arxiv: v1 [quant-ph] 13 Dec 2017

Similar documents
arxiv: v2 [quant-ph] 7 May 2018

4-3 New Regime of Circuit Quantum Electro Dynamics

Entanglement Control of Superconducting Qubit Single Photon System

SUPPLEMENTARY INFORMATION

Theoretical design of a readout system for the Flux Qubit-Resonator Rabi Model in the ultrastrong coupling regime

arxiv: v1 [quant-ph] 13 Jul 2018

Distributing Quantum Information with Microwave Resonators in Circuit QED

Controlling the Interaction of Light and Matter...

Cavity Quantum Electrodynamics with Superconducting Circuits

Coherent Coupling between 4300 Superconducting Flux Qubits and a Microwave Resonator

Cavity Quantum Electrodynamics (QED): Coupling a Harmonic Oscillator to a Qubit

Circuit Quantum Electrodynamics

THE RABI HAMILTONIAN IN THE DISPERSIVE REGIME

10.5 Circuit quantum electrodynamics

10.5 Circuit quantum electrodynamics

Doing Atomic Physics with Electrical Circuits: Strong Coupling Cavity QED

arxiv: v2 [quant-ph] 31 Jan 2017

Superconducting Qubits Lecture 4

Circuit Quantum Electrodynamics. Mark David Jenkins Martes cúantico, February 25th, 2014

Quantum Optics with Electrical Circuits: Strong Coupling Cavity QED

Circuit QED: A promising advance towards quantum computing

arxiv:cond-mat/ v1 [cond-mat.mes-hall] 30 Aug 2006

Thermal Excitation of Multi-Photon Dressed States in Circuit Quantum Electrodynamics

Dispersive Readout, Rabi- and Ramsey-Measurements for Superconducting Qubits

arxiv: v1 [quant-ph] 7 Jul 2009

Qubit-photon interactions in a cavity: Measurement-induced dephasing and number splitting

Simple Scheme for Realizing the General Conditional Phase Shift Gate and a Simulation of Quantum Fourier Transform in Circuit QED

Qubit-oscillator dynamics in the dispersive regime: Analytical theory beyond the rotating-wave approximation

Quantum optics and quantum information processing with superconducting circuits

Quantum computation and quantum optics with circuit QED

Two-photon probe of the Jaynes-Cummings model and symmetry breaking in circuit QED

Introduction to Circuit QED

The Impact of the Pulse Phase Deviation on Probability of the Fock States Considering the Dissipation

Circuit quantum electrodynamics : beyond the linear dispersive regime

Entanglement dynamics in a dispersively coupled qubit-oscillator system

Dynamical Casimir effect in superconducting circuits

Dipole-coupling a single-electron double quantum dot to a microwave resonator

Superposition of two mesoscopically distinct quantum states: Coupling a Cooper-pair box to a large superconducting island

Quantum Optics with Electrical Circuits: Circuit QED

Dressed relaxation and dephasing in a strongly driven two-level system

Solid State Physics IV -Part II : Macroscopic Quantum Phenomena

arxiv:cond-mat/ v1 [cond-mat.mes-hall] 27 Feb 2007

Superconducting Resonators and Their Applications in Quantum Engineering

Superconducting Qubits

Quantum Optics and Quantum Informatics FKA173

Non-linear driving and Entanglement of a quantum bit with a quantum readout

INTRODUCTION TO SUPERCONDUCTING QUBITS AND QUANTUM EXPERIENCE: A 5-QUBIT QUANTUM PROCESSOR IN THE CLOUD

arxiv: v2 [cond-mat.mes-hall] 19 Oct 2010

arxiv: v3 [quant-ph] 18 Jul 2011

Superconducting Qubits Coupling Superconducting Qubits Via a Cavity Bus

arxiv: v2 [quant-ph] 19 Sep 2015

Circuit QED with electrons on helium:

CIRCUIT QUANTUM ELECTRODYNAMICS WITH ELECTRONS ON HELIUM

Exploring the quantum dynamics of atoms and photons in cavities. Serge Haroche, ENS and Collège de France, Paris

Photon shot noise dephasing in the strong-dispersive limit of circuit QED

8 Quantized Interaction of Light and Matter

Quantum-information processing with circuit quantum electrodynamics

Supplementary Figure 1 Level structure of a doubly charged QDM (a) PL bias map acquired under 90 nw non-resonant excitation at 860 nm.

