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

Current address: ]Pritzker School of Molecular Engineering, The University of Chicago, Chicago, Illinois 60637, USA

Quantum control and noise protection of a Floquet 0π0-\pi qubit

Zhaoyou Wang [email protected] [ E. L. Ginzton Laboratory and the Department of Applied Physics, Stanford University, Stanford, CA 94305 USA    Amir H. Safavi-Naeini [email protected] E. L. Ginzton Laboratory and the Department of Applied Physics, Stanford University, Stanford, CA 94305 USA AWS Center for Quantum Computing, Pasadena, CA, 91125, USA
Abstract

Time-periodic systems allow engineering new effective Hamiltonians from limited physical interactions. For example, the inverted position of the Kapitza pendulum emerges as a stable equilibrium with rapid drive of its pivot point. In this work, we propose the Kapitzonium: a Floquet qubit that is the superconducting circuit analog of a mechanical Kapitza pendulum. Under periodic driving, the emerging qubit states are exponentially protected against bit and phase flips caused by dissipation, which is the primary source of decoherence of current qubits. However, we find that dissipation causes leakage out of the Floquet qubit subspace. We engineer a passive cooling scheme to stabilize the qubit subspace, which is crucial for high fidelity quantum control under dissipation. Furthermore, we introduce a hardware-efficient fluorescence-based method for qubit measurement and discuss the experimental implementation of the Floquet qubit. The proposed Kapitzonium is one of the simplest Floquet qubits that can be realized with current technology – and it already has many intriguing features and capabilities. Our work provides the first steps to develop more complex Floquet quantum systems from the ground up to realize large-scale protected engineered dynamics.

I Introduction

Superconducting circuits offer flexible qubit designs, which hold promise for engineering scalable quantum computers Blais et al. (2021); Kjaergaard et al. (2020); Koch et al. (2007); Manucharyan et al. (2009); Brooks et al. (2013); Gyenis et al. (2021a); Kalashnikov et al. (2020); Pechenezhskiy et al. (2020); Ofek et al. (2016); Hu et al. (2019); Campagne-Ibarcq et al. (2020). Quantum information is encoded in eigenstates of the circuit, and qubit properties can be engineered with different circuit designs. Most superconducting qubits today are based on single-node circuits, which have demonstrated long coherence time and high gate fidelity Blais et al. (2021); Kjaergaard et al. (2020); Koch et al. (2007); Manucharyan et al. (2009). More complex circuits have been proposed, such as the 0π0-\pi qubit Brooks et al. (2013), which provide a greater level of protection to decay and dephasing. The intrinsic noise protection of these circuits comes from engineering two degenerate ground states with disjoint wavefunctions Groszkowski et al. (2018); Gyenis et al. (2021b). Many local noise processes are significantly suppressed by these types of states. However, the protected qubits usually require circuit parameters that are demanding for current experiments Groszkowski et al. (2018); Gyenis et al. (2021a, b).

Alternative and less demanding approaches to realizing circuit-level noise protection has emerged over the last decade. In these approaches, time-modulation of the superconducting circuit is used to engineer a subspace that is protected against either bit/phase flips or both. Prominent examples of these schemes include the dissipative cats Mirrahimi et al. (2014); Leghtas et al. (2015); Lescanne et al. (2020), Kerr cats Puri et al. (2017); Grimm et al. (2020), as well as recent proposals and implementations of autonomously corrected qubits Kapit (2016); Li et al. (2023). In all of these works, the modulation frequencies and amplitudes are tuned to induce a certain set of transitions. The drive amplitudes must also be limited to avoid driving higher-order or unwanted transitions.

