{e}, while the small negative anharmonicity of the antisymmetric eigenstate remains constant with Φe\Phi_{e}. The DTC can eliminate both σxσx\sigma_{x}\sigma_{x} and σzσz\sigma_{z}\sigma_{z} interactions at the cancellation flux Φe=Φe(0)\Phi_{e}=\Phi_{e}^{(0)}, irrespective of qubit-qubit detuning. While such cancellations can be achieved in interference based couplers, this necessitates parameter fine-tuning and frequency crowding [37].

Refer to caption
Figure 4: Model parameters as a function of data transmon detuning defined in the a) lab frame and b) frame rotating with the parametric drive at ωd=ωbωa\omega_{d}=\omega_{b}-\omega_{a}, as defined in the main text. The square root denominator Dj=sign(Dj2)Dj2D_{j}=\textrm{sign}(D_{j}^{2})\sqrt{D_{j}^{2}} competes with the magnitude of ga,gb,gC, and gLg_{a},\,g_{b},\,g_{C},\,\textrm{ and }g_{L} to set the size of geffpg^{p}_{\textrm{eff}}. c) The lab frame parameters are chosen such that the rotating frame parameters are constant as a function of data transmon detuning: geffpg^{p}_{\textrm{eff}} [Eq. (61)] then also remains constant for iSWAP- and bSWAP-style exchange interactions. This picture breaks down in the shaded regions where accidental degeneracies arise between the data and coupler transmon transitions. The circular markers are resonant exchange rates obtained from 2D fits to the iSWAP or bSWAP time dependent Schrödinger equation (TDSE) simulations [Eq. (28)] performed for different choices of parametric drive frequency ωd/2π\omega_{d}/2\pi at fixed qubit-qubit detunings (ωbωa)/2π(\omega_{b}-\omega_{a})/2\pi. d) TDSE iSWAP simulations with (ωbωa)/2π(\omega_{b}-\omega_{a})/2\pi equal to 0.08 GHz0.08\text{ GHz} and 4.00 GHz4.00\text{ GHz} are shown in the (left) and (right) panels, respectively. The drive amplitude increases to its maximum value over 50 ns50\text{ ns} (not shown). The left panel shows substantial contributions from counter-rotating terms.

III Coupling with a parametric drive

We now discuss how the DTC can be employed to implement time-dependent parametric coupling. To this end, a parametric modulation of external flux Φe=(Asin(ωdt)+0.25)Φ0\Phi_{e}=(A\sin{(\omega_{d}t)}+0.25)\Phi_{0} can mediate Rabi-like exchange interactions with rate geffpg^{p}_{\textrm{eff}} between fixed-frequency qubits (or between other circuit elements) by tuning ωd\omega_{d} at resonance with the difference of relevant transition frequencies (Fig. 3b). The magnitude of resultant parametric coupling geffpg^{p}_{\textrm{eff}} may be understood in a rotating frame, obtained via the transformation (aap,a1p,abp,a2p)=(aa,a1,abeiωdt,a2eiωdt)(a_{a}^{p},\,a_{1}^{p},\,a_{b}^{p},\,a_{2}^{p})=(a_{a},\,a_{1},\,a_{b}e^{i\omega_{d}t},\,a_{2}e^{i\omega_{d}t}), where qubits aa and bb are rendered degenerate by the parametric driving [Fig. 4(a)]. Neglecting the fast rotating terms, the resulting stationary Hamiltonian is manifestly similar to Eq. (3) with effective coupling given by Eq. (5). Substituting ω2p=ω2ωd\omega_{2}^{p}=\omega_{2}-\omega_{d}, ωbp=ωbωd\omega_{b}^{p}=\omega_{b}-\omega_{d}, and ω¯p=ω¯ωd/2\bar{\omega}^{p}=\bar{\omega}-\omega_{d}/2 into Δjp=ωjpω¯p\Delta_{j}^{p}=\omega_{j}^{p}-\bar{\omega}^{p} and 2δjp=ω1pω2p2\delta_{j}^{p}=\omega_{1}^{p}-\omega_{2}^{p} with gLp=(2πA(2πA)3/8)gL(Φe=0)g_{L}^{p}=(2\pi A-(2\pi A)^{3}/8)g_{L}(\Phi_{e}=0), we obtain the parameters given in Fig. 4(b) and the effective parametric coupling strength as,

geffpj=a,bgagbgLp(1Δjp/ω¯p)4(Djp)2,\displaystyle g^{p}_{\textrm{eff}}\approx\sum_{j=a,b}\frac{-g_{a}g_{b}g_{L}^{p}(1-\Delta^{p}_{j}/\bar{\omega}^{p})}{4(D_{j}^{p})^{2}}, (7)

where (Djp)2=(Δjp)2(δp)2(gLp)2/2gC2(D_{j}^{p})^{2}=(\Delta_{j}^{p})^{2}-(\delta^{p})^{2}-(g_{L}^{p})^{2}/2-g_{C}^{2}. While the qubit detuning and pump frequency need to be swept in tandem to maintain resonance ωd=ωbωa\omega_{d}=\omega_{b}-\omega_{a}, the relative detuning between coupler transmon/qubit can be kept fixed, i.e. |ωaω1||ωbω2||\omega_{a}-\omega_{1}|\sim|\omega_{b}-\omega_{2}|. As a consequence, Δjp\Delta_{j}^{p} and δjp\delta_{j}^{p} also remain constant in the rotating frame as plotted in Fig. 4b. Adjusting ga,gb,andgLpg_{a},\,g_{b},\,\text{and}\,g_{L}^{p} to remain constant as other variables change [appendix A], geffpg^{p}_{\textrm{eff}} can be held constant as a function of flux. Under these circumstances numerical simulation confirm a constant geffpg^{p}_{\textrm{eff}}, as shown in Fig. 4b, in qualitative agreement with Eq. (7). While numerical simulations have included terms up to sixth order in the raising and lowering operators [Fig. 6], the analytical expression in Eq. (7) is derived using a simplified harmonic oscillator model [Eq. (4)] that neglects anharmonicity. However, the expression that generated the dotted line in Fig. 4c incorporates small corrections for the anharmonicity [Eq. (55)]. When the coupling is turned on, numerical simulations show 2\sqrt{2} stronger coupling within the second excitation manifold compared to within the first excitation manifold, qualitatively consistent with the simplified analytical model. The qualitative agreement between this model and numerical simulation implies that the anharmonicity of the coupler transmons does not play a strong role in setting geffg_{\textrm{eff}} or geffpg^{p}_{\textrm{eff}}.

Another consideration for parametric coupling is the rectification of data qubit frequency due to flux modulation. Specifically, ωacj\omega_{\textrm{ac}}^{j} of the driven qubits scales inversely with 1/(ωjω±)1/(\omega_{j}-\omega_{\pm}), which can change nonlinearly with flux; if the parametric drive periodically brings ωj\omega_{j} close to ω±\omega_{\pm}, it can introduce a time-averaged drive amplitude dependence to the Rabi resonance condition that manifests as distortion of the time-domain swap envelope. As shown by the simulations presented in Fig. 4c, such distortion remains negligible if the coupler-induced time-averaged frequency shifts are less than 0.1×geffp0.1\times g^{p}_{\textrm{eff}}. Relative to interference-based couplers, the DTC mitigates the deleterious impact of the nonlinear flux dependence shifts in two ways. First, the DTC eigenenergies tune gently with effective coupling. Second, the freedom to set the detuning and interaction rate gjg_{j} separately for each DTC transmon/qubit pair in fabrication allows the tuning of ωaca\omega_{\textrm{ac}}^{a} and ωacb\omega_{\textrm{ac}}^{b} with flux to be balanced. This balancing strategy roughly maximizes geffpg^{p}_{\textrm{eff}} for a fixed magnitude of nonlinearity. The combination of these factors allows straightforward, and relatively fast, pairwise parametric coupling between non-degenerate fixed-frequency qubits, among other applications. The parameters used in Figs. 2 and 4 represent a worst case scenario for second order nonlinearity generation, where δ0\delta\sim 0. Even so, parametric coupling of qubits with 8080 MHz relative detuning showed Rabi-like state evolution, such as that shown in Fig. 4d, persists as E12E_{12} increases 10%10\% (1%1\%) for the coupler transitions placed below (above) the qubit transitions.

A parametric drive can also induce unwanted direct energy exchange between various states in the Hilbert space, i.e. ωd|ωjω±|\omega_{d}\approx|\omega_{j}-\omega_{\pm}|, that should be avoided through appropriate choice of coupler transitions in fabrication. This is an important consideration for small qubit detuning. For larger qubit detunings, some driving frequencies will excite the DTC directly ωdω±\omega_{d}\sim\omega_{\pm}. A weak, undesirable, multi-photon exchange interactions will also occur in the second excited manifold |1001|0110\mathinner{|1001\rangle}\leftrightarrow\mathinner{|0110\rangle} when ω¯=(ωa+ωb)/2\bar{\omega}=(\omega_{a}+\omega_{b})/2: choosing the level structure as depicted in Figs. 4 and 3a-b avoids this undesirable degeneracy.

IV Coupling to a waveguide

The last use case that we consider is fast tunable coupling between a waveguide (or other open quantum system) and a qubit, as depicted in Fig. 3c. This use case is thematically consistent with previous demonstrations [28, 46]. For this scenario we make the left data qubit and the right DTC transmon degenerate, i.e. ωa=ω2\omega_{a}=\omega_{2}, which invalidates the dispersive approximation used in Eq. (4). Instead, the effective tunable coupling is given by ga(gCgL)/2δg_{a}(g_{C}-g_{L})/2\delta, which goes to zero when the flux is set to Φe(0)\Phi_{e}^{(0)}. If the degeneracy between the qubit and DTC transmon is broken, the qubit relaxation rate into the waveguide under the dispserive limit is given by Γga2(gCgL)2/Da4\Gamma g_{a}^{2}(g_{C}-g_{L})^{2}/D_{a}^{4}, which is typically small. Further, for ωaω2\omega_{a}\neq\omega_{2}, the use cases described in Figs. 3b-c can be combined to realize strong coupling to a continuum without demanding degeneracy between qubit and DTC transitions.

V Energy relaxation and coherence considerations

Our DTC layout (Fig. 1a) uses a grounded coplanar waveguide with impedance Z0=50 ΩZ_{0}=50\text{ }\Omega that is shorted near the three-junction SQuID loop: current in the waveguide induces a controllable flux Φe\Phi_{e} via a mutual inductance MM to the loop. The three junction DTC SQuID allows direct exchange coupling between the DTC states and current excitations in the flux bias line. Using the circuit model discussed in appendix D, the resultant uncorrelated relaxation rate of each DTC transmon into the bias line can be approximated as, Γ1,jM2E122cos2(Φe/φ0)/Z0φ04Cj\Gamma_{1,j}\approx M^{2}E_{12}^{2}\cos^{2}{(\Phi_{e}/\varphi_{0})}/Z_{0}\varphi_{0}^{4}C_{j} for j{1,2}j\in\{1,2\}, which only doubles if gCg_{C} is increased from zero to gCgLg_{C}\sim g_{L}: see the expression for ξj\xi_{j} and the derivation for Γ1,j\Gamma_{1,j} in the appendix D. Therefore, DTCs with small mutual inductive coupling M3 pHM\leq 3\text{ pH} likely will not need additional low pass filters on their current bias lines so long as the thermal population of the flux bias line is low. Conveniently, Γ1\Gamma_{1} also tunes to zero at gL=0g_{L}=0: thus, energy relaxation of the coupler into its flux bias line occurs only when geff0g_{\textrm{eff}}\neq 0 in the gCgLg_{C}\ll g_{L} limit.

Spontaneous emission of the DTC, whatever the underlying source of fluctuations, is a source of Purcell decay for the qubits whose rate can be approximated as, Γ1,a=(ga/Δa)2Γ1,1\Gamma_{1,a}=(g_{a}/\Delta_{a})^{2}\Gamma_{1,1} and Γ1,b=(gb/Δb)2Γ1,2\Gamma_{1,b}=(g_{b}/\Delta_{b})^{2}\Gamma_{1,2} where Γ1,j\Gamma_{1,j} for j{a,1,2,b}j\in\{a,1,2,b\} is the relaxation rate of the relevant transmon. Multi-junction transmon devices routinely have lifetimes in excess of tens of microseconds. Assuming T1>20 μsT_{1}>20\text{ }\mu\text{s} and a Purcell factor (gj/Δj)21/10(g_{j}/\Delta_{j})^{2}\sim 1/10, the coupler transmons impose at worst a 200 μs200\text{ }\mu\text{s} energy relaxation time on each qubit. While the dispersive regime is considered to be (gj/Δj)<0.1(g_{j}/\Delta_{j})<0.1, violating the dispersive approximation condition requires renormalization of (gj/Δj)(g_{j}/\Delta_{j}) and the inclusion of fourth order terms in the Schrieffer-Wolff transformation for quantitative accuracy [47] and accounts for the deviation between Eqs. (55) and (61) and numerical simulations in Figs. 2 and 4, respectively.