Driving Qubit Transitions in J-C Hamiltonian

Supercondcting Qubits

arxiv: v1 [quant-ph] 23 Oct 2014

Engineering the quantum probing atoms with light & light with atoms in a transmon circuit QED system

Electrical quantum engineering with superconducting circuits

Cavity quantum electrodynamics for superconducting electrical circuits: An architecture for quantum computation

Condensed Matter Without Matter Quantum Simulation with Photons

Quantum walks on circles in phase space via superconducting circuit quantum electrodynamics

Strong-coupling Circuit QED

arxiv: v2 [quant-ph] 8 May 2018

Synthesising arbitrary quantum states in a superconducting resonator

arxiv: v1 [cond-mat.supr-con] 5 May 2015

Population Dynamics and Emission Spectrum of a Cascade Three-Level Jaynes Cummings Model with Intensity-Dependent Coupling in a Kerr-like Medium

Entangled Macroscopic Quantum States in Two Superconducting Qubits

Towards new states of matter with atoms and photons

Quantum non-demolition measurement of a superconducting two-level system

Single Photon Nonlinear Optics with Cavity enhanced Quantum Electrodynamics

arxiv: v1 [quant-ph] 6 Oct 2011

Remote entanglement of transmon qubits

Supplementary information for Quantum delayed-choice experiment with a beam splitter in a quantum superposition

Quantum simulation with superconducting circuits

Cavity QED with Rydberg Atoms Serge Haroche, Collège de France & Ecole Normale Supérieure, Paris

Strongly Driven Semiconductor Double Quantum Dots. Jason Petta Physics Department, Princeton University

Superconducting quantum bits. Péter Makk

arxiv: v1 [quant-ph] 21 Sep 2015

Superconducting Qubits. Nathan Kurz PHYS January 2007

Electrical Quantum Engineering with Superconducting Circuits

Multiphoton quantum Rabi oscillations in ultrastrong cavity QED

Preparation of macroscopic quantum superposition states of a cavity field via coupling to a superconducting charge qubit

arxiv: v3 [cond-mat.mes-hall] 25 Feb 2011

Controllable cross-kerr interaction between microwave photons in circuit quantum electrodynamics

Multipartite Entanglement Generation Assisted by Inhomogeneous Coupling

arxiv: v2 [quant-ph] 2 Mar 2015

Quantum Many-Body Phenomena in Arrays of Coupled Cavities

Code: ECTS Credits: 6. Degree Type Year Semester

arxiv: v1 [quant-ph] 12 Sep 2007

Introduction to Cavity QED

PROTECTING QUANTUM SUPERPOSITIONS IN JOSEPHSON CIRCUITS

Synthesizing arbitrary photon states in a superconducting resonator

Quantum Many Body Physics with Strongly Interacting Matter and Light. Marco Schiro Princeton Center for Theoretical Science

Amplification, entanglement and storage of microwave radiation using superconducting circuits

dc measurements of macroscopic quantum levels in a superconducting qubit structure with a time-ordered meter

Transcription:

