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

Hamiltonian of a flux qubit-LC oscillator circuit in the deep-strong-coupling regime

F. Yoshihara [email protected] Advanced ICT Research Institute, National Institute of Information and Communications Technology, 4-2-1, Nukuikitamachi, Koganei, Tokyo 184-8795, Japan    S. Ashhab Advanced ICT Research Institute, National Institute of Information and Communications Technology, 4-2-1, Nukuikitamachi, Koganei, Tokyo 184-8795, Japan Qatar Environment and Energy Research Institute, Hamad Bin Khalifa University, Qatar Foundation, Doha, Qatar    T. Fuse Advanced ICT Research Institute, National Institute of Information and Communications Technology, 4-2-1, Nukuikitamachi, Koganei, Tokyo 184-8795, Japan    M. Bamba Department of Physics, Kyoto University, Kyoto 606-8502, Japan PRESTO, Japan Science and Technology Agency, Kawaguchi 332-0012, Japan    K. Semba Present address: Institute for Photon Science and Technology, The University of Tokyo, Tokyo 113-0033, Japan. Advanced ICT Research Institute, National Institute of Information and Communications Technology, 4-2-1, Nukuikitamachi, Koganei, Tokyo 184-8795, Japan
Abstract

We derive the Hamiltonian of a superconducting circuit that comprises a single-Josephson-junction flux qubit inductively coupled to an LC oscillator, and we compare the derived circuit Hamiltonian with the quantum Rabi Hamiltonian, which describes a two-level system coupled to a harmonic oscillator. We show that there is a simple, intuitive correspondence between the circuit Hamiltonian and the quantum Rabi Hamiltonian. While there is an overall shift of the entire spectrum, the energy level structure of the circuit Hamiltonian up to the seventh excited states can still be fitted well by the quantum Rabi Hamiltonian even in the case where the coupling strength is larger than the frequencies of the qubit and the oscillator, i.e., when the qubit-oscillator circuit is in the deep-strong-coupling regime. We also show that although the circuit Hamiltonian can be transformed via a unitary transformation to a Hamiltonian containing a capacitive coupling term, the resulting circuit Hamiltonian cannot be approximated by the variant of the quantum Rabi Hamiltonian that is obtained using an analogous procedure for mapping the circuit variables onto Pauli and harmonic oscillator operators, even for relatively weak coupling. This difference between the flux and charge gauges follows from the properties of the qubit Hamiltonian eigenstates.

introduction

Superconducting circuits are one of the most promising platforms for realizing large-scale quantum information processing. One of the most important features of superconducting circuits is the freedom they allow in their circuit design. Since the first demonstration of coherent control of a Cooper pair box Nakamura et al. (1999), various types of superconducting circuits have been demonstrated.

The Hamiltonian of a superconducting circuit can be derived using the standard quantization procedure applied to the charge and flux variables in the circuit Vool and Devoret (2017). The Hamiltonian of an LC circuit is well known to be that of a harmonic oscillator. The Hamiltonians of various kinds of superconducting qubits have also been well studied Nakamura et al. (1997); Orlando et al. (1999); Vion et al. (2002); Martinis et al. (2002); Koch et al. (2007) and these Hamiltonians can be numerically diagonalized to obtain eigenenergies and eigenstates that accurately reproduce experimental data. On the other hand, the Hamiltonian of circuits containing two or more components, e.g., qubit-qubit, qubit-oscillator, or oscillator-oscillator systems, are usually treated in such a way that the Hamiltonian of the individual components and the coupling among them are separately obtained Pashkin et al. (2003); Chiorescu et al. (2004); Blais et al. (2004); Niskanen et al. (2006). This separate treatment of individual circuit components works reasonably well in most circuits. Even for flux qubit-oscillator circuits in the ultrastrong-coupling regime Niemczyk et al. (2010); Forn-Diaz et al. (2010), where the coupling strength gg is around 10% of the oscillator’s frequency ω\omega and the qubit minimum frequency Δq\Delta_{q}, or the deep-strong-coupling regime Yoshihara et al. (2017a, b, 2018), where gg is comparable to or larger than Δq\Delta_{q} and ω\omega, the experimental data can be well fitted by the quantum Rabi Hamiltonian Rabi (1937); Jaynes and Cummings (1963); Braak (2011), where a two-level atom and a harmonic oscillator are coupled by a dipole-dipole interaction. However, the more rigorous approach based on the standard quantization procedure has not been applied to such circuits except in a few specific studies Bourassa et al. (2009); Peropadre et al. (2013); Smith et al. (2016), and the validity of describing a flux qubit-oscillator circuit by the quantum Rabi Hamiltonian has been demonstrated in only a few specific circuits Manucharyan et al. (2017); De Bernardis et al. (2018).

In this paper, we apply the standard quantization procedure to a superconducting circuit in which a single-Josephson-junction flux qubit (an rf-SQUID qubit or a fluxonium-equivalent circuit) and an LC oscillator are inductively coupled to each other by a shared inductor [Fig. 1a], and derive the Hamiltonian of the circuit. Our model with a single Josephson junction should be sufficient for the purposes of the present study, and the results can be applied to circuits of commonly used multi-Josephson-junction flux qubits Chiorescu et al. (2003); Robertson et al. (2006), including the fluxonium. Note that single-Josephson-junction flux qubits with different parameters have been experimentally demonstrated using superinductors, which have been realized by high-kinetic-inductance superconductors Peltonen et al. (2018); Hazard et al. (2019), granular aluminum Grünhaupt et al. (2019), and Josephson-junction arrays Manucharyan et al. (2009). The derived circuit Hamiltonian consists of terms associated with the LC oscillator, the flux qubit (and its higher energy levels), and the product of the two flux operators. Excluding the qubit’s energy levels higher than the first excited state, this circuit Hamiltonian takes the form of the quantum Rabi Hamiltonian, which describes a two-level system coupled to a harmonic oscillator. To investigate contributions from the qubit’s higher energy levels, we numerically calculate the transition frequencies of the circuit Hamiltonian. We find that the qubit’s higher energy levels mainly cause a negative shift of the entire spectrum, and that the calculated transition frequencies are well fitted by the quantum Rabi Hamiltonian even when the qubit-oscillator circuit is in the deep-strong-coupling regime. The circuit Hamiltonian can be transformed to one that has a capacitive coupling term by a unitary transformation. We show, however, that the spectrum of the circuit Hamiltonian cannot be fitted by the variant of the quantum Rabi Hamiltonian that has different Pauli operators in the qubit and coupling terms. This situation arises when we perform the mapping from circuit variables to Pauli operators for a circuit Hamiltonian that has a capacitive coupling term, i.e. a Hamiltonian expressed in the charge gauge. We explain the advantage of the flux gauge over the charge gauge in this regard based on the properties of the eigenstates of the flux qubit Hamiltonian.

Refer to caption
Figure 1: Circuit diagrams. a A superconducting circuit in which a single-Josephson-junction flux qubit and an LC oscillator are inductively coupled to each other by a shared inductor. b Equivalent circuit obtained by applying the so-called Y-Δ\Delta transformation to the inductor network in circuit a. c The outer loop of circuit a, which forms an LC oscillator. d The inner loop of circuit a, which forms a single-Josephson-junction flux qubit.

Results

circuit Hamiltonian

Following the standard quantization procedure, nodes are assigned to the circuit as shown in Fig. 1a. Before deriving the circuit Hamiltonian, the circuit in Fig. 1a is transformed to the one shown in Fig. 1b by applying the so-called Y-Δ\Delta transformation, by which a Δ\Delta-shaped network of electrical elements is converted to an equivalent Y-shaped network or vice versa, to the inductor network. Thus, node 3 surrounded by the inductors is eliminated. The inductances of the new set of inductors are given as

Lg1=(LcL1+LcL2+L1L2)/L2,\displaystyle L_{g1}=(L_{c}L_{1}+L_{c}L_{2}+L_{1}L_{2})/L_{2}, (1)
Lg2=(LcL1+LcL2+L1L2)/L1,\displaystyle L_{g2}=(L_{c}L_{1}+L_{c}L_{2}+L_{1}L_{2})/L_{1}, (2)

and

L12=(LcL1+LcL2+L1L2)/Lc.\displaystyle L_{12}=(L_{c}L_{1}+L_{c}L_{2}+L_{1}L_{2})/L_{c}. (3)

The Lagrangian of the circuit can now be obtained relatively easily Vool and Devoret (2017):

^circ=C2Φ^˙12+CJ2Φ^˙22+EJcos(2πΦ^2ΦxΦ0)Φ^122Lg1Φ^222Lg2(Φ^2Φ^1)22L12,\displaystyle\hat{\mathcal{L}}_{circ}=\frac{C}{2}\dot{\hat{\Phi}}_{1}^{2}+\frac{C_{J}}{2}\dot{\hat{\Phi}}_{2}^{2}+E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right)-\frac{\hat{\Phi}_{1}^{2}}{2L_{g1}}-\frac{\hat{\Phi}_{2}^{2}}{2L_{g2}}-\frac{(\hat{\Phi}_{2}-\hat{\Phi}_{1})^{2}}{2L_{12}}, (4)

where CJC_{J} and EJ=IcΦ0/(2π)E_{J}=I_{c}\Phi_{0}/(2\pi) are the capacitance and the Josephson energy of the Josephson junction, IcI_{c} is the critical current of the Josephson junction, Φ0=h/(2e)\Phi_{0}=h/(2e) is the superconducting flux quantum, and Φ^k\hat{\Phi}_{k} and Φ^˙k\dot{\hat{\Phi}}_{k} (k=1,2)(k=1,2) are the node flux and its time derivative for node kk. The node charges are defined as the conjugate momenta of the node fluxes as q^k=^circ/Φ^˙k\hat{q}_{k}=\partial\hat{\mathcal{L}}_{circ}/\partial\dot{\hat{\Phi}}_{k}. After the Legendre transformation, the Hamiltonian is obtained as

^circ\displaystyle\hat{\mathcal{H}}_{circ} =\displaystyle= q^122C+q^222CJEJcos(2πΦ^2ΦxΦ0)+Φ^122LLC+Φ^222LFQΦ^1Φ^2L12\displaystyle\frac{\hat{q}_{1}^{2}}{2C}+\frac{\hat{q}_{2}^{2}}{2C_{J}}-E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right)+\frac{\hat{\Phi}_{1}^{2}}{2L_{LC}}+\frac{\hat{\Phi}_{2}^{2}}{2L_{FQ}}-\frac{\hat{\Phi}_{1}\hat{\Phi}_{2}}{L_{12}}
=\displaystyle= ^1+^2+^12,\displaystyle\hat{\mathcal{H}}_{1}+\hat{\mathcal{H}}_{2}+\hat{\mathcal{H}}_{12},

where

^1\displaystyle\hat{\mathcal{H}}_{1} =\displaystyle= q^122C+Φ^122LLC,\displaystyle\frac{\hat{q}_{1}^{2}}{2C}+\frac{\hat{\Phi}_{1}^{2}}{2L_{LC}}, (6)
^2\displaystyle\hat{\mathcal{H}}_{2} =\displaystyle= q^222CJEJcos(2πΦ^2ΦxΦ0)+Φ^222LFQ,\displaystyle\frac{\hat{q}_{2}^{2}}{2C_{J}}-E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right)+\frac{\hat{\Phi}_{2}^{2}}{2L_{FQ}}, (7)
^12\displaystyle\hat{\mathcal{H}}_{12} =\displaystyle= Φ^1Φ^2L12,\displaystyle-\frac{\hat{\Phi}_{1}\hat{\Phi}_{2}}{L_{12}}, (8)
1LLC\displaystyle\frac{1}{L_{LC}} =\displaystyle= 1Lg1+1L12,\displaystyle\frac{1}{L_{g1}}+\frac{1}{L_{12}}, (9)

and

1LFQ\displaystyle\frac{1}{L_{FQ}} =\displaystyle= 1Lg2+1L12.\displaystyle\frac{1}{L_{g2}}+\frac{1}{L_{12}}. (10)

As can be seen from Eq. (circuit Hamiltonian), the Hamiltonian ^circ\hat{\mathcal{H}}_{circ} can be separated into three parts: the first part ^1\hat{\mathcal{H}}_{1} consisting of the charge and flux operators of node 1, the second part ^2\hat{\mathcal{H}}_{2} consisting of node 2 operators, and the third part ^12\hat{\mathcal{H}}_{12} containing the product of the two flux operators.