The impact of DTC-induced dephasing on the qubits can be computed from the frequency response of its eigenstates in presence of fluctuations on loop flux of the SQuID. The coupler eigenenergies change by at most 4gL4g_{L} over the full flux tuning range, with a maximum slope at the idling flux typically set to the deocupling bias Φe(0)\Phi_{e}^{(0)}. At the idling flux, the slope is then |ω/Φ|=2gL/φ0|\partial\omega/\partial\Phi|=2g_{L}/\varphi_{0}. For typical amplitudes of flux noise spectral density, AΦ2.5μΦ0/Hz\sqrt{A_{\Phi}}\sim 2.5\mu\Phi_{0}/\sqrt{\textrm{Hz}}, the Hahn echo dephasing rate due to 1/f1/f flux noise can be estimated as ΓϕE=Aϕln(2)|ω/Φ|50×103 s1/2\Gamma^{E}_{\phi}=\sqrt{A_{\phi}\textrm{ln}(2)}|\partial\omega/\partial\Phi|\sim 50\times 10^{3}\text{ s}^{-1/2} [48]. The frequency fluctuations of the qubits due to flux- and critical-current noise of the coupler transmons are filtered by the Purcell factor (gj/Δj)2(g_{j}/\Delta_{j})^{2}. Hence for a Purcell factor of 1/101/10, 1/f1/f flux noise would limit the coherence of coupled qubit modes to 200 μs\sim 200\text{ }\mu\text{s} at Φe(0)\Phi_{e}^{(0)}.

In summary, we have theoretically explored the use of hybridization between two transmons, a DTC, as a mechanism for mediating tunable interactions between fixed-frequency qubits. The DTC design offers low noise, along with linear and easy-to-model qubit-qubit interactions. Crucially, the decoupling condition is defined by a flux bias that nulls the hybridization, making the decoupling condition almost entirely independent of the properties of the data qubit and qubit-DTC interaction. The extra design freedom afforded by having two coupler transmons rather than one also enables implementation of fast, parametric coupling to a fixed mode or continuum. Moreover, the DTC may be attractive as a standalone qubit for its gentle eigenfrequency tuning with flux, weak exchange coupling with its current bias line, and level structure. The last benefit may also enable novel qutrit-based superconducting circuit architectures [49, 50].

Acknowledgements.
This material is based upon work supported by the U.S. Department of Energy, Office of Science, under award number DE-SC0019461, and Air Force Office of Scientific Research (AFOSR) under grant FA9550-21-1-0151. DLC and ML would also like to acknowledge independent support from AFOSR. Any opinions, findings, and conclusions or recommendations expressed in this article are those of the authors and do not necessarily reflect the views of the Air Force Research Laboratory (AFRL).

Appendix A Circuit derivation for the coupler alone

Refer to caption
Figure 5: Coupler circuit comprising two transmons qubits arranged in a flux-tunable loop with a third junction (E12,C12)(E_{12},C_{12}).

We write down the Lagrangian for the circuit shown in Fig. 5 using standard circuit parametrization [51, 52, 53]

=\displaystyle\mathcal{L}= 12C1(Φ˙1+Φ˙e1)2+12C12(Φ˙2Φ˙1+Φ˙e12)2\displaystyle\frac{1}{2}C_{1}(\dot{\Phi}_{1}+\dot{\Phi}_{e1})^{2}+\frac{1}{2}C_{12}(\dot{\Phi}_{2}-\dot{\Phi}_{1}+\dot{\Phi}_{e12})^{2}
+12C2(Φ˙2Φ˙e2)2\displaystyle\,+\frac{1}{2}C_{2}(\dot{\Phi}_{2}-\dot{\Phi}_{e2})^{2}
+E1cos((Φ1+Φe1)/φ0)+E2cos((Φ2Φe2)/φ0)\displaystyle\,+E_{1}\cos{((\Phi_{1}+\Phi_{e1})/\varphi_{0})}+E_{2}\cos{((\Phi_{2}-\Phi_{e2})/\varphi_{0})}
+E12cos((Φ2Φ1+Φe12)/φ0),\displaystyle\,+E_{12}\cos{((\Phi_{2}-\Phi_{1}+\Phi_{e12})/\varphi_{0})}, (8)

where we use φ0\varphi_{0} to denote the reduced flux quantum Φ0/2π\Phi_{0}/2\pi. Here Φj(t)\Phi_{j}(t) represent the node fluxes (a “position”-like variable) and Φ˙j(t)\dot{\Phi}_{j}(t) represent the node voltages (associated “velocities”) respectively. The loop of three Josephson junctions is threaded by an external flux, Φe=Φe12+Φe1+Φe2{\Phi_{e}=\Phi_{e12}+\Phi_{e1}+\Phi_{e2}}, with parameters {Φe12,Φe1,Φe2}\{\Phi_{e12},\,\Phi_{e1},\,\Phi_{e2}\} representing the external flux drops across the three junctions {E12,E1,E2}\{E_{12},\,E_{1},\,E_{2}\} respectively.

Performing the Legendre transformation, we arrive at the Hamiltonian,

H=\displaystyle H= C12(Q1+Q2)22C2+C2Q12+C1Q222C2\displaystyle\frac{C_{12}(Q_{1}+Q_{2})^{2}}{2C^{2}}+\frac{C_{2}Q_{1}^{2}+C_{1}Q_{2}^{2}}{2C^{2}}
+Φ˙e12(Q1C2C12+Q2C1C12)C2\displaystyle\,+\frac{\dot{\Phi}_{e12}(-Q_{1}C_{2}C_{12}+Q_{2}C_{1}C_{12})}{C^{2}}
+Φ˙e1(Q1(C1C2+C1C12)+Q2C1C12)C2\displaystyle\,+\frac{\dot{\Phi}_{e1}(Q_{1}(C_{1}C_{2}+C_{1}C_{12})+Q_{2}C_{1}C_{12})}{C^{2}}
Φ˙e2(Q1C1C12+Q2(C1C2+C1C12))C2\displaystyle\,-\frac{\dot{\Phi}_{e2}(Q_{1}C_{1}C_{12}+Q_{2}(C_{1}C_{2}+C_{1}C_{12}))}{C^{2}}
E1cos((Φ1+Φe1)/φ0)E2cos((Φ2Φe2)/φ0)\displaystyle\,-E_{1}\cos{((\Phi_{1}+\Phi_{e1})/\varphi_{0})}-E_{2}\cos{((\Phi_{2}-\Phi_{e2})/\varphi_{0})}
E12cos((Φ2Φ1+Φe12)/φ0),\displaystyle\,-E_{12}\cos{((\Phi_{2}-\Phi_{1}+\Phi_{e12})/\varphi_{0})}, (9)

where C2=C1C2+C1C12+C2C12C^{2}=C_{1}C_{2}+C_{1}C_{12}+C_{2}C_{12}. Using the loop constraint we may parametrize {Φe12,Φe1,Φe2}\{\Phi_{e12},\,\Phi_{e1},\,\Phi_{e2}\} in terms of Φe\Phi_{e} such that all Φ˙e\dot{\Phi}_{e} terms cancel [53]:

H=\displaystyle H= C12(Q1+Q2)22C2+C2Q12+C1Q222C2\displaystyle\frac{C_{12}(Q_{1}+Q_{2})^{2}}{2C^{2}}+\frac{C_{2}Q_{1}^{2}+C_{1}Q_{2}^{2}}{2C^{2}}
E1cos[(Φ1+C12C2C2Φe)/φ0]\displaystyle\,-E_{1}\cos{\left[\left(\Phi_{1}+\frac{C_{12}C_{2}}{C^{2}}\Phi_{e}\right)/\varphi_{0}\right]}
E2cos[(Φ2C12C1C2Φe)/φ0]\displaystyle\,-E_{2}\cos{\left[\left(\Phi_{2}-\frac{C_{12}C_{1}}{C^{2}}\Phi_{e}\right)/\varphi_{0}\right]}
E12cos[(Φ2Φ1+C1C2C2Φe)/φ0].\displaystyle\,-E_{12}\cos{\left[\left(\Phi_{2}-\Phi_{1}+\frac{C_{1}C_{2}}{C^{2}}\Phi_{e}\right)/\varphi_{0}\right]}. (10)

In this parameterization, we can infer that for typical capacitive coupling, C120.1C1C2C_{12}\leq 0.1\sqrt{C_{1}C_{2}}, the participation of E12E_{12} with Φe\Phi_{e} is greater than 80%80\%. Due to the lack of Φ˙e\dot{\Phi}_{e} terms, Eq. (10) is the simplest construction for charge basis simulations of the Hamiltonian.

Another useful parameterization redistributes Φe\Phi_{e} across the junctions to set H/Φj=0\partial H/\partial\Phi_{j}=0 for j={1, 2}j=\{1,\,2\}. This cancels odd order terms in Φj\Phi_{j} and thereby enables a straightforward conversion to the harmonic oscillator basis. We assume Φj2Φ0\langle\Phi_{j}^{2}\rangle\ll\Phi_{0} and expand to fourth order in Φj\Phi_{j}

H\displaystyle H\approx C12(Q1+Q2)22C2+C2Q12+C1Q222C2\displaystyle\frac{C_{12}(Q_{1}+Q_{2})^{2}}{2C^{2}}+\frac{C_{2}Q_{1}^{2}+C_{1}Q_{2}^{2}}{2C^{2}}
+Φ˙e12(Q1C2C12+Q2C1C12)C2\displaystyle\,+\frac{\dot{\Phi}_{e12}(-Q_{1}C_{2}C_{12}+Q_{2}C_{1}C_{12})}{C^{2}}
+E12Φ˙e12E1(Q1(C1C2+C1C12)+Q2C1C12)C2\displaystyle\,+\frac{E^{\prime}_{12}\dot{\Phi}_{e12}}{E^{\prime}_{1}}\frac{(Q_{1}(C_{1}C_{2}+C_{1}C_{12})+Q_{2}C_{1}C_{12})}{C^{2}}
E12Φ˙e12E2(Q1C2C12+Q2(C1C2+C2C12))C2\displaystyle\,-\frac{E^{\prime}_{12}\dot{\Phi}_{e12}}{E^{\prime}_{2}}\frac{(Q_{1}C_{2}C_{12}+Q_{2}(C_{1}C_{2}+C_{2}C_{12}))}{C^{2}}
+E1(Φ122φ02Φ1424φ04)\displaystyle\,+E^{\prime}_{1}\left(\frac{\Phi_{1}^{2}}{2\varphi_{0}^{2}}-\frac{\Phi_{1}^{4}}{24\varphi_{0}^{4}}\right)
+E2(Φ222φ02Φ2424φ04)\displaystyle\,+E^{\prime}_{2}\left(\frac{\Phi_{2}^{2}}{2\varphi_{0}^{2}}-\frac{\Phi_{2}^{4}}{24\varphi_{0}^{4}}\right)
+E12((Φ2Φ1)22φ02(Φ2Φ1)424φ04).\displaystyle\,+E^{\prime}_{12}\left(\frac{(\Phi_{2}-\Phi_{1})^{2}}{2\varphi_{0}^{2}}-\frac{(\Phi_{2}-\Phi_{1})^{4}}{24\varphi_{0}^{4}}\right). (11)

The external flux is then

Φe=Φe12+\displaystyle\Phi_{e}=\Phi_{e12}+ φ0arcsin((E12/E1)sin(Φe12/φ0))\displaystyle\varphi_{0}\arcsin{((E_{12}/E_{1})\sin{(\Phi_{e12}/\varphi_{0})})}
+\displaystyle+ φ0arcsin((E12/E2)sin(Φe12/φ0)),\displaystyle\varphi_{0}\arcsin{((E_{12}/E_{2})\sin{(\Phi_{e12}/\varphi_{0})})}, (12)

which may be numerically inverted. The junction energies are now flux sensitive: E12=E12cos(Φe12/φ0)E^{\prime}_{12}=E_{12}\cos{(\Phi_{e12}/\varphi_{0})} and Ej=Ej2E122sin2(Φe12/φ0)E^{\prime}_{j}=\sqrt{E_{j}^{2}-E_{12}^{2}\sin^{2}{(\Phi_{e12}/\varphi_{0})}}.