Inversion of qubit energy levels in qubit-oscillator circuits in the deep-strong-coupling regime arxiv:171.539v1 [quant-ph] 13 Dec 17 F. Yoshihara, 1, T. Fuse, 1 Z. Ao, S. Ashhab, 3 K. Kakuyanagi, 4 S. Saito, 4 T. Aoki, K. Koshino, 5 and K. Semba 1, 1 Advanced ICT Institute, National Institute of Information and Communications Technology, 4--1, Nukuikitamachi, Koganei, Tokyo 184-8795, Japan Department of Applied Physics, Waseda University, 3-4-1, Ookubo, Shinjuku-ku, Tokyo, Japan 3 Qatar Environment and Energy Research Institute, Hamad Bin Khalifa University, Qatar Foundation, Doha, Qatar 4 NTT Basic Research Laboratories, NTT Corporation, 3-1 Morinosato-Wakamiya, Atsugi, Kanagawa 43-198, Japan 5 College of Liberal Arts and Sciences, Tokyo Medical and Dental University, -8-3, Kounodai, Ichikawa, Chiba 7-87, Japan. (Dated: December 15, 17) We report on experimentally measured light shifts of superconducting flux qubits deep-stronglycoupled to an LC oscillator, where the coupling constant is comparable to the qubit s transition frequency and the oscillator s resonance frequency. By using two-tone spectroscopy, the energies of the six-lowest levels of the coupled circuits are determined. We find a huge Lamb shift that exceeds 9% of the bare qubit frequencies and inversion of the qubits ground and excited states when there is a finite number of photons in the oscillator. Our experimental results agree with theoretical predictions based on the quantum Rabi model. According to quantum theory, the vacuum electromagnetic field has half photon fluctuations, which cause several physical phenomena such as the Lamb shift [1]. A cavity can enhance the interaction between the atom and the electromagnetic field inside the cavity, and more precise measurements on the influence of the vacuum becomes possible. A cavity/circuit-quantum electrodynamics (QED) system in the weak and strong coupling regimes is well described by the Jaynes-Cummings Hamiltonian [, 3]. In the strong coupling regime, when the cavity s resonance frequency ω is on resonance with the atom s transiton frequency, the vacuum Rabi splitting [4 6] and oscillation [7, 8] have been observed. In the off-resonance case, the Lamb shift [9 11] caused by the vacuum fluctuations, and the ac-stark shift proportional to the photon number in the cavity, were observed [1 13]. In the so-called ultrastrong coupling regime [14, 15], where the coupling constant g becomes around 1% of and ω, and the deep-strong-coupling regime [16, 17], where g is comparable or larger than and ω, the rotating wave approximation used in the Jaynes-Cummings Hamiltonian breaks down and the system should be described by the quantum Rabi model Hamiltonian [18 ]. In these regimes, the light shifts of an atom could nonmonotonously change as g increases, and the amount of the shift is not proportional to the photon number in the cavity [1]. In this work, to study the light shift in the case of g ω, we use qubit-oscillator circuits which comprises a superconducting flux qubit and an LC oscillator inductively coupled to each other by sharing a loop of Josephson junctions that serves as a coupler. By using two-tone spectroscopy [], energies of the six lowest energy eigenstates were measured, and photon-numberdependent qubit frequencies were evaluated. We find (a) C Φ q L c L ω p ω d (b) g g ω, g ω g, Δ Δ 1 Δ e ω, e ω e, FIG. 1. (a) Circuit diagram. A superconducting flux qubit (red and black) and a superconducting LC oscillator (blue and black) are inductively coupled to each other by sharing an inductance (black). (b) The diagram of six lowest energylevels of a qubit-oscillator circuit. The energy eigenstates in (i = g, e and n =, 1,, ) indicates that the qubit is in g the ground or e the excited state and the oscillator is in the n-photon Fock state. The arrows indicate transition frequencies between energy eigenstates, and also means that the transitions are allowed. Here, n (n =, 1, ) is the photon-number-dependent qubit frequency. Lamb shifts over 9% of the bare qubit frequency and inversion of the qubit s ground and the excited states when there is a finite number of photons in the oscillator. The circuits that we use for this work can be described as a composite system that comprises a flux qubit [3] inductively coupled to a lumped-element LC oscillator as shown in Fig. 1(a). The circuit design is similar to those of Refs. [16, 17]. The qubit-oscillator circuit is described by the Hamiltonian Ĥ = ( ˆσ x + εˆσ z ) + ωâ â + gˆσ z (â + â ). (1) The first two terms represent the energy of the flux qubit written in the basis of two states with persistent currents e