separate treatment of the qubit-oscillator circuit

Let us consider an alternative treatment of the circuit shown in Fig. 1a, where the circuit is assumed to be naively divided into two well-defined components. The capacitor and the inductors in the outer loop of the circuit in Fig. 1a form an LC oscillator [Fig. 1c]. The Josephson junction and the inductors in the inner loop form a single-Josephson-junction flux qubit [Fig. 1d]. The LC oscillator and the single-Josephson-junction flux qubit share the inductor LcL_{c} at the common part of the two loops. It is now instructive to investigate the relation of the following pairs of Hamiltonians: ^1\hat{\mathcal{H}}_{1} and the Hamiltonian of the LC oscillator shown in Fig. 1c, ^2\hat{\mathcal{H}}_{2} and the Hamiltonian of the flux qubit shown in Fig. 1d, and ^12\hat{\mathcal{H}}_{12} and the Hamiltonian of the inductive coupling between the LC oscillator and the flux qubit MI^LCI^q=LcΦ^1Φ^2/[(Lc+L1)(Lc+L2)]M\hat{I}_{LC}\hat{I}_{q}=-L_{c}\hat{\Phi}_{1}\hat{\Phi}_{2}/[(L_{c}+L_{1})(L_{c}+L_{2})], where we have used the relations of the oscillator current I^LC=Φ^1/(Lc+L1)\hat{I}_{LC}=\hat{\Phi}_{1}/(L_{c}+L_{1}), the qubit current I^q=Φ^2/(Lc+L2)\hat{I}_{q}=\hat{\Phi}_{2}/(L_{c}+L_{2}), and the mutual inductance M=LcM=L_{c}. Actually, only the inductances are different in the Hamiltonians of the two pictures: The inductances of the LC oscillator, the flux qubit, and the coupling Hamiltonians derived from the separate treatment are respectively Lc+L1L_{c}+L_{1}, Lc+L2L_{c}+L_{2}, and (Lc+L1)(Lc+L2)/Lc-(L_{c}+L_{1})(L_{c}+L_{2})/L_{c}, while those in ^circ\hat{\mathcal{H}}_{circ} are LLCL_{LC}, LFQL_{FQ}, and L12-L_{12}. Figure 2 shows the inductances in the separate treatment and in ^circ\hat{\mathcal{H}}_{circ} as functions of LcL_{c} on condition that the inductance sums are kept constant at Lc+L1=800L_{c}+L_{1}=800 pH and Lc+L2=2050L_{c}+L_{2}=2050 pH. As LcL_{c} approaches 0, or more specifically when LcL1,L2L_{c}\ll L_{1},L_{2}, we obtain Lc+L1L_{c}+L_{1} \sim LLCL_{LC}, Lc+L2L_{c}+L_{2} \sim LFQL_{FQ}, and (Lc+L1)(Lc+L2)/Lc(L_{c}+L_{1})(L_{c}+L_{2})/L_{c} \sim L12L_{12}. In this way, the Hamiltonian derived from the separate treatment has the same form as ^circ\hat{\mathcal{H}}_{circ}, and the inductances in the separate treatment approach those of ^circ\hat{\mathcal{H}}_{circ}.

Refer to caption
Figure 2: a The inductances of the LC oscillator LLCL_{LC}, b the inductance of the flux qubit LFQL_{FQ}, and c the inverse inductance of ^12\hat{\mathcal{H}}_{12}, 1/L121/L_{12} obtained by Eqs. (9), (10), and (3) (solid lines) are plotted as functions of LcL_{c} together with their counterparts in the separate treatment (dotted lines) on condition that the inductance sums are kept constant at Lc+L1=800L_{c}+L_{1}=800 pH and Lc+L2=2050L_{c}+L_{2}=2050 pH.

comparison to the quantum Rabi Hamiltonian

We compare the Hamiltonian ^circ\hat{\mathcal{H}}_{circ} with the generalized quantum Rabi Hamiltonian:

^R/\displaystyle\hat{\mathcal{H}}_{R}/\hbar =\displaystyle= ω(a^a^+12)12(εσ^x+Δqσ^z)+gσ^x(a^+a^)\displaystyle\omega\left(\hat{a}^{\dagger}\hat{a}+\frac{1}{2}\right)-\frac{1}{2}(\varepsilon\hat{\sigma}_{x}+\Delta_{q}\hat{\sigma}_{z})+g\hat{\sigma}_{x}(\hat{a}+\hat{a}^{\dagger})
=\displaystyle= (^LC+^FQ+^coup)/.\displaystyle\left(\hat{\mathcal{H}}_{LC}+\hat{\mathcal{H}}_{FQ}+\hat{\mathcal{H}}_{coup}\right)/\hbar.

The first part ^LC\hat{\mathcal{H}}_{LC} represents the energy of the LC oscillator, where a^\hat{a}^{\dagger} and a^\hat{a} are the creation and annihilation operators, respectively. The second part ^FQ\hat{\mathcal{H}}_{FQ} represents the energy of the flux qubit written in the energy eigenbasis at ε=0\varepsilon=0. The operators σ^x,z\hat{\sigma}_{x,z} are the standard Pauli operators. The parameters Δq\hbar\Delta_{q} and ε\hbar\varepsilon are the tunnel splitting and the energy bias between the two states with persistent currents flowing in opposite directions around the qubit loop. The third part ^coup\hat{\mathcal{H}}_{coup} represents the inductive coupling energy.

The relation between ^1\hat{\mathcal{H}}_{1} and ^LC\hat{\mathcal{H}}_{LC} is straightforward. The resonance frequency and the operators in ^LC\hat{\mathcal{H}}_{LC} can be analytically described by the variables and operators in ^1\hat{\mathcal{H}}_{1} as ω=1/LLCC\omega=1/\sqrt{L_{LC}C} and a^+a^Φ^1/(LLCIzpf)\hat{a}+\hat{a}^{\dagger}\to\hat{\Phi}_{1}/(L_{LC}I_{zpf}), where Izpf=ω/(2LLC)I_{zpf}=\sqrt{\hbar\omega/(2L_{LC})} is the zero-point-fluctuation current.

Refer to caption
Figure 3: a Numerically calculated energy levels of ^2\hat{\mathcal{H}}_{2} as functions of Φx\Phi_{x}. b The lowest two energy levels of ^2\hat{\mathcal{H}}_{2}. c The expectation values of the flux operator g|Φ^2|g\left\langle g\left|\hat{\Phi}_{2}\right|g\right\rangle and e|Φ^2|e\left\langle e\left|\hat{\Phi}_{2}\right|e\right\rangle. In b and c, the solid circles are obtained from numerical calculations of ^2\hat{\mathcal{H}}_{2}, while the lines are obtained from fitting by ^FQ\hat{\mathcal{H}}_{FQ}. The black and red colors respectively indicate states |g\left|g\right\rangle and |e\left|e\right\rangle. In the calculation, we used the following parameters: Lc+L2=2050L_{c}+L_{2}=2050 pH, LJ=990L_{J}=990 pH (EJ/h=165.1E_{J}/h=165.1 GHz), and CJ=4.84C_{J}=4.84 fF (EC/h=4.0E_{C}/h=4.0 GHz).

To see the relation between ^2\hat{\mathcal{H}}_{2} and ^FQ\hat{\mathcal{H}}_{FQ}, we numerically calculated the eigenenergies of ^2\hat{\mathcal{H}}_{2} as functions of Φx\Phi_{x}. In the calculation, we used the following parameters: Lc+L2=2050L_{c}+L_{2}=2050 pH, LJ=990L_{J}=990 pH (EJ/h=165.1E_{J}/h=165.1 GHz), and CJ=4.84C_{J}=4.84 fF (EC/h=4.0E_{C}/h=4.0 GHz). As shown in Fig. 3a the lowest two energy levels are well separated from the higher levels, which are more than 40 GHz higher in frequency. The lowest two energy levels of ^2\hat{\mathcal{H}}_{2} are well approximated by ^FQ\hat{\mathcal{H}}_{FQ} [Fig. 3b], which gives almost identical results obtained by the local basis reduction method Consani and Warburton (2020), with σ^x\hat{\sigma}_{x} and σ^z\hat{\sigma}_{z} swapped. The eigen frequencies of the ground state ω0\omega_{0} and the first excited state ω1\omega_{1} are respectively fitted by ωosε2+Δq2/2\omega_{os}-\sqrt{\varepsilon^{2}+\Delta_{q}^{2}}/2 and ωos+ε2+Δq2/2\omega_{os}+\sqrt{\varepsilon^{2}+\Delta_{q}^{2}}/2. Besides the offset ωos\omega_{os}, the fitting parameters are Δq\Delta_{q} and the maximum persistent current IpI_{p}, which is determined as the proportionality constant between the energy bias and the flux bias, ε=2Ip(Φx0.5Φ0)\varepsilon=2I_{p}(\Phi_{x}-0.5\Phi_{0}). We also numerically calculated the expectation values of the flux operator, g|Φ^2|g\left\langle g\left|\hat{\Phi}_{2}\right|g\right\rangle and e|Φ^2|e\left\langle e\left|\hat{\Phi}_{2}\right|e\right\rangle, which are well approximated by g|σ^x|g\left\langle g\left|\hat{\sigma}_{x}\right|g\right\rangle and e|σ^x|e\left\langle e\left|\hat{\sigma}_{x}\right|e\right\rangle, respectively [Fig. 3c]. The expectation values of the flux operator g|Φ^2|g\left\langle g\left|\hat{\Phi}_{2}\right|g\right\rangle and e|Φ^2|e\left\langle e\left|\hat{\Phi}_{2}\right|e\right\rangle are respectively fitted by Φ2maxε/ε2+Δq2-\Phi_{2max}\varepsilon/\sqrt{\varepsilon^{2}+\Delta_{q}^{2}} and Φ2maxε/ε2+Δq2\Phi_{2max}\varepsilon/\sqrt{\varepsilon^{2}+\Delta_{q}^{2}}. Here, |g\left|g\right\rangle and |e\left|e\right\rangle are respectively the ground and excited states of the qubit Hamiltonian ^2\hat{\mathcal{H}}_{2}. The only fitting parameter is determined by the ratio Φ2max=i|Φ^2|i/i|σ^x|i\Phi_{2max}=-\left\langle i\left|\hat{\Phi}_{2}\right|i\right\rangle/\left\langle i\left|\hat{\sigma}_{x}\right|i\right\rangle (i=g,ei=g,e). The Pauli operator σ^x\hat{\sigma}_{x} is therefore identified as being proportional to the flux operator σ^xΦ^2/Φ2max\hat{\sigma}_{x}\to-\hat{\Phi}_{2}/\Phi_{2max}. The relation between ^12\hat{\mathcal{H}}_{12} and ^coup\hat{\mathcal{H}}_{coup} can now be obtained by using the relations for the oscillator and qubit operators identified above. This way we find that the Hamiltonian ^12\hat{\mathcal{H}}_{12} can be expressed as (LLC/L12)IzpfΦ2maxσ^x(a^+a^)-(L_{LC}/L_{12})I_{zpf}\Phi_{2max}\hat{\sigma}_{x}(\hat{a}+\hat{a}^{\dagger}), which is exactly the same form as ^coup\hat{\mathcal{H}}_{coup}, with the coupling strength g=(LLC/L12)IzpfΦ2max\hbar g=-(L_{LC}/L_{12})I_{zpf}\Phi_{2max}. Note that coup\mathcal{H}_{coup} directly derived from the Lagrangian ^circ\hat{\mathcal{L}}_{circ} is in the flux gauge, as the qubit-oscillator coupling term is of the form Φ^1Φ^2\hat{\Phi}_{1}\hat{\Phi}_{2}, which is optimal for our system with a single oscillator mode Stokes and Nazir (2019); Roth et al. (2019); Di Stefano et al. (2019). Excluding the qubit’s energy levels higher than the first excited states, ^circ\hat{\mathcal{H}}_{circ} takes the form of ^R\hat{\mathcal{H}}_{R}. In other words, once the circuit parameters, i.e. Φx\Phi_{x}, LcL_{c}, L1L_{1}, L2L_{2}, CC, EJE_{J}, and CJC_{J}, are given, the corresponding parameters in ^R\hat{\mathcal{H}}_{R} (ω\omega, ε\varepsilon, Δq\Delta_{q}, and gg) are set. The relation between ^circ\hat{\mathcal{H}}_{circ} and ^R\hat{\mathcal{H}}_{R} is summarized in Table 1.

^R\hat{\mathcal{H}}_{R} ^circ\hat{\mathcal{H}}_{circ}
^LC\hat{\mathcal{H}}_{LC} ^1\hat{\mathcal{H}}_{1}
^FQ\hat{\mathcal{H}}_{FQ} ^2\hat{\mathcal{H}}_{2}
^coup\hat{\mathcal{H}}_{coup} ^12\hat{\mathcal{H}}_{12}
a^+a^\hat{a}+\hat{a}^{\dagger} Φ^1/(LLCIzpf)\hat{\Phi}_{1}/(L_{LC}I_{zpf})
(a^a^)/i(\hat{a}-\hat{a}^{\dagger})/i q^1/(CVzpf)\hat{q}_{1}/(CV_{zpf})
σ^x\hat{\sigma}_{x} Φ^2/Φ2max-\hat{\Phi}_{2}/\Phi_{2max}
σ^z\hat{\sigma}_{z}
ω\omega 1/LLCC1/\sqrt{L_{LC}C}
ε\varepsilon 2Ip(Φx0.5Φ0)2I_{p}(\Phi_{x}-0.5\Phi_{0})
Δq\Delta_{q} minimum qubit frequency
(numerically evaluated)
gg (LLC/L12)IzpfΦ2max/(L_{LC}/L_{12})I_{zpf}\Phi_{2max}/\hbar
Table 1: Hamiltonians, operators, and variables in ^R\hat{\mathcal{H}}_{R} and their counterparts in ^circ\hat{\mathcal{H}}_{circ}. Izpf=ω/(2LLC)I_{zpf}=\sqrt{\hbar\omega/(2L_{LC})} and Vzpf=ω/(2C)V_{zpf}=\sqrt{\hbar\omega/(2C)} are the zero-point-fluctuation current and voltage, respectively. The parameters Φ2max\Phi_{2max}, IpI_{p}, and Δq\Delta_{q} are obtained by numerically calculating the eigenenergies of ^2\hat{\mathcal{H}}_{2} as functions of Φx\Phi_{x} and fitting the lowest two energy levels by ^FQ\hat{\mathcal{H}}_{FQ}. Note that there is no analytic expression for σ^z\hat{\sigma}_{z} and Δq\Delta_{q}.