A.1 Harmonic oscillator basis

We convert to the Harmonic oscillator basis by substituting Qj=ωjCSj/2(aj+aj){Q_{j}=\sqrt{\hbar\omega_{j}C_{Sj}/2}(a_{j}+a_{j}^{\dagger})} and Φj=iωjLj/2(ajaj){\Phi_{j}=-i\sqrt{\hbar\omega_{j}L_{j}/2}(a_{j}-a_{j}^{\dagger})} in Eq. (11). This leads to the following approximate Hamiltonian, with parameters defined in Table 1, as

term value
CS1C_{S1} C2C12+C2\frac{C^{2}}{C_{12}+C_{2}}
CS2C_{S2} C2C12+C1\frac{C^{2}}{C_{12}+C_{1}}
LjL_{j} φ02E12+Ej\frac{\varphi_{0}^{2}}{E^{\prime}_{12}+E^{\prime}_{j}}
ECjE_{Cj} e2/2CSje^{2}/2C_{Sj}
gCg_{C} C12ω1ω22(C1+C12)(C2+C12)\frac{C_{12}\sqrt{\omega_{1}\omega_{2}}}{2\sqrt{(C_{1}+C_{12})(C_{2}+C_{12})}}
gLg_{L} 4E12EC1EC22ω1ω2\frac{4E^{\prime}_{12}\sqrt{E_{C1}E_{C2}}}{\hbar^{2}\sqrt{\omega_{1}\omega_{2}}}
ξ1\xi_{1} (2(E1+E12))1/4[E1(E2+E12)C2C12+E2E12(C1C2+C1C12)]2EC21/4E1E2C2\frac{(2(E_{1}^{\prime}+E_{12}^{\prime}))^{1/4}[-E_{1}^{\prime}(E_{2}^{\prime}+E_{12}^{\prime})C_{2}C_{12}+E_{2}^{\prime}E_{12}^{\prime}(C_{1}C_{2}+C_{1}C_{12})]}{2E_{C2}^{1/4}E_{1}^{\prime}E_{2}^{\prime}C^{2}}
ξ2\xi_{2} (2(E2+E12))1/4[E2(E1+E12)C1C12E1E12(C1C2+C2C12)]2EC21/4E1E2C2\frac{(2(E_{2}^{\prime}+E_{12}^{\prime}))^{1/4}[E_{2}^{\prime}(E_{1}^{\prime}+E_{12}^{\prime})C_{1}C_{12}-E_{1}^{\prime}E_{12}^{\prime}(C_{1}C_{2}+C_{2}C_{12})]}{2E_{C2}^{1/4}E_{1}^{\prime}E_{2}^{\prime}C^{2}}
ωj\omega_{j} 8(Ej+E12)ECj\frac{\sqrt{8(E_{j}^{\prime}+E_{12}^{\prime})E_{Cj}}}{\hbar}
νj(4)\nu_{j}^{(4)} ECj12\frac{E_{Cj}}{12\hbar}
νj(6)\nu_{j}^{(6)} 2ECj360(ECjEj+E12)1/2\frac{\sqrt{2}E_{Cj}}{360\hbar}\left(\frac{E_{Cj}}{E_{j}^{\prime}+E_{12}^{\prime}}\right)^{1/2}
μk(4)\mu_{k}^{(4)} (4k)(1)kE1212(EC1E1+E12)k/4(EC2E2+E12)(4k)/4\binom{4}{k}\frac{(-1)^{k}E^{\prime}_{12}}{12\hbar}\left(\frac{E_{C1}}{E_{1}^{\prime}+E_{12}^{\prime}}\right)^{k/4}\left(\frac{E_{C2}}{E_{2}^{\prime}+E_{12}^{\prime}}\right)^{(4-k)/4}
μk(6)\mu_{k}^{(6)} (6k)(1)k2E12360(EC1E1+E12)k/4(EC2E2+E12)(6k)/4\binom{6}{k}\frac{(-1)^{k}\sqrt{2}E^{\prime}_{12}}{360\hbar}\left(\frac{E_{C1}}{E_{1}^{\prime}+E_{12}^{\prime}}\right)^{k/4}\left(\frac{E_{C2}}{E_{2}^{\prime}+E_{12}^{\prime}}\right)^{(6-k)/4}
Table 1: Parameter definitions
H/=\displaystyle H/\hbar= j=12[ωjajajνj(4)(ajaj)4νj(6)(ajaj)6]\displaystyle\sum_{j=1}^{2}\left[\omega_{j}a_{j}^{\dagger}a_{j}-\nu_{j}^{(4)}(a_{j}-a_{j}^{\dagger})^{4}-\nu_{j}^{(6)}(a_{j}-a_{j}^{\dagger})^{6}\right]
+gC(a1+a1)(a2+a2)+gL(a1a1)(a2a2)\displaystyle\,+g_{C}(a_{1}+a_{1}^{\dagger})(a_{2}+a_{2}^{\dagger})+g_{L}(a_{1}-a_{1}^{\dagger})(a_{2}-a_{2}^{\dagger})
+j=12ξjΦ˙e12Φ0(aj+aj)\displaystyle\,+\sum_{j=1}^{2}\frac{\xi_{j}\dot{\Phi}_{e12}}{\Phi_{0}}(a_{j}+a_{j}^{\dagger})
k=13μk(4)(a1a1)k(a2a2)(4k)\displaystyle\,-\sum_{k=1}^{3}\mu_{k}^{(4)}(a_{1}-a_{1}^{\dagger})^{k}(a_{2}-a_{2}^{\dagger})^{(4-k)}
k=13μk(6)(a1a1)k(a2a2)(6k).\displaystyle\,-\sum_{k=1}^{3}\mu_{k}^{(6)}(a_{1}-a_{1}^{\dagger})^{k}(a_{2}-a_{2}^{\dagger})^{(6-k)}. (13)

The terms fourth order in the raising and lowering operators modify the eigenenergies in the harmonic oscillator basis by several hundred MHz and account for the majority of the anharmonicity. The sixth order terms compensate for a residual few 10s of MHz of discrepancy between the charge basis and harmonic oscillator approaches to simulation. At sixth order in the raising and lowering operators the first and second excitation manifolds of eigenenergies for the two simulation approaches agree to within 1 MHz, shown together in Fig. 6.

Refer to caption
Figure 6: Comparison of the simulated eigenenergies found by a harmonic oscillator basis simulation [Eq. (10)] verses a charge basis simulation [13]. We chose parameters Ej=13 GHzE_{j}=13\textrm{ GHz}, E12=1.3 GHzE_{12}=1.3\textrm{ GHz}, Cj=100 fFC_{j}=100\textrm{ fF}, and C12=0 fFC_{12}=0\textrm{ fF}. We simulated 8 levels per qubit for the harmonic oscillator simulation and 14 Cooper pair levels in the charge simulation.

Appendix B Circuit derivation for coupler and data qubits

We now include the data qubit terms to the Lagrangian in Eq. (8), collect the capacitive terms between operators, and break out the time dependence of the flux

=\displaystyle\mathcal{L}= 12ΦTCˇΦ\displaystyle\frac{1}{2}\vec{\Phi}^{T}\check{C}\vec{\Phi}
+Φ˙1(C1Φ˙e1C12Φ˙e12)Φ˙2(C2Φ˙e2C12Φ˙e12)\displaystyle\,+\dot{\Phi}_{1}(C_{1}\dot{\Phi}_{e1}-C_{12}\dot{\Phi}_{e12})-\dot{\Phi}_{2}(C_{2}\dot{\Phi}_{e2}-C_{12}\dot{\Phi}_{e12})
+E1cos((Φ1+Φe1)/φ0)+E2cos((Φ2Φe2)/φ0)\displaystyle\,+E_{1}\cos{((\Phi_{1}+\Phi_{e1})/\varphi_{0})}+E_{2}\cos{((\Phi_{2}-\Phi_{e2})/\varphi_{0})}
+E12cos((Φ2Φ1+Φe12)/φ0)\displaystyle\,+E_{12}\cos{((\Phi_{2}-\Phi_{1}+\Phi_{e12})/\varphi_{0})}
+Eacos(Φa/φ0)+Ebcos(Φb/φ0),\displaystyle\,+E_{a}\cos{(\Phi_{a}/\varphi_{0})}+E_{b}\cos{(\Phi_{b}/\varphi_{0})}, (14)

where Φ=(Φ˙a,Φ˙1,Φ˙2,Φ˙b)T\vec{\Phi}=(\dot{\Phi}_{a},\,\dot{\Phi}_{1},\,\dot{\Phi}_{2},\,\dot{\Phi}_{b})^{T} and the capacitance matrix is given by,

(CaCa100Ca1C1C1200C12C2Cb200Cb2Cb).\displaystyle\begin{pmatrix}C_{a}&C_{a1}&0&0\\ C_{a1}&C_{1}&C_{12}&0\\ 0&C_{12}&C_{2}&C_{b2}\\ 0&0&C_{b2}&C_{b}\end{pmatrix}. (15)

Performing the Legendre transformation, we arrive at the Hamiltonian,

=\displaystyle\mathcal{H}= 12QTCˇ1Q\displaystyle\frac{1}{2}\vec{Q}^{T}\check{C}^{-1}\vec{Q}
+QTCˇ1(0C1Φ˙e1C12Φ˙e12C2Φ˙e2+C12Φ˙e120)\displaystyle\,+\vec{Q}^{T}\check{C}^{-1}\begin{pmatrix}0\\ C_{1}\dot{\Phi}_{e1}-C_{12}\dot{\Phi}_{e12}\\ -C_{2}\dot{\Phi}_{e2}+C_{12}\dot{\Phi}_{e12}\\ 0\end{pmatrix}
+E1cos((Φ1+Φe1)/φ0)+E2cos((Φ2Φe2)/φ0)\displaystyle\,+E_{1}\cos{((\Phi_{1}+\Phi_{e1})/\varphi_{0})}+E_{2}\cos{((\Phi_{2}-\Phi_{e2})/\varphi_{0})}
+E12cos((Φ2Φ1+Φe12)/φ0)\displaystyle\,+E_{12}\cos{((\Phi_{2}-\Phi_{1}+\Phi_{e12})/\varphi_{0})}
+Eacos(Φa/φ0)+Ebcos(Φb/φ0),\displaystyle\,+E_{a}\cos{(\Phi_{a}/\varphi_{0})}+E_{b}\cos{(\Phi_{b}/\varphi_{0})}, (16)

where QT=(Qa,Q1,Q2,Qb)\vec{Q}^{T}=(Q_{a},\,Q_{1},\,Q_{2},\,Q_{b}) and QjQ_{j} for j{a, 1, 2,b}j\in\{a,\,1,\,2,\,b\} are charge operators for the jj’th transmon. If C1aC_{1a}, C2bC_{2b}, and C12C_{12} are all non-zero, then direct capacitive coupling exists between all qubits.

Using the loop constraint we may parametrize {Φe12,Φe1,Φe2}\{\Phi_{e12},\,\Phi_{e1},\,\Phi_{e2}\} such that all Φ˙e\dot{\Phi}_{e} terms cancel [53]:

H=\displaystyle H= 12QTCˇ1Q\displaystyle\frac{1}{2}\vec{Q}^{T}\check{C}^{-1}\vec{Q}
E1cos[(Φ1+C12C2C2Φe)/φ0]\displaystyle\,-E_{1}\cos{\left[\left(\Phi_{1}+\frac{C_{12}C_{2}}{C^{2}}\Phi_{e}\right)/\varphi_{0}\right]}
E2cos[(Φ2C12C1C2Φe)/φ0]\displaystyle\,-E_{2}\cos{\left[\left(\Phi_{2}-\frac{C_{12}C_{1}}{C^{2}}\Phi_{e}\right)/\varphi_{0}\right]}
E12cos[(Φ2Φ1+C1C2C2Φe)/φ0].\displaystyle\,-E_{12}\cos{\left[\left(\Phi_{2}-\Phi_{1}+\frac{C_{1}C_{2}}{C^{2}}\Phi_{e}\right)/\varphi_{0}\right]}. (17)

Interestingly, this parameterization of the flux does not change from that of the coupler alone in Eq. (10). Due to the lack of Φ˙e\dot{\Phi}_{e} terms, Eq. (17) is the simplest construction for charge basis simulations of the full qubit and coupler Hamiltonian.

B.1 Approximations for harmonic oscillator basis

Another useful parameterization involves expanding the cosine terms in the Hamiltonian according to

cos(Φ+Φe)\displaystyle\cos(\Phi+\Phi_{e}) cos(Φ)cos(Φe)+sin(Φ)sin(Φe)\displaystyle\approx\cos(\Phi)\cos(\Phi_{e})+\sin(\Phi)\sin(\Phi_{e}) (18)
cos(Φ)cos(Φe)+Φsin(Φe)\displaystyle\approx\cos(\Phi)\cos(\Phi_{e})+\Phi\sin(\Phi_{e}) (19)

where we have assumed Φj2Φ0\langle\Phi_{j}^{2}\rangle\ll\Phi_{0}, which is appropriate in the transmon limit EJ/EC>50E_{J}/E_{C}>50.