flowing in opposite directions around the qubit loop [3], q and q. ˆσ x,z are Pauli operators. The parameters and ε are the tunnel splitting and the energy bias between q and q, where ε can be controlled by the flux bias through the qubit loop Φ q. The third term in Eq. (1) represents the energy of the LC oscillator, where ω = 1/ (L + L c )C [see Fig. 1(a)] is the resonance frequency, and â and â are the creation and annihiration operators, respectively. The fourth term in Eq. (1) represents the coupling energy. At ε =, the Hamiltonian in Eq. (1) is equivalent to that of the quantum Rabi model: Ĥ Rabi = ˆσ z + ωâ â + gˆσ x (â + â ). () Note that the qubit part of the Hamiltonian is now written in the qubit s energy eigenbasis. When g =, the energy eigenstates are product states: g q n o q + q n o gn g=, e q n o q q n o ẽn g=, (3) where g q and e q are the qubit s ground and excited states; the bonding and anti-bonding states of q and q, and n o is the n-photon Fock state of the oscillator. Here, we name the energy eigenstates in (i = g, e). For nonzero values of g, the energy eigenstates of ĤRabi cannot be written by Eq. (3) any more. However, in the limit ω, the energy eigenstates are well described by Schrödinger-cat-like entangled states between persistentcurrent states of the qubit and displaced Fock states of the oscillator ˆD(±α) n o [1]: gn ω q ˆD( g ω ) n o + q ˆD( g ω ) n o, ẽn ω q ˆD( g ω ) n o q ˆD( g ω ) n o.(4) Here, ˆD(α) = exp(αâ α â) is the displacement operator, and α is the amount of displacement. Note that the displaced vacuum state ˆD(α) o is the coherent state α o = exp( α /) n= αn n o / n. Note also that Eq. (4) is valid for any value of g and reduces to Eq. (3) when g =. The photon-number-dependent qubit frequency n (g/ω) ω is defined as the energy difference between the energy eigenstates gn ω and ẽn ω, and can be calculated as n (g/ω) ω = ẽn ĤRabi ẽn ω gn ĤRabi gn ω n o ˆD ( g/ω) ˆD(g/ω) n o = exp( g /ω )L n (4g /ω ). (5) Here, L n are Laguerre polynomials; L (x) = 1, L 1 (x) = 1 x, L (x) = (x 4x + )/, and so on. The difference between n and the bare qubit frequency can be considered as the n-photon ac-stark shifts n. In particular, is referred to as the Lamb shift. Note that the Bloch-Siegert shift [4, 5], the contribution from the counter-rotating terms, is included in the n-photon ac- Stark shifts. Since L = 1, considerable Lamb shift is expected when g becomes comparable to ω. Considering that L n has n zeros, i.e. points where L n (x) is equal to zero, n (x) also has n zeros, and hence, become both positive and negative. In other words, the qubit s ground and excited states exchange their roles everytime when n = due to n-photon ac-stark shift. Remarkably, n is reduced from by a factor that is determined by the overlap integral of the interaction-caused displaced n- photon Fock states of the oscillator in opposite directions as seen in the second line of Eq. (5). By considering the symmetry of ĤRabi, the expressions of energy eigenstates in can be applied to any parameter sets of Ĥ Rabi [6]. Also, it is found that Eq. (5) can describe the qualitative behavior of n as long as < ω. These two facts allow us to uniquely determine n including its sign for all the parameter sets in this work [6]. To determine parameters of the qubit-oscillator circuits (, ω, and g), spectroscopy was performed by measuring the transmission spectrum through the transmission line that is inductively coupled to the LC oscillator [Fig. 1(a)]. In total, nine sets of parameters (A I) in five circuits were evaluated. The shared inductance of the circuit for set A is a superconducting lead, while that of the circuits for sets B I is a loop of Josephson junctions, where eight flux bias points in four circuits were used. Therefore, much larger g is expected for sets B I. When the frequency of the probe signal ω p matches the frequency ω kl of a transition k l, where stands for the ground state and k with k 1 stands for the kth excited state of the coupled circuit, the transmission amplitude decreases, provided that the transition matrix element k (â + â ) l is not. Note that for nonzero values of ε, we have labeled the energy eigenstates using a single integer k instead of the label in used above. Figure shows the amplitudes of the transmission spectra S1 meas (ε, ω p )/S bg 1 (ω p) for sets A and H. Here, ω p is the probe frequency, and S1 meas (ε, ω p ) and S bg 1 (ω p) are respectively measured and background transmissions [6]. The parameters are obtained from fitting the experimentally measured resonance frequencies to those numerically calculated by diagonalizing Ĥ with, ω and g treated as fitting parameters. In Fig., the right panels show the calculated transition frequencies superimposed on the measured spectra. In Fig. (a), one can see the splitting of and 1 3 transition frequencies around ε =, known as the qubit-state-dependent frequency shifts of the oscillator. From the fitting, the