As previously mentioned, ^R\hat{\mathcal{H}}_{R} considers only the lowest two energy levels of the flux qubit, while ^circ\hat{\mathcal{H}}_{circ} includes all energy levels. To investigate the effect of the qubit’s higher energy levels, we perform numerical calculations and compare the energy levels calculated by ^R\hat{\mathcal{H}}_{R} and ^circ\hat{\mathcal{H}}_{circ}. The details of the numerical diagonalization of ^circ\hat{\mathcal{H}}_{circ} are given in methods. Since the contributions from the qubit’s higher energy levels are expected to become larger as the coupling strength increases, we consider parameters that cover a wide range of coupling strengths from the weak-coupling to the deep-strong-coupling regime. In the calculations, we fix Lc+L1=800L_{c}+L_{1}=800 pH, Lc+L2=2050L_{c}+L_{2}=2050 pH, C=0.87C=0.87 pF, LJ=990L_{J}=990 pH (EJ/h=165.1E_{J}/h=165.1 GHz), and CJ=4.84C_{J}=4.84 fF (EC/h=4.0E_{C}/h=4.0 GHz), and sweep the flux bias Φx\Phi_{x} around Φ0/2\Phi_{0}/2 at various values of LcL_{c}. These parameters are used in all the calculations for Figs. 4-7. Some of our calculations were performed using the QuTiP simulation package Johansson et al. (2013).

Refer to caption
Figure 4: Transition frequencies of the qubit-oscillator circuit fitted with different approximate Hamiltonians. Numerically calculated transition frequencies from ^circ\hat{\mathcal{H}}_{circ} are plotted as circles while the fitting spectra are plotted as lines. Two mutual inductance values, corresponding to different qubit-oscillator coupling strengths, are used: LcL_{c} = 20 pH (a, c, e, g) and LcL_{c} = 350 pH (b, d, f, h). The spectra are fitted with ^R\hat{\mathcal{H}}_{R} (a, b, c, d) and ^R\hat{\mathcal{H}}_{R}^{\prime} (e, f, g, h). a, b The parameters of ^R\hat{\mathcal{H}}_{R} are obtained from the relations described in Table 1: ω/(2π)=6.033\omega/(2\pi)=6.033 GHz, Δq/(2π)=1.240\Delta_{q}/(2\pi)=1.240 GHz, g/(2π)=0.424g/(2\pi)=0.424 GHz, and Ip=281.3I_{p}=281.3 nA for Lc=20L_{c}=20 pH and ω/(2π)=6.272\omega/(2\pi)=6.272 GHz, Δq/(2π)=2.139\Delta_{q}/(2\pi)=2.139 GHz, g/(2π)=7.338g/(2\pi)=7.338 GHz, and Ip=282.5I_{p}=282.5 nA for Lc=350L_{c}=350 pH. c, d The parameters of ^R\hat{\mathcal{H}}_{R} are obtained by fitting the spectra of ^circ\hat{\mathcal{H}}_{circ} to ^R\hat{\mathcal{H}}_{R}: ω/(2π)=6.033\omega/(2\pi)=6.033 GHz, Δq/(2π)=1.240\Delta_{q}/(2\pi)=1.240 GHz, g/(2π)=0.430g/(2\pi)=0.430 GHz, and Ip=281.3I_{p}=281.3 nA for Lc=20L_{c}=20 pH and ω/(2π)=6.064\omega/(2\pi)=6.064 GHz, Δq/(2π)=2.388\Delta_{q}/(2\pi)=2.388 GHz, g/(2π)=7.822g/(2\pi)=7.822 GHz, and Ip=282.9I_{p}=282.9 nA for Lc=350L_{c}=350 pH. e, f The parameters of ^R\hat{\mathcal{H}}_{R}^{\prime} are obtained from the circuit parameters of ^circ\hat{\mathcal{H}}_{circ}^{\prime} in the similar way as in Table 1: ω/(2π)=6.085\omega/(2\pi)=6.085 GHz, Δq/(2π)=1.238\Delta_{q}/(2\pi)=1.238 GHz, g/(2π)=g^{\prime}/(2\pi)=0.043 GHz, and Ip=281.3I_{p}=281.3 nA for Lc=20L_{c}=20 pH, and ω/(2π)=15.66\omega/(2\pi)=15.66 GHz, Δq/(2π)=1.238\Delta_{q}/(2\pi)=1.238 GHz, g/(2π)=g^{\prime}/(2\pi)=0.492 GHz, and Ip=281.3I_{p}=281.3 nA for Lc=350L_{c}=350 pH. g, h The parameters of ^R\hat{\mathcal{H}}_{R}^{\prime} are obtained by fitting the spectra of ^circ\hat{\mathcal{H}}_{circ}^{\prime} to ^R\hat{\mathcal{H}}_{R}^{\prime}: ω/(2π)=6.035\omega/(2\pi)=6.035 GHz, Δq/(2π)=1.220\Delta_{q}/(2\pi)=1.220 GHz, g/(2π)=0.089g/(2\pi)=0.089 GHz, and Ip=281.3I_{p}=281.3 nA for Lc=20L_{c}=20 pH and ω/(2π)=6.034\omega/(2\pi)=6.034 GHz, Δq/(2π)=0.233\Delta_{q}/(2\pi)=0.233 GHz, g/(2π)=0.165g/(2\pi)=0.165 GHz, and Ip=281.4I_{p}=281.4 nA for Lc=350L_{c}=350 pH. Black, gray, orange, magenta, and red colors indicate transitions |0|1\left|0\right\rangle\to\left|1\right\rangle, |0|2\left|0\right\rangle\to\left|2\right\rangle, |0|3\left|0\right\rangle\to\left|3\right\rangle, |1|2\left|1\right\rangle\to\left|2\right\rangle, and |1|3\left|1\right\rangle\to\left|3\right\rangle, respectively.

Transition frequencies of the qubit-oscillator circuit ωij\omega_{ij} corresponding to the transition |i|j\left|i\right\rangle\to\left|j\right\rangle numerically calculated from ^circ\hat{\mathcal{H}}_{circ} around the resonance frequency of the oscillator ω\omega are plotted in Fig. 4 for a Lc=20L_{c}=20 pH and b Lc=350L_{c}=350 pH, where the indices ii and jj label the energy eigenstates according to their order in the energy-level ladder, with the index 0 denoting the ground state. The same circuit parameters as in Fig. 3 are used. In the case Lc=20L_{c}=20 pH, two characteristic features are observed: avoided level crossings between the qubit and oscillator transition signals approximately at Φx/Φ0=0.497\Phi_{x}/\Phi_{0}=0.497 and 0.503, and the dispersive shift that for example creates the separation between the frequencies of the transitions |0|2\left|0\right\rangle\to\left|2\right\rangle and |1|3\left|1\right\rangle\to\left|3\right\rangle, leading to the peak/dip feature at the symmetry point, i.e. Φx/Φ0=0.5\Phi_{x}/\Phi_{0}=0.5. Note that transitions |0|2\left|0\right\rangle\to\left|2\right\rangle and |1|3\left|1\right\rangle\to\left|3\right\rangle around Φx/Φ0=0.5\Phi_{x}/\Phi_{0}=0.5 respectively correspond to transitions |g0|g1\left|g0\right\rangle\to\left|g1\right\rangle and |e0|e1\left|e0\right\rangle\to\left|e1\right\rangle, where “gg” and “ee” denote, respectively, the ground and excited states of the qubit, and “0” and “1” the number of photons in the oscillator’s Fock state. In the case Lc=350L_{c}=350 pH, the characteristic spectrum indicates that the qubit-oscillator circuit is in the deep-strong-coupling regime Yoshihara et al. (2017b). Transition frequencies of ^R\hat{\mathcal{H}}_{R} are also plotted in Figs. 4a and b. It should be mentioned that the parameters of ^R\hat{\mathcal{H}}_{R} are obtained in two different ways. Here, the parameters of ^R\hat{\mathcal{H}}_{R} are obtained from the relations described in Table 1. The overall shapes of the spectra of ^R\hat{\mathcal{H}}_{R} and ^circ\hat{\mathcal{H}}_{circ} look similar. On the other hand, the shift of the entire spectrum becomes as large as more than 200 MHz for Lc=350L_{c}=350 pH. To quantify the difference of the spectra between ^R\hat{\mathcal{H}}_{R} and ^circ\hat{\mathcal{H}}_{circ}, transition frequencies up to the third excited state numerically calculated from ^circ\hat{\mathcal{H}}_{circ} are fitted by ^R\hat{\mathcal{H}}_{R}. In the fitting, ^R\hat{\mathcal{H}}_{R} with an initial approximate set of parameters is numerically diagonalized and then the parameters, ω\omega, Δq\Delta_{q}, gg, and Ip(=ε/[2(Φx0.5Φ0)])I_{p}(=\varepsilon/[2(\Phi_{x}-0.5\Phi_{0})]), are varied to obtain the best fit. Figures 4c and d show that the same spectra of ^circ\hat{\mathcal{H}}_{circ} in Figs. 4a and b are well fitted by ^R\hat{\mathcal{H}}_{R}, but with a different parameter set.

Refer to caption
Figure 5: Parameters of ^R\hat{\mathcal{H}}_{R} obtained from the relations described in Table 1, (red solid lines) and by fitting the spectra of ^circ\hat{\mathcal{H}}_{circ} up to the third excited state to ^R\hat{\mathcal{H}}_{R} (black dashed lines) as a function of LcL_{c}. Panels a-d respectively correspond to the oscillator frequency ω\omega, qubit frequency Δq\Delta_{q}, coupling strength gg, and persistent current IpI_{p} in ^R\hat{\mathcal{H}}_{R}. e The average of the squares of the residuals in the least-squares method for obtaining the parameters of ^R\hat{\mathcal{H}}_{R} by fitting, [δω0i/(2π)]2¯\overline{[\delta\omega_{0i}/(2\pi)]^{2}} (ii = 1, 2, and 3).

Figure 5 shows parameters of ^R\hat{\mathcal{H}}_{R} obtained from the relations described in Table 1, and by fitting the spectra of ^circ\hat{\mathcal{H}}_{circ} up to the third excited state to ^R\hat{\mathcal{H}}_{R} as a function of LcL_{c}. The parameters of ^R\hat{\mathcal{H}}_{R} obtained in these two different ways become quite different from each other at large values of LcL_{c} due to contributions of the qubit’s higher energy levels. Interestingly, the average of the squares of the residuals in the least-squares method for obtaining the parameters of ^R\hat{\mathcal{H}}_{R} by fitting, [δω0i/(2π)]2¯\overline{[\delta\omega_{0i}/(2\pi)]^{2}} (ii = 1, 2, and 3), remains rather small, at most 2525~{}MHz2, which is consistent with the good fitting shown in Figs. 4c and d. In fact, the energy level structure of ^circ\hat{\mathcal{H}}_{circ} up to the seventh excited state can still be fitted well by the quantum Rabi Hamiltonian as shown in Supplementary Information SI .