If, after expanding in Φj\Phi_{j} and Φ˙j\dot{\Phi}_{j}, the largest terms are proportional to Φj2\Phi_{j}^{2} and Φ˙j2\dot{\Phi}_{j}^{2} the harmonic oscillator basis can require fewer states to be modeled relative to the charge basis while achieving equivalent numerical precision. In flux-sensitive systems this can require some extra work since the external flux must be distributed across the junctions so that terms linear in Φj\Phi_{j} cancel. Solving the following equations, separately for Φe1\Phi_{e1} and Φe2\Phi_{e2}, performs this cancellation:

E1sin(Φe1)\displaystyle E_{1}\sin(\Phi_{e1}) =E12sin(Φe12)\displaystyle=E_{12}\sin(\Phi_{e12}) (20)
E2sin(Φe2)\displaystyle E_{2}\sin(\Phi_{e2}) =E12sin(Φe12)\displaystyle=E_{12}\sin(\Phi_{e12})
Φe1\displaystyle\Phi_{e1} =arcsin(E12sin(Φe12)/E1)\displaystyle=\arcsin(E_{12}\sin(\Phi_{e12})/E_{1}) (21)
Φe2\displaystyle\Phi_{e2} =arcsin(E12sin(Φe12)/E2)\displaystyle=\arcsin(E_{12}\sin(\Phi_{e12})/E_{2})
Φ˙e1\displaystyle\dot{\Phi}_{e1} =E12Φ˙e12/E1\displaystyle=E_{12}^{\prime}\dot{\Phi}_{e12}/E_{1}^{\prime} (22)
Φ˙e2\displaystyle\dot{\Phi}_{e2} =E12Φ˙e12/E2\displaystyle=E_{12}^{\prime}\dot{\Phi}_{e12}/E_{2}^{\prime}
E12\displaystyle E_{12}^{\prime} =E12cos(Φe12/φ0)\displaystyle=E_{12}\cos(\Phi_{e12}/\varphi_{0}) (23)
E1\displaystyle E_{1}^{\prime} =E12E122sin2(Φe12/φ0)\displaystyle=\sqrt{E_{1}^{2}-E_{12}^{2}\sin^{2}(\Phi_{e12}/\varphi_{0})} (24)
E2\displaystyle E_{2}^{\prime} =E22E122sin2(Φe12/φ0)\displaystyle=\sqrt{E_{2}^{2}-E_{12}^{2}\sin^{2}(\Phi_{e12}/\varphi_{0})} (25)

Substituting into Eq. (16) and truncating at sixth order in Φj\Phi_{j} we arrive at a suitable starting point for a transformation to the harmonic oscillator basis:

H\displaystyle H\approx 12QTCˇ1Q\displaystyle\frac{1}{2}\vec{Q}^{T}\check{C}^{-1}\vec{Q} (26)
+QTCˇ1(0(C1E12/E1C12)Φ˙e12(C2E12/E2+C12)Φ˙e120)\displaystyle\,+\vec{Q}^{T}\check{C}^{-1}\begin{pmatrix}0\\ (C_{1}E_{12}^{\prime}/E_{1}^{\prime}-C_{12})\dot{\Phi}_{e12}\\ (-C_{2}E_{12}^{\prime}/E_{2}^{\prime}+C_{12})\dot{\Phi}_{e12}\\ 0\end{pmatrix}
+E1(Φ122φ02Φ1424φ04+Φ16720φ06)\displaystyle\,+E^{\prime}_{1}\left(\frac{\Phi_{1}^{2}}{2\varphi_{0}^{2}}-\frac{\Phi_{1}^{4}}{24\varphi_{0}^{4}}+\frac{\Phi_{1}^{6}}{720\varphi_{0}^{6}}\right)
+E2(Φ222φ02Φ2424φ04+Φ26720φ06)\displaystyle\,+E^{\prime}_{2}\left(\frac{\Phi_{2}^{2}}{2\varphi_{0}^{2}}-\frac{\Phi_{2}^{4}}{24\varphi_{0}^{4}}+\frac{\Phi_{2}^{6}}{720\varphi_{0}^{6}}\right)
+E12((Φ2Φ1)22φ02(Φ2Φ1)424φ04+(Φ2Φ1)6720φ06).\displaystyle\,+E^{\prime}_{12}\left(\frac{(\Phi_{2}-\Phi_{1})^{2}}{2\varphi_{0}^{2}}-\frac{(\Phi_{2}-\Phi_{1})^{4}}{24\varphi_{0}^{4}}+\frac{(\Phi_{2}-\Phi_{1})^{6}}{720\varphi_{0}^{6}}\right).

E1E_{1}^{\prime}, E2E_{2}^{\prime} and E12E_{12}^{\prime} are parametrized by Φe12\Phi_{e12}. Φe12\Phi_{e12} is related to the true externally applied flux Φe\Phi_{e} by the following relation

Φe=Φe12+\displaystyle\Phi_{e}=\Phi_{e12}+ φ0arcsin((E12/E1)sin(Φe12/φ0))\displaystyle\varphi_{0}\arcsin{((E_{12}/E_{1})\sin{(\Phi_{e12}/\varphi_{0})})}
+\displaystyle+ φ0arcsin((E12/E2)sin(Φe12/φ0)),\displaystyle\varphi_{0}\arcsin{((E_{12}/E_{2})\sin{(\Phi_{e12}/\varphi_{0})})}, (27)

which may be numerically inverted.

B.2 Hamiltonian for the full circuit in the harmonic oscillator basis

We convert to the Harmonic oscillator basis by substituting Qj=ωjCSj/2(aj+aj){Q_{j}=\sqrt{\hbar\omega_{j}C_{Sj}/2}(a_{j}+a_{j}^{\dagger})} and Φj=iωjLj/2(ajaj){\Phi_{j}=-i\sqrt{\hbar\omega_{j}L_{j}/2}(a_{j}-a_{j}^{\dagger})} in Eq. (26) for j{a, 1, 2,b}j\in\{a,\,1,\,2,\,b\}.

This leads to the following approximate Hamiltonian given in Eq. (28), with parameters defined in Table 2

term value
CSj,jC_{Sj,j} 1/Cj,j11/C^{-1}_{j,j}
LjL_{j} for j{1, 2}j\in\{1,\,2\} φ02/(E12+Ej)\varphi_{0}^{2}/(E^{\prime}_{12}+E^{\prime}_{j})
LjL_{j} for j{a,b}j\in\{a,\,b\} φ02/Ej\varphi_{0}^{2}/E_{j}
ZjZ_{j} Lj/CSj,j\sqrt{L_{j}/C_{Sj,j}}
ECjE_{Cj} e2/2CSj,je^{2}/2C_{Sj,j}
gCj,kg_{Cj,k} Cj,k1/2ZjZkC^{-1}_{j,k}/2\sqrt{Z_{j}Z_{k}}
gLg_{L} E12ZjZk/2φ02E_{12}^{\prime}\sqrt{Z_{j}Z_{k}}/2\varphi_{0}^{2}
ξj\xi_{j} Φ0[Cj,11(C1E12/E1C12)+Cj,21(C2E12/E2+C12)]/2Zj\begin{aligned} \Phi_{0}&\big{[}C^{-1}_{j,1}(C_{1}E_{12}^{\prime}/E_{1}^{\prime}-C_{12})\\ +&C^{-1}_{j,2}(-C_{2}E_{12}^{\prime}/E_{2}^{\prime}+C_{12})\big{]}/\sqrt{2\hbar Z_{j}}\end{aligned}
ωj\hbar\omega_{j} 8(Ej+E12)ECj\sqrt{8(E_{j}^{\prime}+E_{12}^{\prime})E_{Cj}}
νj(4)\nu_{j}^{(4)} ECj12\frac{E_{Cj}}{12\hbar}
νj(6)\nu_{j}^{(6)} 2ECj360(ECjEj+E12)1/2\frac{\sqrt{2}E_{Cj}}{360\hbar}\left(\frac{E_{Cj}}{E_{j}^{\prime}+E_{12}^{\prime}}\right)^{1/2}
μk(4)\mu_{k}^{(4)} (4k)(1)kE1212(EC1E1+E12)k/4(EC2E2+E12)(4k)/4\binom{4}{k}\frac{(-1)^{k}E^{\prime}_{12}}{12\hbar}\left(\frac{E_{C1}}{E_{1}^{\prime}+E_{12}^{\prime}}\right)^{k/4}\left(\frac{E_{C2}}{E_{2}^{\prime}+E_{12}^{\prime}}\right)^{(4-k)/4}
μk(6)\mu_{k}^{(6)} (6k)(1)k2E12360(EC1E1+E12)k/4(EC2E2+E12)(6k)/4\binom{6}{k}\frac{(-1)^{k}\sqrt{2}E^{\prime}_{12}}{360\hbar}\left(\frac{E_{C1}}{E_{1}^{\prime}+E_{12}^{\prime}}\right)^{k/4}\left(\frac{E_{C2}}{E_{2}^{\prime}+E_{12}^{\prime}}\right)^{(6-k)/4}
Table 2: Parameter definitions
H/=\displaystyle H/\hbar= j[ωjajajνj(4)(ajaj)4νj(6)(ajaj)6]\displaystyle\sum_{j}\left[\omega_{j}a_{j}^{\dagger}a_{j}-\nu_{j}^{(4)}(a_{j}-a_{j}^{\dagger})^{4}-\nu_{j}^{(6)}(a_{j}-a_{j}^{\dagger})^{6}\right]
+jkgCj,k(aj+aj)(ak+ak)\displaystyle\,+\sum_{j\neq k}g_{Cj,k}(a_{j}+a_{j}^{\dagger})(a_{k}+a_{k}^{\dagger})
+gL(a1a1)(a2a2)\displaystyle\,+g_{L}(a_{1}-a_{1}^{\dagger})(a_{2}-a_{2}^{\dagger})
+jξjΦ˙e12Φ0(aj+aj)\displaystyle\,+\sum_{j}\frac{\xi_{j}\dot{\Phi}_{e12}}{\Phi_{0}}(a_{j}+a_{j}^{\dagger})
=13μk(4)(a1a1)k(a2a2)(4k)\displaystyle\,-\sum_{\ell=1}^{3}\mu_{k}^{(4)}(a_{1}-a_{1}^{\dagger})^{k}(a_{2}-a_{2}^{\dagger})^{(4-k)}
=13μk(6)(a1a1)k(a2a2)(6k)\displaystyle\,-\sum_{\ell=1}^{3}\mu_{k}^{(6)}(a_{1}-a_{1}^{\dagger})^{k}(a_{2}-a_{2}^{\dagger})^{(6-k)} (28)

where the summations range over j,k{a, 1, 2,b}j,\,k\in\{a,\,1,\,2,\,b\}.

The inverted capacitance matrix results in effective capacitive coupling between all four transmons in the system. We represent the exact analytic form followed by three approximate examples

gj,k\displaystyle g_{j,k} =|As\j,s\k|ωjωk2|As\j,s\j||As\k,s\k|\displaystyle=\frac{|A_{s\backslash j,s\backslash k}|\sqrt{\omega_{j}\omega_{k}}}{2\sqrt{|A_{s\backslash j,s\backslash j}||A_{s\backslash k,s\backslash k}|}} (29)
ga,1\displaystyle g_{a,1} Ca1ωaω1/CS\displaystyle\approx C_{a1}\sqrt{\omega_{a}\omega_{1}}/C_{S} (30)
ga,2\displaystyle g_{a,2} Ca1C12ωaω2/CS2\displaystyle\approx C_{a1}C_{12}\sqrt{\omega_{a}\omega_{2}}/C_{S}^{2} (31)
ga,b\displaystyle g_{a,b} Ca1Cb2C12ωaωb/CS3\displaystyle\approx C_{a1}C_{b2}C_{12}\sqrt{\omega_{a}\omega_{b}}/C_{S}^{3} (32)

where s={a,1,2,b}s=\{a,1,2,b\}, AA is the capacitance matrix from the Lagrangian, |A||A| is the determinant of that matrix, and As\j,s\kA_{s\backslash j,s\backslash k} is the four transmon capacitance matrix with row jj and column kk deleted. In Eqs. (30-32), CSC_{S} is a stand-in for something with the approximate value of the shunt capacitance, such as CSjC_{Sj}. The approximate relations ga,1g_{a,1}, ga,2g_{a,2}, and ga,bg_{a,b} show that the coupling capacitance falls off geometrically as the coupling progresses from nearest neighbor to next-to-nearest neighbor and then to next-to-next-to-nearest neighbor coupling, respectively. Data transmons a and b are therefore well isolated from each other, capacitively, by the coupler circuit.

For notational simplicity we will use the shorthand in Table 3 in the following appendices.

term shorthand
gC1,2g_{C1,2} gCg_{C}
gCa,1g_{Ca,1} gag_{a}
gC2,bg_{C2,b} gbg_{b}
Table 3: Commonly used shorthand notation in the appendices

B.3 Simulation of parametric driving

We simulated the parametric driving that appears in Fig. 4 with Eq. (28). Every term in Eq. (28) is flux sensitive. Those parameters, defined in Table 2, are flux sensitive through the definitions of E12E_{12}^{\prime} and EjE_{j}^{\prime} [Eqs. (23-25)] and Φe12\Phi_{e12} [Eq. (27)]. We apply a 50 ns50\text{ ns} half cosine ramp to a flux drive amplitude AΦ0A\Phi_{0}. The drive Φe(t)=(Asin(ωdt)0.25)Φ0\Phi_{e}(t)=(A\sin{(\omega_{d}t)}-0.25)\Phi_{0} was then held constant for 160 ns160\text{ ns}. The qubit state’s time evolution was modeled using the qutip python package’s time-dependent mesolve ODE solver in the absence of dissipation [54, 55]. Realistic estimates for energy relaxation and decoherence for the proposed tunable coupler are provided in Section D.