3.5 1 1 1. 13 3 ω g π π π π A 1.46 6.365.4 1.36 (1.35) B 1.1 6.96 5.41.33 (.9) C.9 6.88 5.59.193 (.189) D 3.93 5.8 5.8.54 (.539) E 4.88 5.3 5.37.67 (.67) F 4.71 5. 5.46.538 (.54) G 3.53 5.63 5.58.375 (.379) H 1.68 6.345 7.7.17 (.1) I 1.61 6.335 7.48.99 (.99) ωp/π (GHz) 6.4 6.38 (a) 6.36 6.34 1 ωp/π (GHz) 6.6 1 1 ε/π (GHz) (b) 1 ε/π (GHz) 6.4 6. 6. 5 ε/π (GHz) FIG.. Measured transmission spectra for two qubitoscillator circuits as functions of the qubit s energy bias ε and probe frequency ωp. The color scheme is chosen such that the lowest point in each spectrum is red and the highest point is blue. The right panels show the transition frequencies calculated from the Hamiltonian to fit experimental data. The black, gray, orange, pink, and red lines correspond to the transitions i 1i, i i, i 3i, 1i i, and 1i 3i, respectively. The parameters are obtained to be (a) /π = 1.46 GHz, ω/π = 6.365 GHz, and g/π =.4 GHz corresponding to set A; (b) /π = 1.68 GHz, ω/π = 6.345 GHz, and g/π = 7.7 GHz corresponding to set H. These parameters give (a) g/ω =.65 and (b) 1.15. parameters are obtained to be /π = 1.46 GHz, ω/π = 6.365 GHz, and g/π =.4 GHz. The spectrum shown in Fig. (b) looks qualitatively different from that in (a) as discussed in Ref. [17]. The parameters are obtained to be /π = 1.68 GHz, ω/π = 6.345 GHz, and g/π = 7.7 GHz. Here, g is larger than both and ω, indicating that the circuit is in the deep-strong-coupling regime [g & max(ω, ω/)] [1]. The parameters from all the sets are summarized in Table I. To obtain the photon-number-dependent qubit frequency n (n =, 1, ), at least five transition frequencies out of seven allowed transitions [Fig. 1(b)] are necessary. However, in each spectrum at ε =, we see only two signals at frequencies ωf e e correspondg,f and ωe, f i f and ei e i, e respecing to the transitions gi tively. Since the base temperature of the dilution fridge is around mk, only the two lowest energy levels f gi and e are populated, and should be taken into account as ei e caninitial states. Moreover, the transition f gi ei not be measured directly since the transition frequency /π is below the frequency range of the measurement setup (4 GHz to 8 GHz). To access transitions other than f gi f i and e e ei i, two-tone spectroscopy was used, where a drive signal with frequency ωd is applied while the trans- π -.13 (-.13) -.6 (-.59).56 (.64).96 (1.18) 1.5 (1.87).8 (.834).5 (.53).493 (.53) TABLE I. Parameters of qubit-oscillator circuits in GHz., ω, and g are obtained from the (single-tone) transmission spectra. The upper line of n (n =, 1, ) are obtained from two-tone transmission spectra, while the lower line in () are numerically calculated values using H Rabi and the parameters, ω, and g. 6.7 ωp/π (GHz) 5 6.68.7.8.9 1. (a) 6.66 6.64 6.6 ωp/π (GHz) 5 ε/π (GHz) ωp/π (GHz) 5 1 π 1.15 (1.13) -.45 (-.448) -.41 (-.41) -1.51 (-1.53) -1.746 (-1.741) -1.64 (-1.641) -1.55 (-1.44) -.518 (-.514) -.458 (-.451) 6.4 5.8 5.84 5.86 5.88 5.9 (b) 6. 6.7 6.68 g g 6. 5.98 g 6.84 6.86 6.88 6.9 6.9 (c) g g 6.66 6.64.48.5.5.54.56 ωd/π (GHz) g e e e e e e FIG. 3. (left) Measured two-tone transmission spectra as functions of drive frequency ωd and probe frequency ωp. The color scheme is chosen such that the lowest point in each spectrum is red and the highest point is blue. The white dotted lines are calculated transition frequencies considering dressed states due to drive signals. Right panels show energy-level diagrams. The thin green arrows indicate transitions scanned by the probe signal, while thick magenta arrows indicate transitions scanned by drive signals.