Hamiltonian in the charge gauge

The circuit Hamiltonian ^circ\hat{\mathcal{H}}_{circ} is in the flux gauge, as the qubit-oscillator coupling term is of the form Φ^1Φ^2\hat{\Phi}_{1}\hat{\Phi}_{2}. The Hamiltonian can be transformed into the charge gauge:

^circ\displaystyle\hat{\mathcal{H}}_{circ}^{\prime} =\displaystyle= 𝒰^^circ𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\hat{\mathcal{H}}_{circ}\hat{\mathcal{U}}
=\displaystyle= (12C+LLC22CJL122)q^12+Φ^122LLC+q^222CJ+12(1LFQLLCL122)Φ^22EJcos(2πΦ^2ΦxΦ0)LLCCJL12q^1q^2\displaystyle\left(\frac{1}{2C}+\frac{L_{LC}^{2}}{2C_{J}L_{12}^{2}}\right)\hat{q}_{1}^{2}+\frac{\hat{\Phi}_{1}^{2}}{2L_{LC}}+\frac{\hat{q}_{2}^{2}}{2C_{J}}+\frac{1}{2}\left(\frac{1}{L_{FQ}}-\frac{L_{LC}}{L_{12}^{2}}\right)\hat{\Phi}_{2}^{2}-E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right)-\frac{L_{LC}}{C_{J}L_{12}}\hat{q}_{1}\hat{q}_{2}
=\displaystyle= ^1+^2+^12,\displaystyle\hat{\mathcal{H}}_{1}^{\prime}+\hat{\mathcal{H}}_{2}^{\prime}+\hat{\mathcal{H}}_{12}^{\prime},

where

𝒰^\displaystyle\hat{\mathcal{U}} =\displaystyle= exp(1iLLCL12Φ^2q^1),\displaystyle\exp\left(\frac{1}{i\hbar}\frac{L_{LC}}{L_{12}}\hat{\Phi}_{2}\hat{q}_{1}\right), (13)
^1\displaystyle\hat{\mathcal{H}}_{1}^{\prime} =\displaystyle= (12C+LLC22CJL122)q^12+Φ^122LLC,\displaystyle\left(\frac{1}{2C}+\frac{L_{LC}^{2}}{2C_{J}L_{12}^{2}}\right)\hat{q}_{1}^{2}+\frac{\hat{\Phi}_{1}^{2}}{2L_{LC}}, (14)
^2\displaystyle\hat{\mathcal{H}}_{2}^{\prime} =\displaystyle= 12CJq^22+12(1LFQLLCL122)Φ^22EJcos(2πΦ^2ΦxΦ0),\displaystyle\frac{1}{2C_{J}}\hat{q}_{2}^{2}+\frac{1}{2}\left(\frac{1}{L_{FQ}}-\frac{L_{LC}}{L_{12}^{2}}\right)\hat{\Phi}_{2}^{2}-E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right), (15)
^12\displaystyle\hat{\mathcal{H}}_{12}^{\prime} =\displaystyle= LLCCJL12q^1q^2.\displaystyle-\frac{L_{LC}}{C_{J}L_{12}}\hat{q}_{1}\hat{q}_{2}. (16)

The details of the calculations are given in Supplementary Information SI . As can be seen in Eq. (Hamiltonian in the charge gauge), the Hamiltonian ^circ\hat{\mathcal{H}}_{circ}^{\prime} can be separated into three parts: the first part ^1\hat{\mathcal{H}}_{1}^{\prime} consisting of the charge and flux operators of node 1, the second part ^2\hat{\mathcal{H}}_{2}^{\prime} consisting of the charge and flux operators of node 2, and the third part ^12\hat{\mathcal{H}}_{12}^{\prime} containing the product of the two charge operators. It is worth mentioning that the inductance in ^2\hat{\mathcal{H}}_{2}^{\prime} is equal to Lc+L2L_{c}+L_{2}, which is the inductance of the flux qubit in the separate treatment:

1LFQLLCL122\displaystyle\frac{1}{L_{FQ}}-\frac{L_{LC}}{L_{12}^{2}} =\displaystyle= 1L12+1Lg2Lg1L12(L12+Lg1)\displaystyle\frac{1}{L_{12}}+\frac{1}{L_{g2}}-\frac{L_{g1}}{L_{12}(L_{12}+L_{g1})} (17)
=\displaystyle= Lc+L1LcL1+LcL2+L1L2Lc2(LcL1+LcL2+L1L2)(Lc+L2)\displaystyle\frac{L_{c}+L_{1}}{L_{c}L_{1}+L_{c}L_{2}+L_{1}L_{2}}-\frac{L_{c}^{2}}{(L_{c}L_{1}+L_{c}L_{2}+L_{1}L_{2})(L_{c}+L_{2})}
=\displaystyle= (Lc+L1)(Lc+L2)Lc2(LcL1+LcL2+L1L2)(Lc+L2)\displaystyle\frac{(L_{c}+L_{1})(L_{c}+L_{2})-L_{c}^{2}}{(L_{c}L_{1}+L_{c}L_{2}+L_{1}L_{2})(L_{c}+L_{2})}
=\displaystyle= 1Lc+L2.\displaystyle\frac{1}{L_{c}+L_{2}}.
Refer to caption
Figure 6: Numerically calculated matrix elements j|q^2|i\left\langle j\left|\hat{q}_{2}\right|i\right\rangle for i=g,ei=g,e and a j=g,ej=g,e, b j=g,e,f,h,k,lj=g,e,f,h,k,l, where |f\left|f\right\rangle, |h\left|h\right\rangle, |k\left|k\right\rangle, and |l\left|l\right\rangle respectively represent the second, third, fourth, and fifth excited states of ^2\hat{\mathcal{H}}_{2}. Numerically calculated matrix elements j|Φ^2|i\left\langle j\left|\hat{\Phi}_{2}\right|i\right\rangle for i=g,ei=g,e and c j=g,ej=g,e, d j=g,e,f,h,k,lj=g,e,f,h,k,l. Note that the x axis ranges are smaller than those in Fig. 4.

The overall form of ^circ\hat{\mathcal{H}}^{\prime}_{circ} [Eq. (Hamiltonian in the charge gauge)] is almost the same as that of ^circ\hat{\mathcal{H}}_{circ} [Eq. (circuit Hamiltonian)], with the exception that the coupling term is of the form q^1q^2\hat{q}_{1}\hat{q}_{2} instead of Φ^1Φ^2\hat{\Phi}_{1}\hat{\Phi}_{2}. When we examined the mapping between ^2\hat{\mathcal{H}}_{2} and ^FQ\hat{\mathcal{H}}_{FQ}, we explained that the operator Φ^2\hat{\Phi}_{2} can be identified as σ^x\hat{\sigma}_{x}. Similarly, if we calculate the matrix elements of the operator q^2\hat{q}_{2} for the two lowest qubit states, we find that g|q^2|g=e|q^2|e=0\left\langle g\left|\hat{q}_{2}\right|g\right\rangle=\left\langle e\left|\hat{q}_{2}\right|e\right\rangle=0 and g|q^2|e=e|q^2|g0\left\langle g\left|\hat{q}_{2}\right|e\right\rangle=-\left\langle e\left|\hat{q}_{2}\right|g\right\rangle\neq 0 [Fig. 6a]. We can therefore identify the operator q^2\hat{q}_{2} as the qubit operator σ^y\hat{\sigma}_{y}.

It then seems natural to look for a mapping of ^circ\hat{\mathcal{H}}^{\prime}_{circ} to the variant of the quantum Rabi Hamiltonian:

^R/\displaystyle\hat{\mathcal{H}}_{R}^{\prime}/\hbar =\displaystyle= ω(a^a^+12)12(εσ^x+Δqσ^z)+igσ^y(a^a^)\displaystyle\omega\left(\hat{a}^{\dagger}\hat{a}+\frac{1}{2}\right)-\frac{1}{2}(\varepsilon\hat{\sigma}_{x}+\Delta_{q}\hat{\sigma}_{z})+ig^{\prime}\hat{\sigma}_{y}(\hat{a}-\hat{a}^{\dagger})
=\displaystyle= (^LC+^FQ+^coup)/.\displaystyle\left(\hat{\mathcal{H}}_{LC}+\hat{\mathcal{H}}_{FQ}+\hat{\mathcal{H}}_{coup}^{\prime}\right)/\hbar.

Compared with ^R\hat{\mathcal{H}}_{R}, only ^coup\hat{\mathcal{H}}_{coup} is replaced by ^coup\hat{\mathcal{H}}_{coup}^{\prime}. It should be noted, however, that there is a physical difference between the two Hamiltonians and that there is no simple mapping between them. The operator σ^x\hat{\sigma}_{x} in ^coup\hat{\mathcal{H}}_{coup} is the same as one of the Pauli operators in ^FQ\hat{\mathcal{H}}_{FQ}. In contrast, the operator σ^y\hat{\sigma}_{y} in ^coup\hat{\mathcal{H}}_{coup}^{\prime} is different from the two operators (σ^x\hat{\sigma}_{x} and σ^z\hat{\sigma}_{z}) in ^FQ\hat{\mathcal{H}}_{FQ}. As a result, ^R\hat{\mathcal{H}}_{R} and ^R\hat{\mathcal{H}}^{\prime}_{R} are physically different and for example produce different spectra.

In Fig. 4 we plot a few of the transition frequencies in the circuit’s spectrum along with the corresponding transition frequencies obtained from ^R\hat{\mathcal{H}}_{R}^{\prime}. When the parameters of ^R\hat{\mathcal{H}}_{R}^{\prime} are read off ^circ\hat{\mathcal{H}}_{circ}^{\prime}, similarly to what is shown in Table I, Figs. 4e and 4f show that the fitting is poor in most parts of the spectrum, even for the relatively weak coupling case Lc=20L_{c}=20 pH. Here, g=q1zpfq2maxLLC/(CJL12)g^{\prime}=q_{1zpf}q_{2max}L_{LC}/(C_{J}L_{12}), q1zpf=ωC/2q_{1zpf}=\sqrt{\hbar\omega^{\prime}C^{\prime}/2}, ω=1/LLC×(1/C)+(LLC2/CJL122)\omega^{\prime}=1/\sqrt{L_{LC}}\times\sqrt{(1/C)+(L_{LC}^{2}/C_{J}L_{12}^{2})}, 1/C=(1/C)+[LLC2/(CJL122)]1/C^{\prime}=(1/C)+[L_{LC}^{2}/(C_{J}L^{2}_{12})], and q2maxq_{2max} is numerically calculated as shown in Fig. 6a. Contrary to the open grey and red circles obtained from ^circ\hat{\mathcal{H}}_{circ}^{\prime}, the solid grey and red lines obtained from ^R\hat{\mathcal{H}}_{R}^{\prime} are larger and do not show the peaks and dips around the symmetry point. Even with a numerical optimization of the fitting parameters [Figs. 4g and 4h], only parts of the spectrum can be fitted well. In particular, the peaks and dips that occur in the spectrum at Φx/Φ0=0.5\Phi_{x}/\Phi_{0}=0.5 are not reproduced by ^R\hat{\mathcal{H}}_{R}^{\prime}.

Here it is useful to consider the two characteristic features in the spectrum, i.e. the avoided crossings and dispersive shifts, as being the result of energy shifts in the spectrum of an uncoupled circuit when the coupling term is added. The gap of an avoided level crossing, or in other words the Rabi splitting, is proportional to the matrix element of the coupling term between the relevant energy eigenstates of the uncoupled system. The dispersive shift of one energy level caused by another energy level is proportional to the square of the matrix element between the two energy eigenstates according to perturbation theory. The details of the dispersive shifts up to second order in perturbation theory are described in Supplementary information SI .

The difference between the spectra shown in the solid lines in Figs. 4a and 4e is attributed to the difference between the matrix elements of the qubit’s flux and charge operators, Φ^2\hat{\Phi}_{2} and q^2\hat{q}_{2}. The flux bias dependences of the numerically calculated matrix elements of the qubit’s charge and flux operators are shown in Fig. 6. Regarding the matrix elements of the qubit’s flux operator |j|Φ^2|i|\left|\left\langle j\left|\hat{\Phi}_{2}\right|i\right\rangle\right| (i=g,ei=g,e), those involving the higher qubit levels j=f,h,k,lj=f,h,k,l are smaller than those of j=g,ej=g,e. Here, |f\left|f\right\rangle, |h\left|h\right\rangle, |k\left|k\right\rangle, and |l\left|l\right\rangle respectively represent the second, third, fourth, and fifth excited states of ^2\hat{\mathcal{H}}_{2}. Regarding the matrix elements of the qubit’s charge operator |j|q^2|i|\left|\left\langle j\left|\hat{q}_{2}\right|i\right\rangle\right| (i=g,ei=g,e), on the other hand, some of those involving the higher qubit levels j=f,h,k,lj=f,h,k,l are significantly larger than those of j=g,ej=g,e in most of the flux bias range. This difference stems from the fact that the qubit Hamiltonian involves a double-well potential of the flux variables, as explained in Supplementary information SI .

The matrix elements g|Φ^2|e\left\langle g\left|\hat{\Phi}_{2}\right|e\right\rangle and e|Φ^2|g\left\langle e\left|\hat{\Phi}_{2}\right|g\right\rangle have a peak at Φ/Φ0=0.5\Phi/\Phi_{0}=0.5, which directly leads to the peak/dip feature at the symmetry point as shown in the solid lines in Fig. 4a. On the other hand, the matrix elements g|q^2|e\left\langle g\left|\hat{q}_{2}\right|e\right\rangle and e|q^2|g\left\langle e\left|\hat{q}_{2}\right|g\right\rangle are almost constant in the flux bias range of Fig. 6, which is consistent with the absence of a peak/dip feature at the symmetry point in the solid lines in Fig. 4e. Instead, the matrix elements h|q^2|g\left\langle h\left|\hat{q}_{2}\right|g\right\rangle and f|q^2|e\left\langle f\left|\hat{q}_{2}\right|e\right\rangle, which have somewhat similar flux-bias dependence to those of g|Φ^2|e\left\langle g\left|\hat{\Phi}_{2}\right|e\right\rangle and e|Φ^2|g\left\langle e\left|\hat{\Phi}_{2}\right|g\right\rangle, have peaks at Φ/Φ0=0.5\Phi/\Phi_{0}=0.5. As a result, for the flux gauge, the contribution from higher levels is just a small correction, while it cannot be neglected in the charge gauge. For this reason, the flux gauge turns out to be more convenient for purposes of mapping ^circ\hat{\mathcal{H}}_{circ} to the quantum Rabi Hamiltonian.

It is worth mentioning that we are dealing with an inductively coupled circuit, and one might think that the nature of the coupling, i.e. inductive or capacitive, will determine which gauge is more suitable, especially because the two gauges differ mainly by the form of the coupling term. We show that the deciding factor is the qubit rather than the coupling. As a result, if we have a flux qubit capacitively coupled to the oscillator, the flux gauge will be more suitable for the purpose of approximating the circuit Hamiltonian by the quantum Rabi Hamiltonian. It is also worth mentioning that our results should apply to fluxonium-resonator circuits, since the fluxonium Hamiltonian close to the degeneracy point also involves a double-well potential of the flux variables. It should be noted, however, that the fluxonium’s transition frequencies to the higher energy levels, ω02\omega_{02} and ω03\omega_{03}, are smaller than those of flux qubits, and, hence, the contribution from higher levels would be larger in fluxonium-resonator circuits.