Appendix C Analytical derivation of the effective coupling

To apply the standard machinery of the Schrieffer-Wolff tranformation to decouple the states of the coupler from the computational Hilbert space we must first diagonalize the coupler states. Exact analytic diagonalization at second order in raising and lowering operators (this neglects anharmonicity which first appears at fourth order in the raising and lowering operators) may be performed using the Bogoliubov transformation Ψ=TΨ\Psi=T\Psi^{\prime} where Ψ=(a1a1a2a2)\Psi^{\dagger}=\begin{pmatrix}a_{1}&a_{1}^{\dagger}&a_{2}&a_{2}^{\dagger}\end{pmatrix} is the untransformed “vector” of raising and lowering operators. Our goal is to analytically obtain ‘T’. To this end we define ΨKΨ\Psi^{\dagger}K\Psi where KK contains prefactors associated with the quadratic terms of the coupler Hamiltonian [Eq. (13)]

K\displaystyle K =(ω10gCgLgC+gL0ω1gC+gLgCgLgCgLgC+gLω20gC+gLgCgL0ω2).\displaystyle=\begin{pmatrix}\omega_{1}&0&g_{C}-g_{L}&g_{C}+g_{L}\\ 0&\omega_{1}&g_{C}+g_{L}&g_{C}-g_{L}\\ g_{C}-g_{L}&g_{C}+g_{L}&\omega_{2}&0\\ g_{C}+g_{L}&g_{C}-g_{L}&0&\omega_{2}\end{pmatrix}. (33)

Note that we use the shorthand notation introduced in Table 3. The idea is to perform a similarity transformation TT to diagonalize KK and thereby find the desired Bogoliubov transformation.

Interestingly, because [Ψi,Ψj]=(1)iδi,j[\Psi_{i}^{\dagger},\Psi_{j}]=(-1)^{i}\delta_{i,j}, a similarity transformation is likely to produce a set of transformed operators that no longer obey the bosonic commutation relations. We can fix this by identifying the transformation σ=(1)iδi,j\sigma=(-1)^{i}\delta_{i,j}, and applying it to the column vector [Ψi,σΨj]=σσ=δi,j[\Psi_{i}^{\dagger},\sigma\Psi_{j}]=\sigma\sigma=\delta_{i,j}, where the commutation relations are now uniform. A similarity transformation TT may be constructed

σTσσKT=T1σKT=σK\displaystyle\sigma T^{\dagger}\sigma\sigma KT=T^{-1}\sigma KT=\sigma K^{\prime} (34)

where KK^{\prime} is the desired diagonal form of KK and Ψ=TΨ\Psi=T\Psi^{\prime}, see Altland and Simons, page 72 [56]. Using this technique, the diagonal form of the coupler operators, given by Ψ\Psi^{\prime}, can be found by diagonalizing σK\sigma K to obtain the transformation TT.

The diagonal elements of KK^{\prime} are a redundant set of eigenenergies {ω±,ω±}\{\omega_{\pm},-\omega_{\pm}\} otherwise obtainable from the Hamiltonian of the system (again this assumes a perfectly linear system). Starting from Eq. (33) these are,

ω±\displaystyle\omega_{\pm} =2ω¯2ω1ω24gC12gL±2η,\displaystyle=\sqrt{2\bar{\omega}^{2}-\omega_{1}\omega_{2}-4g_{C12}g_{L}\pm 2\eta}, (35)

with η=δ2ω¯2+(gC+gL)2ω1ω24gCgLω¯2\eta=\sqrt{\delta^{2}\bar{\omega}^{2}+(g_{C}+g_{L})^{2}\omega_{1}\omega_{2}-4g_{C}g_{L}\bar{\omega}^{2}}.

An analytic form for the eigenstates is also required to obtain an effective coupling but there exists no short form expression for these. Inspection of Eq. (33) reveals that the terms proportional to gC+gLg_{C}+g_{L}, the counter-rotating terms, and the terms proportional to gCgLg_{C}-g_{L}, the co-rotating terms, interact with each other at second order in (gC±gL)/ωj(g_{C}\pm g_{L})/\omega_{j}. This weak interaction between co- and counter-rotating terms justifies separating the calculation of geffg_{\textrm{eff}} into two parts, greatly simplifying the analytic expression of the eigenstates. The contributions to geffg_{\textrm{eff}} are then added together to obtain a better approximation of the effective coupling.

C.1 Co-rotating terms

Neglecting terms proportional to gC+gLg_{C}+g_{L} in Eq. (33) allows a compact representation of the approximate eigenvectors TT, which we apply to derive an analytic form of the effective coupling as mediated by exchange interactions. The approximate eigenenergies are likewise simplified to

ω±=ω¯±δ2+(gCgL)2,\displaystyle\omega_{\pm}=\bar{\omega}\pm\sqrt{\delta^{2}+(g_{C}-g_{L})^{2}}, (36)

where ω¯=(ω1+ω2)/2\bar{\omega}=(\omega_{1}+\omega_{2})/2 and δ=(ω1ω2)/2\delta=(\omega_{1}-\omega_{2})/2. The corresponding eigenvectors now take the form

a\displaystyle a_{-} =1β2a1+βa2,\displaystyle=-\sqrt{1-\beta^{2}}a_{1}+\beta a_{2}, (37a)
a+\displaystyle a_{+} =βa1+1β2a2,\displaystyle=\beta a_{1}+\sqrt{1-\beta^{2}}a_{2}, (37b)

where β=A/(gCgL)2+A2{\beta=A/\sqrt{(g_{C}-g_{L})^{2}+A^{2}}}, with A=δ+δ2+(gCgL)2{A=\delta+\sqrt{\delta^{2}+(g_{C}-g_{L})^{2}}}. If δ=0\delta=0 then the coupler transmons are degenerate and therefore fully hybridized, leading to β=1/2\beta=1/\sqrt{2} as we might expect. Substituting the eigenoperators a±a_{\pm} for a1,2a_{1,2} in Eq. (13) and retaining terms to second order, we obtain

H\displaystyle H =H0+V\displaystyle=H_{0}+V (38)

with

H0/\displaystyle H_{0}/\hbar =ωaaaaa+ωbabab+ωaa+ω+a+a+,\displaystyle=\omega_{a}a_{a}^{\dagger}a_{a}+\omega_{b}a_{b}^{\dagger}a_{b}+\omega_{-}a_{-}^{\dagger}a_{-}+\omega_{+}a_{+}^{\dagger}a_{+}, (39a)
V/\displaystyle V/\hbar =(1β2ga+βgCa,2)(aa+aa)(a+a)\displaystyle=(-\sqrt{1-\beta^{2}}g_{a}+\beta g_{Ca,2})(a_{a}+a_{a}^{\dagger})(a_{-}+a_{-}^{\dagger})
+(βgb1β2gCb,1)(ab+ab)(a+a)\displaystyle\quad+(\beta g_{b}-\sqrt{1-\beta^{2}}g_{Cb,1})(a_{b}+a_{b}^{\dagger})(a_{-}+a_{-}^{\dagger})
+(βga+1β2gCa,2)(aa+aa)(a++a+)\displaystyle\quad+(\beta g_{a}+\sqrt{1-\beta^{2}}g_{Ca,2})(a_{a}+a_{a}^{\dagger})(a_{+}+a_{+}^{\dagger})
+(1β2gb+βgCb,1)(ab+ab)(a++a+).\displaystyle\quad+(\sqrt{1-\beta^{2}}g_{b}+\beta g_{Cb,1})(a_{b}+a_{b}^{\dagger})(a_{+}+a_{+}^{\dagger}). (39b)

We identify a Schrieffer-Wolff generator SS such that [H0,S]=V[H_{0},S]=V,

S/=\displaystyle S/\hbar= 1β2ga+βgCa2Δa(aaaaaa)\displaystyle\frac{-\sqrt{1-\beta^{2}}g_{a}+\beta g_{Ca2}}{\Delta_{a-}}(a_{a}^{\dagger}a_{-}-a_{a}a_{-}^{\dagger})
+\displaystyle+ βgb1β2gCb,1Δb(abaaba)\displaystyle\frac{\beta g_{b}-\sqrt{1-\beta^{2}}g_{Cb,1}}{\Delta_{b-}}(a_{b}^{\dagger}a_{-}-a_{b}a_{-}^{\dagger})
+\displaystyle+ βga+1β2gCa,2Δa+(aaa+aaa+)\displaystyle\frac{\beta g_{a}+\sqrt{1-\beta^{2}}g_{Ca,2}}{\Delta_{a+}}(a_{a}^{\dagger}a_{+}-a_{a}a_{+}^{\dagger})
+\displaystyle+ 1β2gb+βgCb,1Δb+(aba+aba+)\displaystyle\frac{\sqrt{1-\beta^{2}}g_{b}+\beta g_{Cb,1}}{\Delta_{b+}}(a_{b}^{\dagger}a_{+}-a_{b}a_{+}^{\dagger})
+\displaystyle+ 1β2ga+βgCa2Σa(aaaaaa)\displaystyle\frac{-\sqrt{1-\beta^{2}}g_{a}+\beta g_{Ca2}}{\Sigma_{a-}}(a_{a}^{\dagger}a_{-}^{\dagger}-a_{a}a_{-})
+\displaystyle+ βgb1β2gCb,1Σb(abaaba)\displaystyle\frac{\beta g_{b}-\sqrt{1-\beta^{2}}g_{Cb,1}}{\Sigma_{b-}}(a_{b}^{\dagger}a_{-}^{\dagger}-a_{b}a_{-})
+\displaystyle+ βga+1β2gCa,2Σa+(aaa+aaa+)\displaystyle\frac{\beta g_{a}+\sqrt{1-\beta^{2}}g_{Ca,2}}{\Sigma_{a+}}(a_{a}^{\dagger}a_{+}^{\dagger}-a_{a}a_{+})
+\displaystyle+ 1β2gb+βgCb,1Σb+(aba+aba+),\displaystyle\frac{\sqrt{1-\beta^{2}}g_{b}+\beta g_{Cb,1}}{\Sigma_{b+}}(a_{b}^{\dagger}a_{+}^{\dagger}-a_{b}a_{+}),

where Δj±=ωjω±=Δj±δ2+(gCgL)2\Delta_{j\pm}=\omega_{j}-\omega_{\pm}=\Delta_{j}\pm\sqrt{\delta^{2}+(g_{C}-g_{L})^{2}}, Σj±=ωj+ω±=Σj±δ2+(gCgL)2\Sigma_{j\pm}=\omega_{j}+\omega_{\pm}=\Sigma_{j}\pm\sqrt{\delta^{2}+(g_{C}-g_{L})^{2}}, Σj=ωj+ω¯\Sigma_{j}=\omega_{j}+\bar{\omega}, and as defined in the main text Δj=ωjω¯\Delta_{j}=\omega_{j}-\bar{\omega}. Then

H=eSHeS=H0+[S,V]/2+𝒪(V3)\displaystyle H^{\prime}=e^{S}He^{-S}=H_{0}+[S,V]/2+\mathcal{O}(V^{3}) (41)

is the leading order diagonalized Hamiltonian with the pre-factor of the lowest order term [S,V]/2[S,V]/2 setting the effective coupling

geffco=\displaystyle g_{\textrm{eff}}^{\textrm{co}}= (gagb+gCa,2gCb,1)β1β22\displaystyle\frac{(g_{a}g_{b}+g_{Ca,2}g_{Cb,1})\beta\sqrt{1-\beta^{2}}}{2}
×\displaystyle\times j=a,b[1Δj+1Δj+1Σj+1Σj]\displaystyle\sum_{j=a,b}\Big{[}\frac{1}{\Delta_{j+}}-\frac{1}{\Delta_{j-}}+\frac{1}{\Sigma_{j+}}-\frac{1}{\Sigma_{j-}}\Big{]}
+\displaystyle+ gCa,2gb2\displaystyle\frac{g_{Ca,2}g_{b}}{2}
×\displaystyle\times j=a,b[1β2Δj++β2Δj+1β2Σj++β2Σj]\displaystyle\sum_{j=a,b}\Big{[}\frac{1-\beta^{2}}{\Delta_{j+}}+\frac{\beta^{2}}{\Delta_{j-}}+\frac{1-\beta^{2}}{\Sigma_{j+}}+\frac{\beta^{2}}{\Sigma_{j-}}\Big{]}
+\displaystyle+ gCb,1ga2\displaystyle\frac{g_{Cb,1}g_{a}}{2}
×\displaystyle\times j=a,b[β2Δj++1β2Δj+β2Σj++1β2Σj]\displaystyle\sum_{j=a,b}\Big{[}\frac{\beta^{2}}{\Delta_{j+}}+\frac{1-\beta^{2}}{\Delta_{j-}}+\frac{\beta^{2}}{\Sigma_{j+}}+\frac{1-\beta^{2}}{\Sigma_{j-}}\Big{]}

We may further condense terms

geffco=\displaystyle g_{\textrm{eff}}^{\textrm{co}}= (gagb+gCa,2gCb,1)(gCgL)j=a,b[12Dj2+12Sj2]\displaystyle(g_{a}g_{b}+g_{Ca,2}g_{Cb,1})(g_{C}-g_{L})\sum_{j=a,b}\Big{[}\frac{1}{2D_{j}^{2}}+\frac{1}{2S_{j}^{2}}\Big{]}
+\displaystyle+ gCa,2gbj=a,b[ωjω22Dj2+ωjω22Sj2]\displaystyle g_{Ca,2}g_{b}\sum_{j=a,b}\Big{[}\frac{\omega_{j}-\omega_{2}}{2D_{j}^{2}}+\frac{\omega_{j}-\omega_{2}}{2S_{j}^{2}}\Big{]}
+\displaystyle+ gCb,1gaj=a,b[ωjω12Dj2+ωjω12Sj2]\displaystyle g_{Cb,1}g_{a}\sum_{j=a,b}\Big{[}\frac{\omega_{j}-\omega_{1}}{2D_{j}^{2}}+\frac{\omega_{j}-\omega_{1}}{2S_{j}^{2}}\Big{]} (43)
Dj2\displaystyle D_{j}^{2} =Δj2δ2(gCgL)2\displaystyle=\Delta_{j}^{2}-\delta^{2}-(g_{C}-g_{L})^{2} (44)
Sj2\displaystyle S_{j}^{2} =Σj2δ2(gCgL)2.\displaystyle=\Sigma_{j}^{2}-\delta^{2}-(g_{C}-g_{L})^{2}. (45)

The terms proportional to Sj2S_{j}^{-2} and gCa,2gCb,1g_{Ca,2}g_{Cb,1} are typically small and may be neglected.