4 mission of a probe signal with frequency ω p around the frequencies ω g, or ω ẽ,ẽ1 is measured. When ω d is equal to the frequency of an allowed transition involving at least one of these four states, the Autler-Townes splitting [7] takes place, which results in the increase of the probe transmission. Figure 3 shows the measured two-tone transmission spectra from set H. An avoided crossing between a horizontal line and a diagonal line is observed in each panel. Interestingly, the slope of the diagonal line is ω p / ω d = 1 for panels (a) and (b), and 1 for panel (c), which indicate that ω p + ω d = const. for panels (a) and (b), while ω p ω d = const. for panel (c). Together with the frequencies numerically calculated from ĤRabi, the corresponding transitions are determined as shown in the right-hand side of each spectrum. The spectrum in panel (c) demonstrates that the energy of is higher than that of ẽ1 and hence 1 is negative. In other words, the qubit s energy levels are inverted. Moreover, from these three two-tone transmission spectra, five transition frequencies, ω g,, ω, g, ω ẽ,ẽ1, ωẽ1,ẽ, and 1, can be evaluated; For panel (a), the holizontal line represents the condition of one-photon resonance, ω p = ω g,, whereas the diagonal line represents the condition of two-photon resonance, ω p = ω g, + ω, g ω d. For panel (b), similarly, the holizontal line is ω p = ωẽ,ẽ1 and the diagonal line is ω p = ωẽ,ẽ1 + ω ẽ1,ẽ ω d. For panel (c), the holizontal line is ω p = ω g, and the diagonal line is ω p = ω g, + 1+ω d. From these five transition frequencies, all the eigenenergies up to the fifth-excited state can be determined. One thing is worth emphasizing here. In the two-tone spectroscopy of a deep-strongly-coupled qubit-oscillator circuit, the states of the qubit are doubly dressed: one is a conventional dressing by a classical strong drive field while the other is in the quantum regime due to deepstrong coupling to the oscillator, where the oscillator s states are displaced. The experimental results demonstrate that the two kinds of dressed states coexist. In Fig. 4, normalized photon-number-dependent qubit frequencies n / (n =, 1, ) obtained from the twotone transmission spectra are plotted for nine sets of parameters. The x axis is the normalized coupling constant g/ω. The parameters, ω, and g are obtained from the transmission spectra. These results demonstrate huge Lamb shifts, some of them exceeds 9% of the bare qubit frequencies. These results also demonstrate that 1-photon and -photon ac-stark shifts are so large that 1 and change their signs depending on g/ω. The solid lines are theoretically predicted values given by Eq. (5). Table I shows a comparison between the measured and the numerically calculated n using ĤRabi and the parameters, ω, and g. In many circuits, the measured is smaller than the numerically calculated one, while the agreement of and 1 are good, where the deviations are at most 1 MHz. As discussed in [6], the Δ n /Δ 1..5..5..5 1. g/ω photon 1 photon photon FIG. 4. Photon-number-dependent normalized qubit frequencies n/ as functions of g/ω. The parameters, ω, and g are obtained from the (single-tone) transmission spectra. The black, red, and blue solid circles are respectively the qubit frequencies, 1, and obtained from the two-tone transmission spectra. The solid lines are n obtained from Eq. (5). numerically calculated in the range.8 g/ω 1.1 is larger than given by Eq. (5) and hence the agreement of in Fig. 4 is a coincidence. In this way, our results can be used to check how well the flux qubit-lc oscillator circuits realize a system that is described by the quantum Rabi model Hamiltonian. A possible source of the deviation in is higher energy levels of the flux qubit. As discussed in Ref. [17], the second or higher excited states can shift energy levels of the qubit-oscillator circuit, even though there is energy difference at least GHz between the first and the second excited states. Consideration of the higher energy levels are necessary to identify the origin of the deviation in. In conclusion, we have studied strongly-coupled and deep-strongly-coupled flux qubit-lc oscillator circuits. By using two-tone spectroscopy, frequencies of the six lowest energy eigenstates are measured, and photonnumber-dependent qubit frequencies were evaluated. We find Lamb shifts over 9% of the bare qubit frequency, and inversions of the qubit s ground and excited states caused by the 1-photon and -photon ac-stark shifts. The results agree with the quantum Rabi model, giving further support to the validity of the quantum Rabi model in describing these circuits in the deep-strongcoupling regime. We thank Masahiro Takeoka for stimulating discussions. This work was supported by the Scientific Research (S) Grant No. JP561 by the Japanese Society for the Promotion of Science (JSPS), and JST CREST Grant Number JPMJCR1775, Japan.