We note here that the differences between the energy spectra obtained from ^circ\hat{\mathcal{H}}_{circ}, ^R\hat{\mathcal{H}}_{R}, and ^R\hat{\mathcal{H}}_{R}^{\prime} in the case ε=0\varepsilon=0, Δq=ω\Delta_{q}=\omega, and a quadratic-plus-quartic potential energy function were extensively studied in Ref. De Bernardis et al. (2018). Contrary to the case of Δq<ω\Delta_{q}<\omega, the spectra obtained from ^circ\hat{\mathcal{H}}_{circ} and ^R\hat{\mathcal{H}}_{R} are almost identical for the lowest energy levels in the whole range of g/ωg/\omega while that of ^R\hat{\mathcal{H}}_{R}^{\prime} deviates from the other two when g/ω0.1g/\omega\gtrsim 0.1. The Rabi splitting between the first two excited energy levels, which is symmetric and is proportional to gg, is clearly visible for g/ω0.1g/\omega\ll 0.1, while the dispersive shift due to the higher energy levels is proportional to g2g^{2}, and can be observed only when g/ω0.1g/\omega\gtrsim 0.1. We also note that for Lc=20L_{c}=20 pH, g/(2π)=0.043g^{\prime}/(2\pi)=0.043 GHz, which is much smaller than g/(2π)=0.424g/(2\pi)=0.424 GHz. On the other hand, the gap of the avoided level crossings plotted as lines in Fig. 4e is not much smaller than that in Fig. 4a considering g0.1gg^{\prime}\sim 0.1g. The gap of the avoided level crossings for ^R\hat{\mathcal{H}}_{R}^{\prime} is 2g2g^{\prime} while that for ^R\hat{\mathcal{H}}_{R} is 2gΔq/ω2g\Delta_{q}/\omega. The factor Δq/ω=0.206\Delta_{q}/\omega=0.206 partly explains why the avoided level crossings for ^R\hat{\mathcal{H}}_{R} and ^R\hat{\mathcal{H}}_{R}^{\prime} are not very different from each other.

expectation values of the photon number and the field operator

Refer to caption
Figure 7: Expectation values of photon number, flux, and current operators for a-d ^circ\hat{\mathcal{H}}_{circ} and for e-f ^circ\hat{\mathcal{H}}_{circ}^{\prime}. a, e Expectation numbers of photons in the oscillator in the ground state as a function of LcL_{c}. The solid curve corresponds to 0|^1/|0/(1/LLCC)0.5\left\langle 0\left|\hat{\mathcal{H}}_{1}/\hbar\right|0\right\rangle/(1/\sqrt{L_{LC}C})-0.5, while the dashed line in panel a indicates the simple expression (g/ω)2(g/\omega)^{2}, which is valid in the limit Δqω\Delta_{q}\ll\omega. b, f [d, h] Expectation values of fluxes 2πΦ^k/Φ02\pi\left\langle\hat{\Phi}_{k}\right\rangle/\Phi_{0} [Expectation values of currents I^k\left\langle\hat{I}_{k}\right\rangle] as functions of the flux bias Φx\Phi_{x} in the case Lc=350L_{c}=350 pH. Black and red colors respectively indicate k=1k=1 and 2, and solid and dashed lines indicate the ground and the excited states, respectively. c, g Expectation value of flux 2πΦ^1/Φ02\pi\left\langle\hat{\Phi}_{1}\right\rangle/\Phi_{0} in the ground state as a function of LcL_{c} at Φx/Φ0=0.498\Phi_{x}/\Phi_{0}=0.498.

One of the most paradoxical features of ^R\hat{\mathcal{H}}_{R} in the deep-strong-coupling regime is the non-negligible number of photons in the oscillator in the ground state Ashhab and Nori (2010). In terms of the creation and annihilation operators, the photon number operator is a^a^\hat{a}^{\dagger}\hat{a}. Considering the mapping between ^circ\hat{\mathcal{H}}_{circ} and ^R\hat{\mathcal{H}}_{R}, the photon number operator in ^circ\hat{\mathcal{H}}_{circ} is ^1/(1/LLCC)0.5\hat{\mathcal{H}}_{1}/(1/\sqrt{L_{LC}C})-0.5. As shown in Fig. 7a, a nonzero number of photons in the oscillator is obtained. Another paradoxical feature of ^R\hat{\mathcal{H}}_{R} in the deep-strong-coupling regime is a non-negligible expectation value of the field operator a^+a^\left\langle\hat{a}+\hat{a}^{\dagger}\right\rangle when the qubit flux bias is Φx0.5Φ0\Phi_{x}\neq 0.5\Phi_{0}. The corresponding operator in ^circ\hat{\mathcal{H}}_{circ} is Φ^1/(LLCIzpf)\hat{\Phi}_{1}/(L_{LC}I_{zpf}). Fig. 7b shows the expectation values of the fluxes, Φ^1\left\langle\hat{\Phi}_{1}\right\rangle and Φ^2\left\langle\hat{\Phi}_{2}\right\rangle, as functions of the flux bias Φx\Phi_{x} for the ground and the first excited states in the case Lc=350L_{c}=350 pH. At Φx/Φ00.5\Phi_{x}/\Phi_{0}\neq 0.5, a nonzero expectation value Φ^1\left\langle\hat{\Phi}_{1}\right\rangle is demonstrated. One may suspect that a nonzero Φ^1\left\langle\hat{\Phi}_{1}\right\rangle is unphysical because it might result in a nonzero DC current through the inductor of an LC oscillator in an energy eigenstate. The relation between the flux and current operators is explicitly given in the circuit model [Fig. 1a]:

(Lc+L1LcLcLc+L2)(I^1I^2)=(Φ^1Φ^2),\displaystyle\begin{pmatrix}L_{c}+L_{1}&L_{c}\\ L_{c}&L_{c}+L_{2}\\ \end{pmatrix}\begin{pmatrix}\hat{I}_{1}\\ \hat{I}_{2}\\ \end{pmatrix}=\begin{pmatrix}\hat{\Phi}_{1}\\ \hat{\Phi}_{2}\\ \end{pmatrix}, (19)

where I^k\hat{I}_{k} (k=1,2k=1,2) is defined as the operator of the current flowing from the ground node to node kk. In the case Lc=0L_{c}=0, Eq. (19) reduces to I^1=Φ^1/L1\hat{I}_{1}=\hat{\Phi}_{1}/L_{1} and I^2=Φ^2/L2\hat{I}_{2}=\hat{\Phi}_{2}/L_{2} which means that the operator Φ^1\hat{\Phi}_{1} can indeed be understood as a current operator. As shown in Fig. 7c, at Φx/Φ0=0.498\Phi_{x}/\Phi_{0}=0.498, the expectation value 2πΦ^1/Φ02\pi\left\langle\hat{\Phi}_{1}\right\rangle/\Phi_{0} in the ground state is proportional to LcL_{c} and is zero only when Lc=0L_{c}=0. Fig. 7d shows the expectation values of the currents I^1\left\langle\hat{I}_{1}\right\rangle and I^2\left\langle\hat{I}_{2}\right\rangle as functions of Φx\Phi_{x} in the case Lc=350L_{c}=350 pH. The expectation value I^2\left\langle\hat{I}_{2}\right\rangle is nonzero at Φx/Φ00.5\Phi_{x}/\Phi_{0}\neq 0.5, while I^1\left\langle\hat{I}_{1}\right\rangle is exactly zero at all values of Φx\Phi_{x}. In this way, although Φ^1\left\langle\hat{\Phi}_{1}\right\rangle is nonzero in the case Lc0L_{c}\neq 0, an unphysical DC current through the inductor of an LC oscillator is not predicted.

Next, we consider the case of ^circ\hat{\mathcal{H}}_{circ}^{\prime}. Since unitary transformations do not change the eigenenergies of Hamiltonians, ^circ\hat{\mathcal{H}}_{circ} and ^circ\hat{\mathcal{H}}_{circ}^{\prime} have exactly the same eigenenergies. On the other hand, the photon-number and flux operators do not commute with the gauge transformation:

𝒰^^1𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\hat{\mathcal{H}}_{1}\hat{\mathcal{U}} =\displaystyle= 𝒰^(q^122C+Φ^122LLC)𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\left(\frac{\hat{q}^{2}_{1}}{2C}+\frac{\hat{\Phi}^{2}_{1}}{2L_{LC}}\right)\hat{\mathcal{U}} (20)
=\displaystyle= [q^122C+(Φ^1+LLCL12Φ^2)212LLC],\displaystyle\left[\frac{\hat{q}^{2}_{1}}{2C}+\left(\hat{\Phi}_{1}+\frac{L_{LC}}{L_{12}}\hat{\Phi}_{2}\right)^{2}\frac{1}{2L_{LC}}\right],

and

𝒰^Φ^1𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\hat{\Phi}_{1}\hat{\mathcal{U}} =\displaystyle= Φ^1+LLCL12Φ^2.\displaystyle\hat{\Phi}_{1}+\frac{L_{LC}}{L_{12}}\hat{\Phi}_{2}. (21)

The corresponding photon number operator in ^circ\hat{\mathcal{H}}_{circ}^{\prime} is ^1/(ω)0.5\hat{\mathcal{H}}_{1}^{\prime}/(\hbar\omega^{\prime})-0.5, where ω=1/LLC×(1/C)+(LLC2/CJL122)\omega^{\prime}=1/\sqrt{L_{LC}}\times\sqrt{(1/C)+(L_{LC}^{2}/C_{J}L_{12}^{2})} is the resonance frequency of ^1\hat{\mathcal{H}}_{1}^{\prime}. The corresponding field operator in ^circ\hat{\mathcal{H}}_{circ}^{\prime} is Φ^1/(LLCIzpf)\hat{\Phi}_{1}/(L_{LC}I_{zpf}). As shown in Fig. 7e, the expectation number of photons in the charge gauge is much smaller than that in the flux gauge, and the non-negligible expectation number of photons in the oscillator in the ground state arises only in the flux-gauge. We note here that the resonance frequency of ^1\hat{\mathcal{H}}_{1}^{\prime} is larger than 1/LLCC6.031/\sqrt{L_{LC}C}\simeq 6.03~{}GHz: ω/(2π)=15.66\omega^{\prime}/(2\pi)=15.66 GHz in the case Lc=350L_{c}=350 pH. We can see that the expectation value of the operator Φ^1\hat{\Phi}_{1} in the case Lc=350L_{c}=350 pH is zero at all values of Φx\Phi_{x} [Fig. 7f], which are also different from the case of the flux-gauge Hamiltonian. The expectation value of the current operator I^1\hat{I}_{1} is zero at all values of Φx\Phi_{x}, which is true in the flux gauge as well.

discussion

We have derived the Hamiltonian of a superconducting circuit that comprises a single-Josephson-junction flux qubit and an LC oscillator using the standard quantization procedure. Excluding the qubit’s higher energy levels, the derived circuit Hamiltonian takes the form of the quantum Rabi Hamiltonian. We show that the Hamiltonian derived from the separate treatment, where the circuit is assumed to be naively divided into the two well-defined components, has the same form as the circuit Hamiltonian, and the inductances in the separate treatment approach those of the circuit Hamiltonian as LcL_{c} approaches 0. The qubit’s higher energy levels mainly cause a negative shift of the entire spectrum, but the energy level structure can still be fitted well by the quantum Rabi Hamiltonian even when the qubit-oscillator circuit is in the deep-strong-coupling regime. We also show that although the circuit Hamiltonian can be transformed to a Hamiltonian containing a capacitive coupling term, the resulting circuit Hamiltonian cannot be approximated by the capacitive-coupling variant of the quantum Rabi Hamiltonian.

methods

As a simple example, let us consider ^1\hat{\mathcal{H}}_{1} in the main text:

^1\displaystyle\hat{\mathcal{H}}_{1} =\displaystyle= q^122C+Φ^122LLC\displaystyle\frac{\hat{q}_{1}^{2}}{2C}+\frac{\hat{\Phi}_{1}^{2}}{2L_{LC}} (22)
=\displaystyle= 4ECn^2+12ELϕ^2,\displaystyle 4E_{C}\hat{n}^{2}+\frac{1}{2}E_{L}\hat{\phi}^{2},

where, EC=e2/(2C)E_{C}=e^{2}/(2C), EL=[Φ0/(2π)]2/LLCE_{L}=[\Phi_{0}/(2\pi)]^{2}/L_{LC}, and ϕ^\hat{\phi} and n^\hat{n} are operators of dimensionless magnetic flux and charge, respectively, and satisfy [ϕ^,n^]=i[\hat{\phi},\hat{n}]=i. Using the relation ϕ^=(1/i)/n\hat{\phi}=-(1/i)\partial/\partial n, the Hamiltonian can be rewritten as

^1\displaystyle\hat{\mathcal{H}}_{1} =\displaystyle= 4ECn^212EL2n2.\displaystyle 4E_{C}\hat{n}^{2}-\frac{1}{2}E_{L}\frac{\partial^{2}}{\partial n^{2}}. (23)