The ac Stark shifts on the data qubits may be calculated similarly. Here we neglect the terms proportional to gCa,2g_{Ca,2} and gCb,1g_{Cb,1}

ωaac=ga22[\displaystyle\omega^{\textrm{ac}}_{a}=\frac{g_{a}^{2}}{2}\Big{[} β2Δa++1β2Δaβ2Σa+1β2Σa]\displaystyle\frac{\beta^{2}}{\Delta_{a+}}+\frac{1-\beta^{2}}{\Delta_{a-}}-\frac{\beta^{2}}{\Sigma_{a+}}-\frac{1-\beta^{2}}{\Sigma_{a-}}\Big{]}
ωaac=ga22[\displaystyle\omega^{\textrm{ac}}_{a}=\frac{g_{a}^{2}}{2}\Big{[} ωaω1Da2+ωaω1Sa2]\displaystyle\frac{\omega_{a}-\omega_{1}}{D_{a}^{2}}+\frac{-\omega_{a}-\omega_{1}}{S_{a}^{2}}\Big{]}
ωbac=gb22[\displaystyle\omega^{\textrm{ac}}_{b}=\frac{g_{b}^{2}}{2}\Big{[} 1β2Δb++β2Δb1β2Σb+β2Σb]\displaystyle\frac{1-\beta^{2}}{\Delta_{b+}}+\frac{\beta^{2}}{\Delta_{b-}}-\frac{1-\beta^{2}}{\Sigma_{b+}}-\frac{\beta^{2}}{\Sigma_{b-}}\Big{]}
ωbac=gb22[\displaystyle\omega^{\textrm{ac}}_{b}=\frac{g_{b}^{2}}{2}\Big{[} ωbω2Db2+ωbω2Sb2].\displaystyle\frac{\omega_{b}-\omega_{2}}{D_{b}^{2}}+\frac{-\omega_{b}-\omega_{2}}{S_{b}^{2}}\Big{]}.

C.2 Counter-rotating terms

Neglecting now the terms proportional to gCgLg_{C}-g_{L} in Eq. (33), the direct exchange interaction terms, we can calculate the approximate eigenenergies,

ω±=±δ+ω¯2(gC+gL)2,\displaystyle\omega_{\pm}=\pm\delta+\sqrt{\bar{\omega}^{2}-(g_{C}+g_{L})^{2}}, (47)

and the eigenvectors,

a+\displaystyle a_{+} =1+α2a1αa2,\displaystyle=\sqrt{1+\alpha^{2}}a_{1}-\alpha a_{2}^{\dagger}, (48a)
a\displaystyle a_{-} =1+α2a2αa1,\displaystyle=\sqrt{1+\alpha^{2}}a_{2}-\alpha a_{1}^{\dagger}, (48b)

with α=(gC+gL)/(gC+gL)2+D2{\alpha=(g_{C}+g_{L})/\sqrt{-(g_{C}+g_{L})^{2}+D^{2}}}, D=ω¯+ω¯2(gC+gL)2{D=\bar{\omega}+\sqrt{\bar{\omega}^{2}-(g_{C}+g_{L})^{2}}}. The coefficients of Eqs. (48a-48b) satisfy the bosonic commutation relations [a±,a±]=1[a_{\pm},a^{\dagger}_{\pm}]=1 and [a+,a]=0[a_{+},a_{-}]=0. Substituting the transformed operators for a1,2a_{1,2} in Eq. (13) leads to the interaction,

V/\displaystyle V/\hbar =αga(aa+aa)(a+a)\displaystyle=\alpha g_{a}(a_{a}+a_{a}^{\dagger})(a_{-}+a_{-}^{\dagger})
+1+α2gb(ab+ab)(a+a)\displaystyle\quad+\sqrt{1+\alpha^{2}}g_{b}(a_{b}+a_{b}^{\dagger})(a_{-}+a_{-}^{\dagger})
+1+α2ga(aa+aa)(a++a+)\displaystyle\quad+\sqrt{1+\alpha^{2}}g_{a}(a_{a}+a_{a}^{\dagger})(a_{+}+a_{+}^{\dagger})
+αgb(ab+ab)(a+a+),\displaystyle\quad+\alpha g_{b}(a_{b}+a_{b}^{\dagger})(a_{+}-a_{+}^{\dagger}), (49)

which can be transformed using the generator,

S/\displaystyle S/\hbar =αgaΔa(aaaaaa)+1+α2gbΔb(abaaba)\displaystyle=\frac{\alpha g_{a}}{\Delta_{a-}}(a_{a}^{\dagger}a_{-}-a_{a}a_{-}^{\dagger})+\frac{\sqrt{1+\alpha^{2}}g_{b}}{\Delta_{b-}}(a_{b}^{\dagger}a_{-}-a_{b}a_{-}^{\dagger})
+\displaystyle+ 1+α2gaΔa+(aaa+aaa+)+αgbΔb+(aba+aba+)\displaystyle\frac{\sqrt{1+\alpha^{2}}g_{a}}{\Delta_{a+}}(a_{a}^{\dagger}a_{+}-a_{a}a_{+}^{\dagger})+\frac{\alpha g_{b}}{\Delta_{b+}}(a_{b}^{\dagger}a_{+}-a_{b}a_{+}^{\dagger})
+\displaystyle+ αgaΣa(aaaaaa)+1+α2gbΣb(abaaba)\displaystyle\frac{\alpha g_{a}}{\Sigma_{a-}}(a_{a}^{\dagger}a_{-}^{\dagger}-a_{a}a_{-})+\frac{\sqrt{1+\alpha^{2}}g_{b}}{\Sigma_{b-}}(a_{b}^{\dagger}a_{-}^{\dagger}-a_{b}a_{-})
+\displaystyle+ 1+α2gaΣa+(aaa+aaa+)+αgbΣb+(aba+aba+),\displaystyle\frac{\sqrt{1+\alpha^{2}}g_{a}}{\Sigma_{a+}}(a_{a}^{\dagger}a_{+}^{\dagger}-a_{a}a_{+})+\frac{\alpha g_{b}}{\Sigma_{b+}}(a_{b}^{\dagger}a_{+}^{\dagger}-a_{b}a_{+}),

to obtain the effective qubit-qubit coupling,

geffcounter=α1+α2gagb2j=a,b[1Δj+1Δj+1Σj1Σj+]\displaystyle g_{\textrm{eff}}^{\textrm{counter}}=\frac{\alpha\sqrt{1+\alpha^{2}}g_{a}g_{b}}{2}\sum_{j=a,b}\left[\frac{1}{\Delta_{j-}}+\frac{1}{\Delta_{j+}}-\frac{1}{\Sigma_{j-}}-\frac{1}{\Sigma_{j+}}\right] (51)

Approximating α1+α2(gC+gL)/2ω¯\alpha\sqrt{1+\alpha^{2}}\approx(g_{C}+g_{L})/2\bar{\omega} we may further condense the terms in Eq. (51)

geffcounter=gagb(gC+gL)j=a,b[\displaystyle g_{\textrm{eff}}^{\textrm{counter}}=-g_{a}g_{b}(g_{C}+g_{L})\sum_{j=a,b}\Big{[} Δj/ω¯2Dj2Σj/ω¯2Sj2].\displaystyle\frac{\Delta_{j}/\bar{\omega}}{2D_{j}^{2}}-\frac{\Sigma_{j}/\bar{\omega}}{2S_{j}^{2}}\Big{]}. (52)

C.3 Effective coupling used in the main text

We approximate the effective coupling for four coupled harmonic oscillators (remember that all the previous analysis in this section neglects non-linearity) as the sum of the coupling contributions from the co- and counter-rotating terms geffHO=geffco+geffcounterg_{\textrm{eff}}^{\textrm{HO}}=g_{\textrm{eff}}^{\textrm{co}}+g_{\textrm{eff}}^{\textrm{counter}},

geffHO=j=a,b[\displaystyle g_{\textrm{eff}}^{\textrm{HO}}=\sum_{j=a,b}\Big{[} gagb[gCgL(gC+gL)Δjω¯]2Dj\displaystyle\frac{g_{a}g_{b}\left[g_{C}-g_{L}-(g_{C}+g_{L})\frac{\Delta_{j}}{\bar{\omega}}\right]}{2D_{j}}
+\displaystyle+ gCa,2gbj=a,bωjω22Dj2\displaystyle g_{Ca,2}g_{b}\sum_{j=a,b}\frac{\omega_{j}-\omega_{2}}{2D_{j}^{2}}
+\displaystyle+ gCb,1gaj=a,bωjω12Dj2]\displaystyle g_{Cb,1}g_{a}\sum_{j=a,b}\frac{\omega_{j}-\omega_{1}}{2D_{j}^{2}}\Big{]} (53)

We found geffHOg_{\textrm{eff}}^{\textrm{HO}} approximates the numerically determined coupling to several percent accuracy when νj(4),νj(6),μk(4),μk(6)=0\nu_{j}^{(4)},\,\nu_{j}^{(6)},\,\mu_{k}^{(4)},\,\mu_{k}^{(6)}=0

Our estimate of geffg_{\textrm{eff}} when νj(4),νj(6),μk(4),μk(6)0\nu_{j}^{(4)},\,\nu_{j}^{(6)},\,\mu_{k}^{(4)},\,\mu_{k}^{(6)}\neq 0 can be improved by incorporating the contribution of terms forth order in the raising and lowering operators of H in Eq. (28) at second order in the raising and lowering operators. We then re-scale terms previously defined in geffHOg_{\textrm{eff}}^{\textrm{HO}}.