5 SUPPLEMENTAL MATERIAL SYMMETRY OF QUANTUM RABI MODEL AND STATE ASSIGNMENT FROM THE SPECTRA The parity operator of the qubit-oscillator system is defined as ˆP = ˆP q ˆP o ˆσ z ( 1)â â, where, ˆPq ˆσ z and ˆP o ( 1)â â are respectively the parity operators of the qubit and the oscillator. The parities of states and operators are defined as follows; For the eigen states, the parity is + when the eigen value is 1, and the parity is when the eigen value is 1. The parity of an operator  is + when [Â, ˆP ] =  ˆP  ˆP =, and is when {Â, ˆP } =  ˆP + ˆP =. The parity symmetry in the states and operators that appear in the quantum Rabi model Hamiltonian Ĥ Rabi = ˆσ z + ωâ â + gˆσ x (â + â ), (S1) is summarized in Table SI. Because both ˆσ x and (â + â ) have negative parity, their product has positive parity, meaning that all three terms in ĤRabi have positive parity. Therefore, [ĤRabi, ˆP ] =, and hence, the energy eigenstates are also eigenstates of ˆP. Note that this property does not depend on the values of, ω, and g. Although the energy eigenstates of Ĥ Rabi cannot be described simply for arbitrary values of, ω, and g, the symmetry allows to define energy eigenstates of ĤRabi as in, where (i = g, e) indicate that the qubit is in g the ground or e the excited state and the oscillator is in the n-photon Fock state, Since the parity of (â + â ) is, the transition matrix elements ĩm (â+â ) jn is non-zero when the parities of the energy eigenstates ĩm and jn are different, and is zero when the parities are same. From the transition frequencies alone, the energy eigenstates cannot be determined uniquely. However, by using the parity symmetry discussed above, energy eigenstates are recursively determined as long as < ω in the following way. (i) The ground and the first excited states of a coupled circuit are respectively g and ẽ, since there is no energy-level crossing between them. (ii) Between the (n)th and (n + 1)th excited states (n 1), the state having nonzero transition matrix element with the state gn is gn + 1. In this way, photon-numberdependent qubit frequency n can be uniquely defined for all the parameter sets in this work. parity + qubit state g q = q + q e q = q q photon state even number o odd number o qubit operator ˆσ z ˆσ x photon operator â â â + â TABLE SI. The parity symmetry in the states and operators that appeard in ĤRabi. where S 1 (ε, ω p ) = 1 (Q L /Q e )e iφ 1 + iq L ω p ω (ε) ω (ε), (S3) and we assumed that a background S bg 1 (ω p) is independent of energy bias ε and is written by a polynomial of the probe photon frequency ω p. Eq. (S3) can be applied to a transmission line that is inductively and capacitively coupled to an LC oscillator [8], where, Q L is the total quality factor of the resonator, Q e is the external quality factor due to the coupling to the transmission line, and φ is a phase factor that accounts for the asymmetry of the resonance lineshape. As written in the main text, S 1 (ε, ω p ) becomes larger than 1 depending on the value value of φ. NUMERICALLY CALCULATED n In Fig. S1, normalized photon-number-dependent qubit frequencies n / obtained from the two-tone spectroscopies are plotted in open stars for set E, which has largest value of /ω =.933. The solid lines are theoretically predicted values in the limit ω: n (g/ω) ω exp( g /ω )L n (4g /ω ), (S4) which is also given in the main text. The dotted lines are numerically calculated values from ĤRabi and the parameter /ω =.933. Although there are clear deviations in smaller values of g/ω, the qualitative behaviors of solid and dotted lines are similar. Interestingly, the blue open star (measured ) is on the solid line rather than the dotted line, where the latter is expected to give more accurate prediction. The numerically calculated in the range.8 g/ω 1.1 is larger than given by Eq. (S4) and hence the agreement of the blue open star and the solid line is a coincidence. BACKGROUND TRANSMISSION AMPLITUDE The amplitudes of the measured transmission spectra S meas 1 (ε, ω p ) are fitted by the following formula: S meas 1 (ε, ω p ) = S bg 1 (ω p)s 1 (ε, ω p ), (S) fumiki@nict.go.jp semba@nict.go.jp [1] W. E. Lamb and R. C. Retherford, Phys. Rev. 7, 41 (1947).