Let us calculate wavefunctions ψ(n)\psi(n) and their eigenenergies EE of this Hamiltonian. We expand the wavefunction with plane waves as

ψ(n)=kψkeikn2nmax.\displaystyle\psi(n)=\sum_{k}\psi_{k}\frac{\mathrm{e}^{ikn}}{\sqrt{2n_{max}}}. (24)

Here, 2nmax2n_{max} is the length of the nn-space. Considering the periodic boundary condition ψ(nmax)=ψ(nmax)\psi(-n_{max})=\psi(n_{max}), the wave number kk is given by

k=2π2nmaxη,(η=0,±1,±2,).\displaystyle k=\frac{2\pi}{2n_{max}}\eta,\,(\eta=0,\pm 1,\pm 2,\dots). (25)

In numerical calculations in this work, we have used 32 waves for the qubit and 64 waves for the oscillator. Then, the equation for determining ψk\psi_{k} and EE is obtained as

^1ψ(n)\displaystyle\hat{\mathcal{H}}_{1}\psi(n) =\displaystyle= Eψ(n)\displaystyle E\psi(n)
k[4ECn2+12ELk2]ψkeikn2nmax\displaystyle\sum_{k^{\prime}}\left[4E_{C}n^{2}+\frac{1}{2}E_{L}k^{\prime 2}\right]\psi_{k^{\prime}}\frac{\mathrm{e}^{ik^{\prime}n}}{2n_{max}} =\displaystyle= Ekψkeikn2nmax\displaystyle E\sum_{k^{\prime}}\psi_{k^{\prime}}\frac{\mathrm{e}^{ik^{\prime}n}}{2n_{max}}
nmaxnmax𝑑neikn2nmaxk[4ECn2+12ELk2]ψkeikn2nmax\displaystyle\int_{-n_{max}}^{n_{max}}dn\frac{\mathrm{e}^{-ikn}}{\sqrt{2n_{max}}}\sum_{k^{\prime}}\left[4E_{C}n^{2}+\frac{1}{2}E_{L}k^{\prime 2}\right]\psi_{k^{\prime}}\frac{\mathrm{e}^{ik^{\prime}n}}{2n_{max}} =\displaystyle= nmaxnmax𝑑neikn2nmaxEkψkeikn2nmax\displaystyle\int_{-n_{max}}^{n_{max}}dn\frac{\mathrm{e}^{-ikn}}{\sqrt{2n_{max}}}E\sum_{k^{\prime}}\psi_{k^{\prime}}\frac{\mathrm{e}^{ik^{\prime}n}}{2n_{max}}
k[4ECfkk(n2)+δk,k12ELk2]ψk\displaystyle\sum_{k^{\prime}}\left[4E_{C}f_{k-k^{\prime}}(n^{2})+\delta_{k,k^{\prime}}\frac{1}{2}E_{L}k^{2}\right]\psi_{k^{\prime}} =\displaystyle= Eψk,\displaystyle E\psi_{k}, (26)

where

fkk(n2)12nmaxnmaxnmax𝑑nei(kk)nn2.\displaystyle f_{k-k^{\prime}}(n^{2})\equiv\frac{1}{2n_{max}}\int_{-n_{max}}^{n_{max}}dn\mathrm{e}^{-i(k-k^{\prime})n}n^{2}. (27)

The set of wavefunctions and eigenenergies are obtained by solving Eq. (26). Numerically, we can get fkk(n2)f_{k-k^{\prime}}(n^{2}) by the fast Fourier transform (FFT) for discretized nn-space. After solving the eigenvalue equation in Eq. (26), the wavefunctions ψ(n)\psi(n) are also obtained from ψk\psi_{k} by the FFT. For the calculation of ^2\hat{\mathcal{H}}_{2} in the main text, we use the following equation instead of Eq. (26),

k{4ECJfkk(n2)+δk,k[EJcos(kkx)+12ELFQk2]}ψk\displaystyle\sum_{k^{\prime}}\left\{4E_{CJ}f_{k-k^{\prime}}(n^{2})+\delta_{k,k^{\prime}}\left[-E_{J}\cos\left(k-k_{x}\right)+\frac{1}{2}E_{LFQ}k^{2}\right]\right\}\psi_{k^{\prime}} =\displaystyle= Eψk,\displaystyle E\psi_{k}, (28)

where ECJ=e2/(2CJ)E_{CJ}=e^{2}/(2C_{J}), ELFQ=[Φ0/(2π)]2/LFQE_{LFQ}=[\Phi_{0}/(2\pi)]^{2}/L_{FQ}, and kx=2πΦx/Φ0k_{x}=2\pi\Phi_{x}/\Phi_{0}. Note that a fluxonium-resonator circuit that has the identical circuit diagram but different circuit parameters was numerically diagonalized using the harmonic oscillator basis Smith et al. (2016).

Acknowledgements.
We are grateful to M. Devoret for valuable discussions. This work was supported by Japan Society for the Promotion of Science (JSPS) Grants-in-Aid for Scientific Research (KAKENHI) (No. JP19H01831 and JP19K03693), Japan Science and Technology Agency (JST) Precursory Research for Embryonic Science and Technology (PRESTO) (Grant No. JPMJPR1767), JST Core Research for Evolutionary Science and Technology (CREST) (Grant No. JPMJCR1775), and MEXT Quantum Leap Flagship Program (MEXT Q-LEAP) Grant Number JPMXS0120319794.

Author Contributions

FY conceived the main idea of the paper. FY, SA, and MB constructed equations of the paper. FY, SA, TF, and MB conducted numerical calculations. FY wrote the manuscript with feedback from all authors. FY and KS supervised the project.

Supplementary Information

S1 Fitting of circ\mathcal{H}_{circ} up to the seventh excited state

Refer to caption
Figure S1: (a) Transition frequencies of the qubit-oscillator circuit up to the seventh excited state for Lc=350L_{c}=350 pH. Numerically calculated transition frequencies from ^circ\hat{\mathcal{H}}_{circ} are plotted as circles, while the results of the fitting by ^R\hat{\mathcal{H}}_{R} are plotted as lines. The parameters of ^R\hat{\mathcal{H}}_{R} obtained by fitting the spectra of ^circ\hat{\mathcal{H}}_{circ} to ^R\hat{\mathcal{H}}_{R} are ω/(2π)=6.054\omega/(2\pi)=6.054 GHz, Δq/(2π)=2.133\Delta_{q}/(2\pi)=2.133 GHz, g/(2π)=7.562g/(2\pi)=7.562 GHz, and Ip=282.2I_{p}=282.2 nA. Red, green, blue, cyan, magenta, yellow, and black colors indicate transition frequencies ω01\omega_{01}, ω02\omega_{02}, ω03\omega_{03}, ω04\omega_{04}, ω05\omega_{05}, ω06\omega_{06}, and ω07\omega_{07}, respectively. (b), (c), (d) The same spectra with smaller range of frequency around 3ω3\omega, 2ω2\omega, and ω\omega.

Transition frequencies of the qubit-oscillator circuit numerically calculated from ^circ\hat{\mathcal{H}}_{circ} up to the seventh excited state are plotted in Figure S1 for Lc=350L_{c}=350 pH using the same circuit parameters as in Fig. 3. The calculated spectra can still be fitted well by ^R\hat{\mathcal{H}}_{R}. In the fitting, ^R\hat{\mathcal{H}}_{R} are numerically diagonalized and then the parameters ω\omega, Δq\Delta_{q}, gg, and IpI_{p}, are varied to obtain the best fit. The average of the squares of the residuals in the least-squares method of obtaining the parameters of ^R\hat{\mathcal{H}}_{R} by fitting, [δω0i/(2π)]2¯\overline{[\delta\omega_{0i}/(2\pi)]^{2}} (ii = 1,2,3,4,5,6, and 7), is 152152~{}MHz2. This value is almost 10 times larger than the case of the fitting up to the third excited state, but still within a moderate level considering that the largest transition frequency is more than 20 GHz.

S2 Gauge transformation

The circuit Hamiltonian in the flux gauge ^circ\hat{\mathcal{H}}_{circ} can be written as

^circ\displaystyle\hat{\mathcal{H}}_{circ} =\displaystyle= q^122C+q^222CJEJcos(2πΦ^2ΦxΦ0)+Φ^122LLC+Φ^222LFQΦ^1Φ^2L12\displaystyle\frac{\hat{q}_{1}^{2}}{2C}+\frac{\hat{q}_{2}^{2}}{2C_{J}}-E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right)+\frac{\hat{\Phi}_{1}^{2}}{2L_{LC}}+\frac{\hat{\Phi}_{2}^{2}}{2L_{FQ}}-\frac{\hat{\Phi}_{1}\hat{\Phi}_{2}}{L_{12}} (S1)
=\displaystyle= q^122C+q^222CJEJcos(2πΦ^2ΦxΦ0)\displaystyle\frac{\hat{q}_{1}^{2}}{2C}+\frac{\hat{q}_{2}^{2}}{2C_{J}}-E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right)
+12LLC(Φ^1LLCL12Φ^2)2+12(1LFQLLCL122)Φ^22.\displaystyle+\frac{1}{2L_{LC}}\left(\hat{\Phi}_{1}-\frac{L_{LC}}{L_{12}}\hat{\Phi}_{2}\right)^{2}+\frac{1}{2}\left(\frac{1}{L_{FQ}}-\frac{L_{LC}}{L_{12}^{2}}\right)\hat{\Phi}_{2}^{2}.

The Hamiltonian can now be transformed into the charge gauge rather easily. First we note that the unitary operator

𝒰^\displaystyle\hat{\mathcal{U}} =\displaystyle= exp(1iαΦ^2q^1)\displaystyle\exp\left(\frac{1}{i\hbar}\alpha\hat{\Phi}_{2}\hat{q}_{1}\right) (S2)

transforms the flux and charge operators as follows:

𝒰^Φ^1𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\hat{\Phi}_{1}\hat{\mathcal{U}} =\displaystyle= Φ^1+αΦ^2,\displaystyle\hat{\Phi}_{1}+\alpha\hat{\Phi}_{2}, (S3)
𝒰^Φ^2𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\hat{\Phi}_{2}\hat{\mathcal{U}} =\displaystyle= Φ^2,\displaystyle\hat{\Phi}_{2}, (S4)
𝒰^q^1𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\hat{q}_{1}\hat{\mathcal{U}} =\displaystyle= q^1,\displaystyle\hat{q}_{1}, (S5)
𝒰^q^2𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\hat{q}_{2}\hat{\mathcal{U}} =\displaystyle= q^2αq^1.\displaystyle\hat{q}_{2}-\alpha\hat{q}_{1}. (S6)

Then, if we set α=LLC/L12\alpha=L_{LC}/L_{12}, the circuit Hamiltonian is transformed into

^circ\displaystyle\hat{\mathcal{H}}_{circ}^{\prime} =\displaystyle= 𝒰^^circ𝒰^\displaystyle\hat{\mathcal{U}}^{\dagger}\hat{\mathcal{H}}_{circ}\hat{\mathcal{U}} (S7)
=\displaystyle= q^122C+12CJ(q^2LLCL12q^1)2\displaystyle\frac{\hat{q}_{1}^{2}}{2C}+\frac{1}{2C_{J}}\left(\hat{q}_{2}-\frac{L_{LC}}{L_{12}}\hat{q}_{1}\right)^{2}
EJcos(2πΦ^2ΦxΦ0)+Φ^122LLC+12(1LFQLLCL122)Φ^22\displaystyle-E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right)+\frac{\hat{\Phi}_{1}^{2}}{2L_{LC}}+\frac{1}{2}\left(\frac{1}{L_{FQ}}-\frac{L_{LC}}{L_{12}^{2}}\right)\hat{\Phi}_{2}^{2}
=\displaystyle= (12C+LLC22CJL122)q^12+Φ^122LLC+12CJq^22\displaystyle\left(\frac{1}{2C}+\frac{L_{LC}^{2}}{2C_{J}L_{12}^{2}}\right)\hat{q}_{1}^{2}+\frac{\hat{\Phi}_{1}^{2}}{2L_{LC}}+\frac{1}{2C_{J}}\hat{q}_{2}^{2}
+12(1LFQLLCL122)Φ^22EJcos(2πΦ^2ΦxΦ0)LLCCJL12q^1q^2.\displaystyle+\frac{1}{2}\left(\frac{1}{L_{FQ}}-\frac{L_{LC}}{L_{12}^{2}}\right)\hat{\Phi}_{2}^{2}-E_{J}\cos\left(2\pi\frac{\hat{\Phi}_{2}-\Phi_{x}}{\Phi_{0}}\right)-\frac{L_{LC}}{C_{J}L_{12}}\hat{q}_{1}\hat{q}_{2}.