ωa,b=ωa,b12νa,b\displaystyle\omega_{a,b}^{\prime}=\omega_{a,b}-12\nu_{a,b} (54a)
ω1,2=ω1,22μ2(4)12ν1,2\displaystyle\omega_{1,2}^{\prime}=\omega_{1,2}-2\mu_{2}^{(4)}-12\nu_{1,2} (54b)
ω¯=(ω1+ω2)/2\displaystyle\bar{\omega}^{\prime}=(\omega_{1}^{\prime}+\omega_{2}^{\prime})/2 (54c)
δ¯=(ω1ω2)/2\displaystyle\bar{\delta}^{\prime}=(\omega_{1}^{\prime}-\omega_{2}^{\prime})/2 (54d)
Δj=ωjω¯\displaystyle\Delta_{j}^{\prime}=\omega_{j}^{\prime}-\bar{\omega}^{\prime} (54e)
gL=gL3(μ1(4)+μ3(4))\displaystyle g_{L}^{\prime}=g_{L}-3(\mu_{1}^{(4)}+\mu_{3}^{(4)}) (54f)
Dj2=Δj2+δ2+(gCgL)2\displaystyle D_{j}^{2^{\prime}}=\Delta_{j}^{2^{\prime}}+\delta^{2^{\prime}}+(g_{C}-g_{L}^{\prime})^{2} (54g)

With these substitutions, we obtain the definition of the effective coupling used in Fig. 2 of the main text.

geff=\displaystyle g_{\textrm{eff}}= j=a,bgagb[gCgL(gC+gL)Δjω¯]2Dj\displaystyle\sum_{j=a,b}\frac{g_{a}g_{b}\left[g_{C}-g_{L}^{\prime}-(g_{C}+g_{L}^{\prime})\frac{\Delta_{j}^{\prime}}{\bar{\omega}^{\prime}}\right]}{2D_{j}^{\prime}}
+gCa,2gbj=a,bωjω22Dj2\displaystyle\quad+g_{Ca,2}g_{b}\sum_{j=a,b}\frac{\omega_{j}^{\prime}-\omega_{2}^{\prime}}{2D_{j}^{2^{\prime}}}
+gCb,1gaj=a,bωjω12Dj2\displaystyle\quad+g_{Cb,1}g_{a}\sum_{j=a,b}\frac{\omega_{j}^{\prime}-\omega_{1}^{\prime}}{2D_{j}^{2^{\prime}}} (55)

C.4 Effective parametric coupling used in the main text

For the purposes of calculating geffpg_{\textrm{eff}}^{p}, we evaluate all the parameters of Table 2 at Φe12=0.25Φ0\Phi_{e12}=-0.25\Phi_{0} except gCg_{C} and E12=E12cos(Φe12/φ0)E_{12}^{\prime}=E_{12}\cos{(\Phi_{e12}/\varphi_{0})} in gLg_{L}. We substitute Φe12=Asin(ωdt)Φ00.25Φ0\Phi_{e12}=A\sin{(\omega_{d}t)}\Phi_{0}-0.25\Phi_{0} into the latter term. ΦeΦe12\Phi_{e}\approx\Phi_{e12} is an approximation of Eq. (27) that is valid in the regime E12EjE_{12}\ll E_{j}. The drive flux offset 0.25Φ0-0.25\Phi_{0} effectively transforms the cosine into a sine, which we expand about small ‘A’

sin(AΦ0sin(ωdt)/φ0)\displaystyle\sin{(A\Phi_{0}\sin{(\omega_{d}t)}/\varphi_{0})}
2πAsin(ωdt)(2πA)3sin3(ωdt)/6\displaystyle\quad\quad\approx 2\pi A\sin{(\omega_{d}t)}-(2\pi A)^{3}\sin^{3}{(\omega_{d}t)}/6 (56)
(2πA(2πA)3/8)sin(ωdt)\displaystyle\quad\quad\approx(2\pi A-(2\pi A)^{3}/8)\sin{(\omega_{d}t)} (57)
sin3(ωdt)=34sin(ωdt)14sin(3ωdt).\displaystyle\sin^{3}{(\omega_{d}t)}=\frac{3}{4}\sin{(\omega_{d}t)}-\frac{1}{4}\sin{(3\omega_{d}t)}. (58)

Substituting into gLg_{L} we obtain for gLpg_{L}^{p}

gL\displaystyle g_{L} E12(2πA(2πA)3/8)sin(ωdt)Z1Z2/2φ02\displaystyle\approx E_{12}(2\pi A-(2\pi A)^{3}/8)\sin{(\omega_{d}t)}\sqrt{Z_{1}Z_{2}}/2\varphi_{0}^{2} (59)
gLp\displaystyle g_{L}^{p} (2πA(2πA)3/8)gL(0Φ0)/2.\displaystyle\approx(2\pi A-(2\pi A)^{3}/8)g_{L}(0\Phi_{0})/2. (60)

The terms proportional to sin(ωdt)\sin{(\omega_{d}t)} are large when they include gLg_{L} and μk(4)\mu_{k}^{(4)} in the numerator of Eq. (55). By contrast, gCg_{C} changes very little as a function of flux and may be set to zero. Most other parameters in Table 2 can be approximately modeled as their static value at Φe12=0.25Φ0\Phi_{e12}=-0.25\Phi_{0}. In the denominator we retained terms without time dependence. This motivates the substitution (gCgL)2(gLp)2/2+gC2(g_{C}-g_{L})^{2}\rightarrow(g_{L}^{p})^{2}/2+g_{C}^{2}.

geffpj=a,bgagbgLp(1Δjp/ω¯p)4(Djp)2.\displaystyle g^{p}_{\textrm{eff}}\approx\sum_{j=a,b}\frac{-g_{a}g_{b}g_{L}^{p}(1-\Delta^{p}_{j}/\bar{\omega}^{p})}{4(D_{j}^{p})^{2}}. (61)
(Djp)2=(Δjp)2(δp)2(gLp)2/2gC2\displaystyle(D_{j}^{p})^{2}=(\Delta_{j}^{p})^{2}-(\delta^{p})^{2}-(g_{L}^{p})^{2}/2-g_{C}^{2} (62)

Appendix D Noise analysis

D.1 Coherence

The Gaussian pure dephasing rate due to 1/f1/f flux noise is ΓϕE=AΦln2|ω/Φ|\Gamma_{\phi}^{E}=\sqrt{A_{\Phi}\text{ln}2}|\partial\omega/\partial\Phi| for Hahn echo measurements. It depends upon the noise amplitude AΦ\sqrt{A_{\Phi}}, defined at ω/2π=1 Hz\omega/2\pi=1\text{ Hz}, of the power spectral density S(ω)Φ=AΦ/|ω|S(\omega)_{\Phi}=A_{\Phi}/|\omega| and the slope of the energy dispersion with flux ω/Φ\hbar\partial\omega/\partial\Phi. Careful engineering gives a noise amplitude AΦ2.5μΦ0\sqrt{A_{\Phi}}\sim 2.5\mu\Phi_{0}.

On the coupler, the peak to peak difference in the frequency dispersion is 0.6 GHz\sim 0.6\text{ GHz}, giving ΓϕE2.5μΦ0ln(2)(2π)2×0.3 GHz/Φ0=1/40 μs\Gamma_{\phi}^{E}\approx 2.5\mu\Phi_{0}\sqrt{\text{ln}(2)}(2\pi)^{2}\times 0.3\text{ GHz}/\Phi_{0}=1/40\text{ }\mu\text{s}. The noise on geffg_{\textrm{eff}} from the flux sensitivity on gLg_{L} is then reduced by a factor gagb/(2Δj22δ2)g_{a}g_{b}/(2\Delta_{j}^{2}-2\delta^{2}). Similarly, frequency noise on the qubits from the flux sensitivity on ω¯\bar{\omega} is reduced by a factor of ga2/(2Δj22δ2)\sim g_{a}^{2}/(2\Delta_{j}^{2}-2\delta^{2}). For geff=2π×60 MHzg_{\textrm{eff}}=2\pi\times 60\text{ MHz}, the coupler limits the pure dephasing lifetime to TϕE400 μsT_{\phi}^{E}\sim 400\text{ }\mu\text{s}.

D.2 Energy relaxation

Energy relaxation of a qubit into its nearest neighbor coupler transmon is approximately given by the Purcell formula. Using the definitions in the main text and supplement the coupler induces relaxation of qubit a

Γ1,aPurcell\displaystyle\Gamma_{1,a}^{\textrm{Purcell}} ga2(Γ1,+β2Δa+2+Γ1,(1β2)Δa2)\displaystyle\approx g_{a}^{2}\left(\frac{\Gamma_{1,+}\beta^{2}}{\Delta_{a+}^{2}}+\frac{\Gamma_{1,-}(1-\beta^{2})}{\Delta_{a-}^{2}}\right)
Γ1,1ga2(ωaω1)2.\displaystyle\sim\frac{\Gamma_{1,1}g_{a}^{2}}{(\omega_{a}-\omega_{1})^{2}}. (63)

Similarly, for qubit b

Γ1,bPurcell\displaystyle\Gamma_{1,b}^{\textrm{Purcell}} gb2(Γ1,β2Δb2+Γ1,+(1β2)Δb+2)\displaystyle\approx g_{b}^{2}\left(\frac{\Gamma_{1,-}\beta^{2}}{\Delta_{b-}^{2}}+\frac{\Gamma_{1,+}(1-\beta^{2})}{\Delta_{b+}^{2}}\right)
Γ1,2gb2(ωbω2)2.\displaystyle\sim\frac{\Gamma_{1,2}g_{b}^{2}}{(\omega_{b}-\omega_{2})^{2}}. (64)

The energy relaxation rates for coupler transmons 1 and 2 are given by Γ1,1\Gamma_{1,1} and Γ1,2\Gamma_{1,2}, respectively. While these quantities are not true observables of the system since the coupler transmons can be strongly hybridized, we want to emphasize that qubit a is not very sensitive to relaxation channels local to coupler transmon 2, nor is qubit b sensitive to relaxation channels local to coupler transmon 1. In the case of direct measurement of coupler relaxation rates, or if we expect correlated relaxation processes, then Γ1,+\Gamma_{1,+} and Γ1,\Gamma_{1,-} describe the energy relaxation rate of the upper and lower hybridized states, respectively.

Plugging in 20 μs20\text{ }\mu\text{s} as a reasonable lower bound on each coupler transmon’s energy relaxation lifetime, the induced T1T_{1} on a neighboring data qubit is 220 μs220\text{ }\mu\text{s} for a g/Δg/\Delta ratio of 0.30.3.

D.3 Coupling a qubit to an open quantum system

We consider coupling qubit ‘aa’ to a bath as mediated by the tunable coupler. In this scenario a deliberate interaction with the bath induces coupler transmon 2 to relax with rate Γ1,2\Gamma_{1,2} into the bath.

Γ1,a\displaystyle\Gamma_{1,a}\approx Γ1,2ga2β2(1β2)(1Δa+2+1Δa2)\displaystyle\Gamma_{1,2}g_{a}^{2}\beta^{2}(1-\beta^{2})\left(\frac{1}{\Delta_{a+}^{2}}+\frac{1}{\Delta_{a-}^{2}}\right)
\displaystyle\approx Γ1,2(gCgL)2ga2(Δa2δ2(gCgL)2)2\displaystyle\frac{\Gamma_{1,2}(g_{C}-g_{L})^{2}g_{a}^{2}}{(\Delta_{a}^{2}-\delta^{2}-(g_{C}-g_{L})^{2})^{2}} (65)

We see that the coupler isolates the qubit from dissipation on the next-to-nearest-neighbor coupler transmon to fourth order in gCgL,ga/Δag_{C}-g_{L},\,g_{a}/\Delta_{a}. Although Γ1,a\Gamma_{1,a} turns off at gCgL=0g_{C}-g_{L}=0, it is difficult to achieve sizeable ‘on’ Γa\Gamma_{a} values using the coupler in dispersive operation. This weak ‘on’ interaction motivates the alternative approach taken in the main text.

D.4 Energy relaxation into the flux bias line

A critical consideration for choosing an appropriate mutual inductance between the coupler SQuID and its flux bias line is the relaxation rate of the coupler induced due to this inductive coupling. This relaxation can lead to strong correlations if δ|gCgL|\delta\ll|g_{C}-g_{L}|. In this circumstance the ‘bright’ state is the one that tunes strongly with flux and will relax, at worst, at the sum of the individual transmon relaxation rates. In the other regime δ|gCgL|\delta\gg|g_{C}-g_{L}|, the eigenstates |±\mathinner{|\pm\rangle} are closely approximated by independent transmon eigenstates, such that we can approximately map ξjξ±\xi_{j}\leftrightarrow\xi_{\pm} for j{1, 2}j\in\{1,\,2\}. Assuming Φ˙e=MI˙\dot{\Phi}_{e}=M\dot{I}, this allows us to write the effective decay rate as,

Γ1,±fb\displaystyle\Gamma_{1,\pm}^{fb} (ξ±(Φe)MΦ0)2SI˙I˙(ω)\displaystyle\sim\left(\frac{\xi_{\pm}(\Phi_{e})M}{\Phi_{0}}\right)^{2}S_{\dot{I}\dot{I}}(\omega)
=2ξ±2(Φe)M2ω±3Φ02Z0,\displaystyle=\frac{2\hbar\xi_{\pm}^{2}(\Phi_{e})M^{2}\omega_{\pm}^{3}}{\Phi_{0}^{2}Z_{0}}, (66)

where in the first line we have used the flux participation ratio as prescribed by the third line of Eq. (13), and in the second line we have assumed that the magnitude of current fluctuations is set by their vacuum expectation value, i.e. SI˙I˙(ω)=ω±2SII(ω)=2ω±3/Z0S_{\dot{I}\dot{I}}(\omega)=\omega_{\pm}^{2}S_{II}(\omega)=2\hbar\omega_{\pm}^{3}/Z_{0}.

Given a mutual inductance M=3.7 pHM=3.7\textrm{ pH}, capacitive coupling C12C1,C2C_{12}\ll C_{1},\,C_{2}, dimensionless coupling constant

ξ±(0)\displaystyle\xi_{\pm}(0) (Φ0/2Z1)E12(0)/E1(0)\displaystyle\sim(\Phi_{0}/\sqrt{2\hbar Z_{1}})E_{12}^{\prime}(0)/E_{1}^{\prime}(0)
or (Φ0/2Z2)E12(0)/E2(0),\displaystyle\textrm{or }(\Phi_{0}/\sqrt{2\hbar Z_{2}})E_{12}^{\prime}(0)/E_{2}^{\prime}(0),

transition frequency ω±/2π=5 GHz\omega_{\pm}/2\pi=5\textrm{ GHz}, and bath impedance of 50 Ohms50\textrm{ Ohms}, the equation above leads to an estimated Γ1,±fb1×103 s1\Gamma_{1,\pm}^{fb}\sim 1\times 10^{3}\textrm{ s}^{-1}, before additional low pass filtering of the flux bias. We note that these are worst case calculations since ξ±(Φe)cos(2πΦe/Φ0)\xi_{\pm}(\Phi_{e})\propto\cos{(2\pi\Phi_{e}/\Phi_{0})}, which at Φe(0)0.25Φ0\Phi_{e}^{(0)}\sim 0.25\Phi_{0} causes the dimensionless coupling constant ξ(Φe)\xi(\Phi_{e}) to vanish.