6 Δ n /Δ 1..5..5..5 1. g/ω Eq. (S4) H Rabi photon 1 photon photon FIG. S1. Photon-number-dependent normalized qubit frequencies n/ as functions of g/ω. The parameters, ω, and g are obtained from the transmission spectra. The black, red, and blue solid lines are respectively, 1, and obtained from Eq. (S4). The dotted lines are numerically calculated n from ĤRabi for /ω =.933 corresponding to set E. The open stars are the qubit frequencies obtained from two-tone spectroscopies for set E. [] A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin, and R. J. Schoelkopf, Phys. Rev. A 69, 63 (4). [3] D. F. Walls and G. J. Milburn, Quantum optics (Springer Science & Business Media, 7). [4] R. J. Thompson, G. Rempe, and H. J. Kimble, Phys. Rev. Lett. 68, 113 (199). [5] A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang, J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, Nature 431, 16 (4). [6] S. Kato and T. Aoki, Phys. Rev. Lett. 115, 9363 (15). [7] M. Brune, F. Schmidt-Kaler, A. Maali, J. Dreyer, E. Hagley, J. M. Raimond, and S. Haroche, Phys. Rev. Lett. 76, 18 (1996). [8] J. Johansson, S. Saito, T. Meno, H. Nakano, M. Ueda, K. Semba, and H. Takayanagi, Phys. Rev. Lett. 96, 176 (6). [9] D. J. Heinzen and M. S. Feld, Phys. Rev. Lett. 59, 63 (1987). [1] M. Brune, P. Nussenzveig, F. Schmidt-Kaler, F. Bernardot, A. Maali, J. M. Raimond, and S. Haroche, Phys. Rev. Lett. 7, 3339 (1994). [11] A. Fragner, M. Göppl, J. M. Fink, M. Baur, R. Bianchetti, P. J. Leek, A. Blais, and A. Wallraff, Science 13, 1357 (8). [1] D. I. Schuster, A. Wallraff, A. Blais, L. Frunzio, R.-S. Huang, J. Majer, S. M. Girvin, and R. J. Schoelkopf, Phys. Rev. Lett. 94, 136 (5). [13] D. Schuster, A. Houck, J. Schreier, A. Wallraff, J. Gambetta, A. Blais, L. Frunzio, J. Majer, B. Johnson, M. Devoret, et al., Nature 445, 515 (7). [14] 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, Nature Phys. 6, 77 (1). [15] 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. 15, 371 (1). [16] F. Yoshihara, T. Fuse, S. Ashhab, K. Kakuyanagi, S. Saito, and K. Semba, Nature Physics 13, 44 (17). [17] F. Yoshihara, T. Fuse, S. Ashhab, K. Kakuyanagi, S. Saito, and K. Semba, Phys. Rev. A 95, 5384 (17). [18] I. I. Rabi, Phys. Rev. 51, 65 (1937). [19] E. T. Jaynes and F. W. Cummings, Proceedings of the IEEE 51, 89 (1963). [] D. Braak, Phys. Rev. Lett. 17, 141 (11). [1] S. Ashhab and F. Nori, Phys. Rev. A 81, 4311 (1). [] J. Fink, M. Göppl, M. Baur, R. Bianchetti, P. Leek, A. Blais, and A. Wallraff, Nature 454, 315 (8). [3] J. E. Mooij, T. P. Orlando, L. Levitov, L. Tian, C. H. van der Wal, and S. Lloyd, Science 85, 136 (1999). [4] H. Bloch and A. Siegert, Phys. Rev. 57, 5 (194). [5] C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Atom - Photon Interactions: Basic Process and Applications (John Wiley and Sons, Inc., New York, 199), Chap. 6. [6] See Supplemental Material. [7] S. H. Autler and C. H. Townes, Phys. Rev. 1, 73 (1955). [8] M. S. Khalil, M. J. A. Stoutimore, F. C. Wellstood, and K. D. Osborn, Journal of Applied Physics 111, 5451 (1).