S3 Energy shifts up to second order in perturbation theory

To investigate the reason for the poor reproducibility of the characteristic spectra of ^circ\hat{\mathcal{H}}_{circ}^{\prime} by ^R\hat{\mathcal{H}}_{R}^{\prime}, we calculate the energy shifts of the four lowest energy levels using perturbation theory. Let us consider the Hamiltonian ^circ,λ^1+^2+λ^12\hat{\mathcal{H}}_{circ,\lambda}\equiv\hat{\mathcal{H}}_{1}+\hat{\mathcal{H}}_{2}+\lambda\hat{\mathcal{H}}_{12}. Here, we assume that the eigenenergies and the eigenstates of ^circ,λ\hat{\mathcal{H}}_{circ,\lambda} can be described by a power series in λ\lambda as Eni,λ=k=0λkEni(k)E_{ni,\lambda}=\sum_{k=0}^{\infty}\lambda^{k}E_{ni}^{(k)} and |niλ=k=0λk|ni(k)\left|ni\right\rangle_{\lambda}=\sum_{k=0}^{\infty}\lambda^{k}\left|ni^{(k)}\right\rangle, where Eni(0)E_{ni}^{(0)} and |ni(0)\left|ni^{(0)}\right\rangle are the eigenenergies and the eigenstates of the non-interacting Hamiltonian ^1+^2\hat{\mathcal{H}}_{1}+\hat{\mathcal{H}}_{2}, nn is the number of photons in the oscillator, and i=g,ei=g,e represents the eigenstate of ^2\hat{\mathcal{H}}_{2}. By taking the eigenenergies and the eigenstates of ^circ,λ\hat{\mathcal{H}}_{circ,\lambda} and multiplying by mj(k)|\left\langle mj^{(k)}\right| from the left, the correction terms can be written as

Eni(1)\displaystyle E_{ni}^{(1)} =\displaystyle= ni(0)|^12|ni(0)\displaystyle\left\langle ni^{(0)}\right|\hat{\mathcal{H}}_{12}\left|ni^{(0)}\right\rangle (S8)

and

Eni(2)\displaystyle E_{ni}^{(2)} =\displaystyle= m,j|mj(0)|^12|ni(0)|2Eni(0)Emj(0)m,jχni,mj.\displaystyle\sum_{m,j}\frac{\left|\left\langle mj^{(0)}\left|\hat{\mathcal{H}}_{12}\right|ni^{(0)}\right\rangle\right|^{2}}{E_{ni}^{(0)}-E_{mj}^{(0)}}\equiv\sum_{m,j}\hbar\chi_{ni,mj}. (S9)

Here, χni,mj\chi_{ni,mj} is the energy shift of the state |ni\left|ni\right\rangle due to the interaction with the state |mj\left|mj\right\rangle, where m=0,1,2,m=0,1,2,\dots, j=g,e,f,h,j=g,e,f,h,\dots, and |f\left|f\right\rangle and |h\left|h\right\rangle respectively represent the second and third excited states of ^2\hat{\mathcal{H}}_{2}. We find that Eni(1)=0E_{ni}^{(1)}=0 for all the combinations of n=0,1n=0,1 and i=g,ei=g,e, and the total energy shifts up to the second-order perturbation χni\chi_{ni} are simply determined by the second-order correction terms: χni=m,jχni,mj\chi_{ni}=\sum_{m,j}\chi_{ni,mj}. We also find that the third- or higher-order correction terms are much smaller than the second-order correction term: the numerator of the nnth order correction term is the product of nn matrix elements of H^12\hat{H}_{12}, which are at most several hundred MHz, while the denominator of the nnth order correction term is the product of n1n-1 energy-level differences, which are in the order of GHz or more.

Refer to caption
Figure S2: Nonzero energy level shifts χni,mj\chi_{ni,mj} of the non-interacting Hamiltonian ^1+^2\hat{\mathcal{H}}_{1}+\hat{\mathcal{H}}_{2} due to the term ^12\hat{\mathcal{H}}_{12} and the state |mj\left|mj\right\rangle for the four lowest eigenstates |ni=\left|ni\right\rangle= (a) |0g\left|0g\right\rangle, (b) |0e\left|0e\right\rangle, (c) |1g\left|1g\right\rangle, and (d) |1e\left|1e\right\rangle for Lc=20L_{c}=20 pH. (e) the net change of the transition frequencies δω01g=χ1gχ0g\delta\omega_{01}^{g}=\chi_{1g}-\chi_{0g} and δω01e=χ1eχ0e\delta\omega_{01}^{e}=\chi_{1e}-\chi_{0e}, obtained by combining the shifts in Panels (a)-(d).

Figures S2(a)-(d) show nonzero energy level shifts χni,mj\chi_{ni,mj} of the non-interacting Hamiltonian ^1+^2\hat{\mathcal{H}}_{1}+\hat{\mathcal{H}}_{2} due to the interaction term ^12\hat{\mathcal{H}}_{12} for the lowest four eigenstates |ni=|0g\left|ni\right\rangle=\left|0g\right\rangle, |0e\left|0e\right\rangle, |1g\left|1g\right\rangle, and |1e\left|1e\right\rangle of the non-interacting Hamiltonian ^1+^2\hat{\mathcal{H}}_{1}+\hat{\mathcal{H}}_{2} for Lc=20L_{c}=20 pH. Although some energy shifts χni,mj\chi_{ni,mj} appear to be equal to zero everywhere, they are in fact finite but extremely small. The level shifts χni,mj\chi_{ni,mj} that involve the higher qubit levels j=f,hj=f,h plotted using dashed lines are all close to zero, and the net level shifts are almost completely determined by the eigenstates involving the two lowest energy levels of ^2\hat{\mathcal{H}}_{2}. Figure S2(e) shows the net change of the transition frequencies δω01g=χ1gχ0g\delta\omega_{01}^{g}=\chi_{1g}-\chi_{0g} and δω01e=χ1eχ0e\delta\omega_{01}^{e}=\chi_{1e}-\chi_{0e} from the resonance frequency of the oscillator described by ^1\hat{\mathcal{H}}_{1}, which reproduce the peaks and dips that occur in the spectrum at Φx/Φ0=0.5\Phi_{x}/\Phi_{0}=0.5 shown in Fig. 4a in the main text.

Refer to caption
Figure S3: Nonzero energy level shifts χni,mj\chi_{ni,mj} of the non-interacting Hamiltonian ^1+^2\hat{\mathcal{H}}_{1}^{\prime}+\hat{\mathcal{H}}_{2}^{\prime} due to the term ^12\hat{\mathcal{H}}_{12}^{\prime} and the state |mj\left|mj\right\rangle for the four lowest eigenstates |ni=\left|ni\right\rangle= (a) |0g\left|0g\right\rangle, (b) |0e\left|0e\right\rangle, (c) |1g\left|1g\right\rangle, and (d) |1e\left|1e\right\rangle for Lc=20L_{c}=20 pH. (e) the net changes of the transition frequencies δω01g=χ1gχ0g\delta\omega_{01}^{g}=\chi_{1g}-\chi_{0g} and δω01e=χ1eχ0e\delta\omega_{01}^{e}=\chi_{1e}-\chi_{0e}, obtained by combining the shifts in Panels (a)-(d). Numerically calculated transition frequencies of ^circ\hat{\mathcal{H}}_{circ}, ω02ω\omega_{02}-\omega^{\prime} and ω13ω\omega_{13}-\omega^{\prime}, are plotted as circles, where ω\omega^{\prime} is the resonance frequency of ^circ\hat{\mathcal{H}}^{\prime}_{circ}.

We now consider the case of the circuit Hamiltonian ^circ\hat{\mathcal{H}}_{circ}^{\prime}. Figures S3(a)-(d) show nonzero energy level shifts χni,mj\chi_{ni,mj} of the non-interacting Hamiltonian ^1+^2\hat{\mathcal{H}}_{1}^{\prime}+\hat{\mathcal{H}}_{2}^{\prime} due to the interaction with the state |mj\left|mj\right\rangle for eigenstates |ni\left|ni\right\rangle (n=0,1n=0,1, i=g,ei=g,e) for Lc=20L_{c}=20 pH. Figure S3(e) shows the net change of the transition frequencies δω01g=χ1gχ0g\delta\omega_{01}^{g}=\chi_{1g}-\chi_{0g} and δω01e=χ1eχ0e\delta\omega_{01}^{e}=\chi_{1e}-\chi_{0e} from the resonance frequency of the oscillator described by ^1\hat{\mathcal{H}}^{\prime}_{1}. Although the qualitative agreement is still not excellent, the peaks and dips that occur in the spectrum at Φx/Φ0\Phi_{x}/\Phi_{0} are qualitatively reproduced. The agreement of δω01g\delta\omega_{01}^{g} is better than that of δω01e\delta\omega_{01}^{e}. Contrary to the case of the circuit Hamiltonian ^circ\hat{\mathcal{H}}_{circ}, the level shifts χni,mj\chi_{ni,mj} that involve the higher qubit levels j=f,hj=f,h plotted using dashed lines are larger than those including j=g,ej=g,e in most of the flux bias conditions, and the level shifts are mainly determined by the third and the fourth lowest energy levels of ^2\hat{\mathcal{H}}_{2}^{\prime}. These results explain the difference between the spectra of ^circ\hat{\mathcal{H}}_{circ}^{\prime} and those of ^R\hat{\mathcal{H}}_{R}^{\prime}, in which ^FQ\hat{\mathcal{H}}^{\prime}_{FQ} contains only the lowest two energy levels, as shown in Fig. 4e in the main text: The spectrum represented by solid grey and red lines obtained from ^R\hat{\mathcal{H}}_{R}^{\prime} are similar to those of an uncoupled qubit-oscillator circuit, whereas the spectrum represented by the open grey and red circles obtained from ^circ\hat{\mathcal{H}}_{circ}^{\prime} show the peaks and dips around the symmetry point and the overall frequency is lower.

S4 Matrix elements of the charge and flux operators

In Section S3, the energy level shifts of the non-interacting Hamiltonian caused by the interaction term were studied using perturbation theory. The energy shifts of the states |ni\left|ni\right\rangle are given in Eq. (S9). Considering that the states |ni(0)\left|ni^{(0)}\right\rangle and |mj(0)\left|mj^{(0)}\right\rangle are separable, the following matrix elements can be written as products of matrix elements of the oscillator and the qubit:

mj(0)|Φ^1Φ^2|ni(0)=m|Φ^1|nj|Φ^2|i\displaystyle\left\langle mj^{(0)}\left|\hat{\Phi}_{1}\hat{\Phi}_{2}\right|ni^{(0)}\right\rangle=\left\langle m\left|\hat{\Phi}_{1}\right|n\right\rangle\left\langle j\left|\hat{\Phi}_{2}\right|i\right\rangle (S10)

and

mj(0)|q^1q^2|ni(0)=m|q^1|nj|q^2|i.\displaystyle\left\langle mj^{(0)}\right|\hat{q}_{1}\hat{q}_{2}\left|ni^{(0)}\right\rangle=\left\langle m\left|\hat{q}_{1}\right|n\right\rangle\left\langle j\left|\hat{q}_{2}\right|i\right\rangle. (S11)

The matrix elements of the oscillator operators Φ^1\hat{\Phi}_{1} and q^1\hat{q}_{1} are analytically given as

m|Φ^1|n\displaystyle\left\langle m\left|\hat{\Phi}_{1}\right|n\right\rangle =\displaystyle= LLCIzpfm|(a^+a^)|n\displaystyle L_{LC}I_{zpf}\left\langle m\left|(\hat{a}+\hat{a}^{\dagger})\right|n\right\rangle (S12)
=\displaystyle= LLCIzpf(nδm,n1+n+1δm,n+1)\displaystyle L_{LC}I_{zpf}(\sqrt{n}\delta_{m,n-1}+\sqrt{n+1}\delta_{m,n+1})

and

m|q^1|n\displaystyle\left\langle m\left|\hat{q}_{1}\right|n\right\rangle =\displaystyle= iCVzpfm|(a^a^)|n\displaystyle-iCV_{zpf}\left\langle m\left|(\hat{a}-\hat{a}^{\dagger})\right|n\right\rangle (S13)
=\displaystyle= iCVzpf(nδm,n1n+1δm,n+1).\displaystyle-iCV_{zpf}(\sqrt{n}\delta_{m,n-1}-\sqrt{n+1}\delta_{m,n+1}).

Only matrix elements that satisfy m=n±1m=n\pm 1 are nonzero. The matrix elements of the qubit operators Φ^2\hat{\Phi}_{2} and q^2\hat{q}_{2} for i=g,ei=g,e and j=g,e,f,h,k,lj=g,e,f,h,k,l, where |k\left|k\right\rangle and |l\left|l\right\rangle respectively represent the fourth and fifth excited states of ^2\hat{\mathcal{H}}_{2}, are numerically calculated as shown in Fig. 6 in the main text. Note that the matrix elements of the qubit operators at ε=0\varepsilon=0 and with a quadratic-plus-quartic potential energy function were studied in Ref. De Bernardis et al. (2018).