References

  • Cross [2018] A. Cross, in APS March Meeting Abstracts, Vol. 2018 (2018) pp. L58–003.
  • Arute et al. [2019] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, R. Biswas, S. Boixo, F. G. Brandao, D. A. Buell, et al., Nature 574, 505 (2019).
  • Gong et al. [2021] M. Gong, S. Wang, C. Zha, M.-C. Chen, H.-L. Huang, Y. Wu, Q. Zhu, Y. Zhao, S. Li, S. Guo, et al., Science 372, 948 (2021).
  • AI [2021] G. Q. AI, Nature 595, 383 (2021).
  • Wu et al. [2021] Y. Wu, W.-S. Bao, S. Cao, F. Chen, M.-C. Chen, X. Chen, T.-H. Chung, H. Deng, Y. Du, D. Fan, et al., Physical Review Letters 127, 180501 (2021).
  • Niskanen et al. [2006] A. O. Niskanen, Y. Nakamura, and J.-S. Tsai, Phys. Rev. B 73, 094506 (2006).
  • Niskanen et al. [2007] A. O. Niskanen, K. Harrabi, F. Yoshihara, Y. Nakamura, S. Lloyd, and J. S. Tsai, Science 316, 723 (2007)https://www.science.org/doi/pdf/10.1126/science.1141324 .
  • Allman et al. [2014] M. S. Allman, J. D. Whittaker, M. Castellanos-Beltran, K. Cicak, F. da Silva, M. P. DeFeo, F. Lecocq, A. Sirois, J. D. Teufel, J. Aumentado, and R. W. Simmonds, Phys. Rev. Lett. 112, 123601 (2014).
  • Whittaker et al. [2014] J. D. Whittaker, F. C. S. da Silva, M. S. Allman, F. Lecocq, K. Cicak, A. J. Sirois, J. D. Teufel, J. Aumentado, and R. W. Simmonds, Phys. Rev. B 90, 024513 (2014).
  • Bialczak et al. [2011] R. C. Bialczak, M. Ansmann, M. Hofheinz, M. Lenander, E. Lucero, M. Neeley, A. D. O’Connell, D. Sank, H. Wang, M. Weides, J. Wenner, T. Yamamoto, A. N. Cleland, and J. M. Martinis, Phys. Rev. Lett. 106, 060501 (2011).
  • Chen et al. [2014] Y. Chen, C. Neill, P. Roushan, N. Leung, M. Fang, R. Barends, J. Kelly, B. Campbell, Z. Chen, B. Chiaro, A. Dunsworth, E. Jeffrey, A. Megrant, J. Y. Mutus, P. J. J. O’Malley, C. M. Quintana, D. Sank, A. Vainsencher, J. Wenner, T. C. White, M. R. Geller, A. N. Cleland, and J. M. Martinis, Phys. Rev. Lett. 113, 220502 (2014), publisher: American Physical Society.
  • Geller et al. [2015] M. R. Geller, E. Donate, Y. Chen, M. T. Fang, N. Leung, C. Neill, P. Roushan, and J. M. Martinis, Phys. Rev. A 92, 012320 (2015).
  • Roushan et al. [2016] P. Roushan, C. Neill, A. Megrant, Y. Chen, R. Babbush, R. Barends, B. Campbell, Z. Chen, B. Chiaro, A. Dunsworth, and et al., Nature Physics 13, 146–151 (2016).
  • Neill et al. [2018] C. Neill, P. Roushan, K. Kechedzhi, S. Boixo, S. V. Isakov, V. Smelyanskiy, A. Megrant, B. Chiaro, A. Dunsworth, K. Arya, R. Barends, B. Burkett, Y. Chen, Z. Chen, A. Fowler, B. Foxen, M. Giustina, R. Graff, E. Jeffrey, T. Huang, J. Kelly, P. Klimov, E. Lucero, J. Mutus, M. Neeley, C. Quintana, D. Sank, A. Vainsencher, J. Wenner, T. C. White, H. Neven, and J. M. Martinis, Science 360, 195 (2018)https://www.science.org/doi/pdf/10.1126/science.aao4309 .
  • Noh et al. [2021] T. Noh, Z. Xiao, K. Cicak, X. Y. Jin, E. Doucet, J. Teufel, J. Aumentado, L. C. G. Govia, L. Ranzani, A. Kamal, and R. W. Simmonds, Strong parametric dispersive shifts in a statically decoupled multi-qubit cavity qed system (2021), arXiv:2103.09277 [quant-ph] .
  • Yan et al. [2018] F. Yan, P. Krantz, Y. Sung, M. Kjaergaard, D. L. Campbell, T. P. Orlando, S. Gustavsson, and W. D. Oliver, Phys. Rev. Applied 10, 054062 (2018), publisher: American Physical Society.
  • Sung et al. [2021] 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, Phys. Rev. X 11, 021058 (2021).
  • Li et al. [2020] X. Li, T. Cai, H. Yan, Z. Wang, X. Pan, Y. Ma, W. Cai, J. Han, Z. Hua, X. Han, Y. Wu, H. Zhang, H. Wang, Y. Song, L. Duan, and L. Sun, Phys. Rev. Applied 14, 024070 (2020), publisher: American Physical Society.
  • 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, Phys. Rev. Lett. 125, 240502 (2020).
  • Xu et al. [2020] Y. Xu, J. Chu, J. Yuan, J. Qiu, Y. Zhou, L. Zhang, X. Tan, Y. Yu, S. Liu, J. Li, F. Yan, and D. Yu, Phys. Rev. Lett. 125, 240503 (2020).
  • Stehlik et al. [2021] J. Stehlik, D. M. Zajac, D. L. Underwood, T. Phung, J. Blair, S. Carnevale, D. Klaus, G. A. Keefe, A. Carniol, M. Kumph, and et al., Physical Review Letters 12710.1103/physrevlett.127.080505 (2021).
  • Sete et al. [2021a] E. A. Sete, A. Q. Chen, R. Manenti, S. Kulshreshtha, and S. Poletto, Physical Review Applied 15, 064063 (2021a).
  • Marxer et al. [2022] F. Marxer, A. Vepsäläinen, S. W. Jolin, J. Tuorila, A. Landra, C. Ockeloen-Korppi, W. Liu, O. Ahonen, A. Auer, L. Belzane, et al., arXiv preprint arXiv:2208.09460  (2022).
  • Kandala et al. [2021] A. Kandala, K. X. Wei, S. Srinivasan, E. Magesan, S. Carnevale, G. Keefe, D. Klaus, O. Dial, and D. McKay, Physical Review Letters 127, 130501 (2021).
  • Moskalenko et al. [2022] I. N. Moskalenko, I. A. Simakov, N. N. Abramov, D. O. Moskalev, A. A. Pishchimova, N. S. Smirnov, E. V. Zikiy, I. A. Rodionov, and I. S. Besedin, arXiv preprint arXiv:2203.16302  (2022).
  • Zhao et al. [2022] P. Zhao, Y. Zhang, G. Xue, Y. Jin, and H. Yu, Applied Physics Letters 121, 032601 (2022)https://doi.org/10.1063/5.0097521 .
  • Zakka-Bajjani et al. [2011] E. Zakka-Bajjani, F. Nguyen, M. Lee, L. R. Vale, R. W. Simmonds, and J. Aumentado, Nature Physics 7, 599–603 (2011).
  • Besse et al. [2020] J.-C. Besse, K. Reuer, M. C. Collodo, A. Wulff, L. Wernli, A. Copetudo, D. Malz, P. Magnard, A. Akin, M. Gabureac, G. J. Norris, J. I. Cirac, A. Wallraff, and C. Eichler, Nature Communications 11, 4877 (2020).
  • McKay et al. [2016] D. C. McKay, S. Filipp, A. Mezzacapo, E. Magesan, J. M. Chow, and J. M. Gambetta, Physical Review Applied 610.1103/physrevapplied.6.064007 (2016).
  • Naik et al. [2017] R. K. Naik, N. Leung, S. Chakram, P. Groszkowski, Y. Lu, N. Earnest, D. C. McKay, J. Koch, and D. I. Schuster, Nature Communications 810.1038/s41467-017-02046-6 (2017).
  • Strand et al. [2013] J. D. Strand, M. Ware, F. Beaudoin, T. A. Ohki, B. R. Johnson, A. Blais, and B. L. T. Plourde, Phys. Rev. B 87, 220505 (2013).
  • Sete et al. [2021b] E. A. Sete, N. Didier, A. Q. Chen, S. Kulshreshtha, R. Manenti, and S. Poletto, Phys. Rev. Applied 16, 024050 (2021b).
  • 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 410.1126/sciadv.aao3603 (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, and et al., Physical Review Applied 1010.1103/physrevapplied.10.034050 (2018).
  • Majer et al. [2007] J. Majer, J. M. Chow, J. M. Gambetta, J. Koch, B. R. Johnson, J. A. Schreier, L. Frunzio, D. I. Schuster, A. A. Houck, A. Wallraff, and et al., Nature 449, 443–447 (2007).
  • Chow et al. [2013] J. M. Chow, J. M. Gambetta, A. W. Cross, S. T. Merkel, C. Rigetti, and M. Steffen, New Journal of Physics 15, 115012 (2013).
  • Goto [2022] H. Goto, Phys. Rev. Appl. 18, 034038 (2022).
  • Kubo and Goto [2022] K. Kubo and H. Goto, Fast parametric two-qubit gate for highly detuned fixed-frequency superconducting qubits using a double-transmon coupler (2022).
  • 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).
  • 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).
  • Mundada et al. [2019] P. Mundada, G. Zhang, T. Hazard, and A. Houck, Phys. Rev. Applied 12, 054023 (2019), publisher: American Physical Society.
  • 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, arXiv:2011.07050 [quant-ph]  (2020), arXiv: 2011.07050.
  • 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).
  • Gambetta et al. [2011] J. M. Gambetta, A. A. Houck, and A. Blais, Phys. Rev. Lett. 106, 030502 (2011).
  • 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, and et al., Nature Communications 710.1038/ncomms12964 (2016).
  • Kannan et al. [2022] B. Kannan, A. Almanakly, Y. Sung, A. Di Paolo, D. A. Rower, J. Braumüller, A. Melville, B. M. Niedzielski, A. Karamlou, K. Serniak, A. Vepsäläinen, M. E. Schwartz, J. L. Yoder, R. Winik, J. I.-J. Wang, T. P. Orlando, S. Gustavsson, J. A. Grover, and W. D. Oliver, On-demand directional photon emission using waveguide quantum electrodynamics (2022).
  • Xiao et al. [2021] Z. Xiao, E. Doucet, T. Noh, L. Ranzani, R. W. Simmonds, L. C. G. Govia, and A. Kamal, Perturbative diagonalization for time-dependent strong interactions (2021), arXiv:2103.09260 [quant-ph] .
  • Braumüller et al. [2020] J. Braumüller, L. Ding, A. P. Vepsäläinen, Y. Sung, M. Kjaergaard, T. Menke, R. Winik, D. Kim, B. M. Niedzielski, A. Melville, and et al., Physical Review Applied 1310.1103/physrevapplied.13.054079 (2020).
  • Kiktenko et al. [2020] E. Kiktenko, A. Nikolaeva, P. Xu, G. Shlyapnikov, and A. Fedorov, Physical Review A 101, 022304 (2020).
  • Baker et al. [2020] J. M. Baker, C. Duckering, and F. T. Chong, in 2020 IEEE 50th International Symposium on Multiple-Valued Logic (ISMVL) (IEEE, 2020) pp. 303–308.
  • Riwar and DiVincenzo [2021] R.-P. Riwar and D. P. DiVincenzo, Circuit quantization with time-dependent magnetic fields for realistic geometries (2021), arXiv:2103.03577 [cond-mat.mes-hall] .
  • Vool and Devoret [2017] U. Vool and M. Devoret, International Journal of Circuit Theory and Applications 45, 897–934 (2017).
  • You et al. [2019] X. You, J. A. Sauls, and J. Koch, Phys. Rev. B 99, 174512 (2019).
  • Johansson et al. [2012] J. Johansson, P. Nation, and F. Nori, Computer Physics Communications 183, 1760 (2012).
  • Johansson et al. [2013] J. Johansson, P. Nation, and F. Nori, Computer Physics Communications 184, 1234 (2013).
  • Altland and Simons [2010] A. Altland and B. D. Simons, Condensed Matter Field Theory, 2nd ed. (Cambridge University Press, 2010).