Regarding the matrix elements |j|Φ^2|i|\left|\left\langle j\left|\hat{\Phi}_{2}\right|i\right\rangle\right| (i=g,ei=g,e), those involving the higher qubit levels j=f,h,k,lj=f,h,k,l are smaller than those of j=g,ej=g,e. Together with the fact that the energy difference |Eni(0)Emj(0)|\left|E_{ni}^{(0)}-E_{mj}^{(0)}\right| of j=f,h,k,lj=f,h,k,l is larger than those of j=g,ej=g,e, which appears in the denominator of Eq. (S9), this result explains why the level shifts are almost completely determined by the eigenstates involving the two lowest energy levels of ^2\hat{\mathcal{H}}_{2}. Regarding the matrix elements |j|q^2|i|\left|\left\langle j\left|\hat{q}_{2}\right|i\right\rangle\right| (i=g,ei=g,e) on the other hand, some of those involving the higher qubit levels j=f,h,k,lj=f,h,k,l are significantly larger than those of j=g,ej=g,e in most of the flux bias range. Although the energy difference |Eni(0)Emj(0)|\left|E_{ni}^{(0)}-E_{mj}^{(0)}\right| is larger than those of j=g,ej=g,e, the large matrix elements result in a situation where the level shifts are mainly determined by the third and fourth lowest energy levels of ^2\hat{\mathcal{H}}^{\prime}_{2}. We note here that the matrix elements involving the levels j=k,lj=k,l are smaller than those involving the levels j=f,hj=f,h and the energy difference is larger than those of j=f,hj=f,h. Hence, the level shifts caused by higher levels with j=k,lj=k,l are smaller than those with j=f,hj=f,h.

The relation between the different matrix elements can be intuitively understood using an analogy to a basic quantum physics problem. The qubit Hamiltonians ^2\hat{\mathcal{H}}_{2} and ^2\hat{\mathcal{H}}^{\prime}_{2} have the same form as the Hamiltonian of a single particle in a trapping potential, with Φ^2\hat{\Phi}_{2} and q^2\hat{q}_{2} playing the roles of the position and momentum variables, respectively. Around the symmetry point, the trapping potential is a double-well potential. The energy eigenstates in the double-well potential are superpositions of the energy eigenstates in the two separate wells, which are approximately harmonic oscillator potentials. We therefore have superpositions of harmonic oscillator states whose centers are separated by a distance that is larger than their widths.

Refer to caption
Figure S4: Numerically calculated wavefunctions of the eigenstates of ^2\hat{\mathcal{H}}_{2}, Ψi\Psi_{i}, as well as the combinations Ψi×(2πΦ2/Φ0)\Psi_{i}\times(2\pi\Phi_{2}/\Phi_{0}), and (Φ0/2πΦ2)Ψi/Φ2(\Phi_{0}/2\pi\Phi_{2})\partial\Psi_{i}/\partial\Phi_{2} (i=g,e,f,h)(i=g,e,f,h), as functions of 2πΦ2/Φ02\pi\Phi_{2}/\Phi_{0} for Φx/Φ0=0.5\Phi_{x}/\Phi_{0}=0.5.

Figures S4(a)-(d) show wave functions of the eigenstates of ^2\hat{\mathcal{H}}_{2}, Ψi\Psi_{i} (i=g,e,f,h)(i=g,e,f,h), as functions of Φ2\Phi_{2} at the flux bias point Φx/Φ0=0.5\Phi_{x}/\Phi_{0}=0.5. Matrix elements of the position operator can be written using the wave-function representation:

j|Φ^2|i=𝑑Φ2(ΨjΦ2Ψi).\displaystyle\left\langle j\left|\hat{\Phi}_{2}\right|i\right\rangle=\int d\Phi_{2}(\Psi_{j}^{*}\Phi_{2}\Psi_{i}). (S14)

Matrix elements that combine a pair of functions Ψj\Psi_{j} and Φ2Ψi\Phi_{2}\Psi_{i} [Figs. S4(e)-(h)] with similar shapes are large. At the symmetry point, the matrix elements |e|Φ^2|g|\left|\left\langle e\left|\hat{\Phi}_{2}\right|g\right\rangle\right| = |g|Φ^2|e|\left|\left\langle g\left|\hat{\Phi}_{2}\right|e\right\rangle\right| \sim |h|Φ^2|f|\left|\left\langle h\left|\hat{\Phi}_{2}\right|f\right\rangle\right| = |f|Φ^2|h|\left|\left\langle f\left|\hat{\Phi}_{2}\right|h\right\rangle\right| are large and the others are small or zero. Away from the symmetry point, the matrix elements |g|Φ^2|g|\left|\left\langle g\left|\hat{\Phi}_{2}\right|g\right\rangle\right|, |e|Φ^2|e|\left|\left\langle e\left|\hat{\Phi}_{2}\right|e\right\rangle\right|, |f|Φ^2|f|\left|\left\langle f\left|\hat{\Phi}_{2}\right|f\right\rangle\right|, and |h|Φ^2|h|\left|\left\langle h\left|\hat{\Phi}_{2}\right|h\right\rangle\right| also become large. The matrix elements between states with i=g,ei=g,e and states with j=f,hj=f,h are therefore small compared to matrix elements involving only gg and ee. Similarly, the matrix elements of the momentum operator can be written using the wave-function representation:

j|q^2|i=𝑑Φ2(ΨjiΨiΦ2).\displaystyle\left\langle j\left|\hat{q}_{2}\right|i\right\rangle=\int d\Phi_{2}\left(\Psi_{j}^{*}\frac{\hbar}{i}\frac{\partial\Psi_{i}}{\partial\Phi_{2}}\right). (S15)

Matrix elements that have a pair of functions Ψj\Psi_{j} and Ψi/Φ2\partial\Psi_{i}/\partial\Phi_{2} [Figs. S4(i)-(l)] with similar shapes are large. At the symmetry point, the matrix elements |h|q^2|g|\left|\left\langle h\left|{\color[rgb]{0,0,0}\hat{q}_{2}}\right|g\right\rangle\right| = |g|q^2|h|\left|\left\langle g\left|{\color[rgb]{0,0,0}\hat{q}_{2}}\right|h\right\rangle\right| \sim |f|q^2|e|\left|\left\langle f\left|{\color[rgb]{0,0,0}\hat{q}_{2}}\right|e\right\rangle\right| = |e|q^2|f|\left|\left\langle e\left|{\color[rgb]{0,0,0}\hat{q}_{2}}\right|f\right\rangle\right| are large and the others are small or zero. Away from the symmetry point, the matrix elements |f|q^2|g|\left|\left\langle f\left|{\color[rgb]{0,0,0}\hat{q}_{2}}\right|g\right\rangle\right|, |g|q^2|f|\left|\left\langle g\left|{\color[rgb]{0,0,0}\hat{q}_{2}}\right|f\right\rangle\right|, |h|q^2|e|\left|\left\langle h\left|{\color[rgb]{0,0,0}\hat{q}_{2}}\right|e\right\rangle\right|, and |e|q^2|h|\left|\left\langle e\left|{\color[rgb]{0,0,0}\hat{q}_{2}}\right|h\right\rangle\right| also become large. The matrix elements between states with i=g,ei=g,e and states with j=f,hj=f,h are therefore large compared to matrix elements involving only gg and ee.

References

  • Nakamura et al. (1999) Y. Nakamura, Y. A. Pashkin,  and J. S. Tsai, Nature 398, 786 (1999).
  • Vool and Devoret (2017) U. Vool and M. Devoret, International Journal of Circuit Theory and Applications 45, 897 (2017).
  • Nakamura et al. (1997) Y. Nakamura, C. D. Chen,  and J. S. Tsai, Phys. Rev. Lett. 79, 2328 (1997).
  • Orlando et al. (1999) T. P. Orlando, J. E. Mooij, L. Tian, C. H. van der Wal, L. S. Levitov, S. Lloyd,  and J. J. Mazo, Phys. Rev. B 60, 15398 (1999).
  • Vion et al. (2002) D. Vion, A. Aassime, A. Cottet, P. Joyez, H. Pothier, C. Urbina, D. Esteve,  and M. H. Devoret, Science 296, 886 (2002).
  • Martinis et al. (2002) J. M. Martinis, S. Nam, J. Aumentado,  and C. Urbina, Physical review letters 89, 117901 (2002).
  • 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).
  • Pashkin et al. (2003) Y. A. Pashkin, T. Yamamoto, O. Astafiev, Y. Nakamura, D. Averin,  and J. S. Tsai, Nature 421, 823 (2003).
  • Chiorescu et al. (2004) I. Chiorescu, P. Bertet, K. Semba, Y. Nakamura, C. J. P. M. Harmans,  and J. E. Mooij, Nature 431, 159 (2004).
  • Blais et al. (2004) A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin,  and R. J. Schoelkopf, Phys. Rev. A 69, 062320 (2004).
  • Niskanen et al. (2006) A. O. Niskanen, K. Harrabi, F. Yoshihara, Y. Nakamura,  and J. S. Tsai, Phys. Rev. B 74, 220503(R) (2006).
  • Niemczyk et al. (2010) T. Niemczyk, F. Deppe, H. Huebl, E. P. Menzel, F. Hocke, M. J. Schwarz, J. J. Garcia-Ripoll, D. Zueco, T. Hümmer, E. Solano, A. Marx,  and R. Gross, Nat. Phys. 6, 772 (2010).
  • Forn-Diaz et al. (2010) P. Forn-Diaz, J. Lisenfeld, D. Marcos, J. J. Garcia-Ripoll, E. Solano, C. J. P. M. Harmans,  and J. E. Mooij, Phys. Rev. Lett. 105, 237001 (2010).
  • Yoshihara et al. (2017a) F. Yoshihara, T. Fuse, S. Ashhab, K. Kakuyanagi, S. Saito,  and K. Semba, Nat. Phys. 13, 44 (2017a).
  • Yoshihara et al. (2017b) F. Yoshihara, T. Fuse, S. Ashhab, K. Kakuyanagi, S. Saito,  and K. Semba, Phys. Rev. A 95, 053824 (2017b).
  • Yoshihara et al. (2018) F. Yoshihara, T. Fuse, Z. Ao, S. Ashhab, K. Kakuyanagi, S. Saito, T. Aoki, K. Koshino,  and K. Semba, Phys. Rev. Lett. 120, 183601 (2018).
  • Rabi (1937) I. I. Rabi, Phys. Rev. 51, 652 (1937).
  • Jaynes and Cummings (1963) E. T. Jaynes and F. W. Cummings, Proc. IEEE 51, 89 (1963).
  • Braak (2011) D. Braak, Phys. Rev. Lett. 107, 100401 (2011).
  • Bourassa et al. (2009) J. Bourassa, J. M. Gambetta, A. A. Abdumalikov Jr, O. Astafiev, Y. Nakamura,  and A. Blais, Physical Review A 80, 032109 (2009).
  • Peropadre et al. (2013) B. Peropadre, D. Zueco, D. Porras,  and J. J. García-Ripoll, Phys. Rev. Lett. 111, 243602 (2013).
  • Smith et al. (2016) W. C. Smith, A. Kou, U. Vool, I. M. Pop, L. Frunzio, R. J. Schoelkopf,  and M. H. Devoret, Phys. Rev. B 94, 144507 (2016).
  • Manucharyan et al. (2017) V. E. Manucharyan, A. Baksic,  and C. Ciuti, Journal of Physics A: Mathematical and Theoretical 50, 294001 (2017).
  • De Bernardis et al. (2018) D. De Bernardis, P. Pilar, T. Jaako, S. De Liberato,  and P. Rabl, Phys. Rev. A 98, 053819 (2018).
  • Chiorescu et al. (2003) I. Chiorescu, Y. Nakamura, C. J. P. M. Harmans,  and J. E. Mooij, Science 299, 1869 (2003).
  • Robertson et al. (2006) T. L. Robertson, B. L. T. Plourde, P. A. Reichardt, T. Hime, C.-E. Wu,  and J. Clarke, Phys. Rev. B 73, 174526 (2006).
  • Peltonen et al. (2018) J. Peltonen, P. Coumou, Z. Peng, T. Klapwijk, J. S. Tsai,  and O. Astafiev, Scientific reports 8, 1 (2018).
  • Hazard et al. (2019) T. M. Hazard, A. Gyenis, A. Di Paolo, A. T. Asfaw, S. A. Lyon, A. Blais,  and A. A. Houck, Phys. Rev. Lett. 122, 010504 (2019).
  • Grünhaupt et al. (2019) L. Grünhaupt, M. Spiecker, D. Gusenkova, N. Maleeva, S. T. Skacel, I. Takmakov, F. Valenti, P. Winkel, H. Rotzinger, W. Wernsdorfer, et al., Nature materials 18, 816 (2019).
  • Manucharyan et al. (2009) V. E. Manucharyan, J. Koch, L. I. Glazman,  and M. H. Devoret,  326, 113 (2009).
  • Consani and Warburton (2020) G. Consani and P. A. Warburton, New Journal of Physics 22, 053040 (2020).
  • Stokes and Nazir (2019) A. Stokes and A. Nazir, Nature communications 10, 1 (2019).
  • Roth et al. (2019) M. Roth, F. Hassler,  and D. P. DiVincenzo, Physical Review Research 1, 033128 (2019).
  • Di Stefano et al. (2019) O. Di Stefano, A. Settineri, V. Macrì, L. Garziano, R. Stassi, S. Savasta,  and F. Nori, Nature Physics 15, 803 (2019).
  • Johansson et al. (2013) J. R. Johansson, P. D. Nation,  and F. Nori, Comp. Phys. Commun. 184, 1234 (2013).
  • (36) See Supplementary Information for (i) fitting of ^circ\hat{\mathcal{H}}_{circ} up to the seventh excited state, (ii) gauge transformation, (iii) energy shifts up to second order in perturbation theory, (iv) matrix elements of the charge and flux operators. .
  • Ashhab and Nori (2010) S. Ashhab and F. Nori, Phys. Rev. A 81, 042311 (2010).