Here we propose a qubit design based on circuits that are periodically modulated in time. The modulation in our case is highly nonperturbative and not tuned to any particular frequency. In general, the dynamics of a time-periodic system is governed by the effective Hamiltonian from the Floquet theory Eckardt (2017); Venkatraman et al. (2022). Engineering the Floquet Hamiltonian enables new designs of superconducting qubits Didier et al. (2019); Huang et al. (2021); Gandon et al. (2022). The Floquet qubit we study is the superconducting circuit analog to the mechanical Kapitza pendulum Landau and Lifshits (1976). In Kapitza pendulum, the pivot point of the pendulum is periodically moved up and down (Fig. 1(a)), leading to qualitatively new dynamics. Notably, Kapitza pendulum has two stable equilibria: one at ϕ=0\phi=0 and the other at ϕ=π\phi=\pi. In the qubit that we propose here, these correspond to the qubit basis states. We thus name this Floquet 0π0-\pi qubit the Kapitzonium.

Our paper is organized as follows. In Sec. II, we introduce Kapitzonium and its unitary gates. In Sec. III, we consider the open system dynamics of Kapitzonium where heating effects induced by charge noise are suppressed with engineered cooling. In Sec. IV, we discuss some technical aspects towards the experimental implementation of the Kapitzonium. In Sec. V, we conclude our discussion with potential directions for future study.

II Unitary dynamics of Kapitzonium

II.1 Kapitzonium Hamiltonian

The Kapitzonium circuit (Fig. 1(b)) is identical to that of a capacitively shunted superconducting quantum interference device (SQUID). The external flux Φext(t)\Phi_{\text{ext}}(t) threading the SQUID loop is used to emulate the effect of a time-modulated pivot point of a pendulum. The Kapitzonium Hamiltonian is derived with the established circuit quantization procedures Vool and Devoret (2017); You et al. (2019); Rajabzadeh et al. (2022); Chitta et al. (2022). The branch flux variable φ^k\hat{\varphi}_{k} across each Josephson junction and the conjugate charge variable n^k\hat{n}_{k} satisfy [φ^k,n^k]=i[\hat{\varphi}_{k},\hat{n}_{k}]=i for k=1,2k=1,2. The circuit Hamiltonian is

H^=k=1,24ECkn^k2EJkcosφ^k,\hat{H}=\sum_{k=1,2}4E_{Ck}\hat{n}^{2}_{k}-E_{Jk}\cos\hat{\varphi}_{k}, (1)

where EJkE_{Jk} are the Josephson energies and ECk=e2/2CkE_{Ck}=e^{2}/2C_{k} are the charging energies with CkC_{k} being the junction capacitances.

Refer to caption
Figure 1: Schematics of Kapitzonium. (a) The Kapitza pendulum. (b) Superconducting circuit implementing the Kapitzonium. (c) Effective double well potential showing the disjoint wavefunctions and degeneracy of |0\left|0\right\rangle and |π\left|\pi\right\rangle.

Flux quantization sets the constraint of φ^1φ^2=Φext(t)\hat{\varphi}_{1}-\hat{\varphi}_{2}=\Phi_{\text{ext}}(t). In the symmetric case of EJ1=EJ2E_{J1}=E_{J2} and EC1=EC2E_{C1}=E_{C2}, Eq. (1) reduces to the Kapitzonium Hamiltonian

H^(t)=4ECn^2EJcosϕext(t)cosϕ^\hat{H}(t)=4E_{C}\hat{n}^{2}-E_{J}\cos\phi_{\text{ext}}(t)\cos\hat{\phi} (2)

where ϕext(t)=Φext(t)/2\phi_{\text{ext}}(t)=\Phi_{\text{ext}}(t)/2. The flux variable ϕ^=(φ^1+φ^2)/2\hat{\phi}=(\hat{\varphi}_{1}+\hat{\varphi}_{2})/2 corresponds to the pendulum rotation angle, and n^=n^1+n^2\hat{n}=\hat{n}_{1}+\hat{n}_{2} is the conjugate charge variable satisfying [ϕ^,n^]=i[\hat{\phi},\hat{n}]=i. The charging energy is EC=e2/2(C1+C2)E_{C}=e^{2}/2(C_{1}+C_{2}) and the Josephson energy is EJ=EJ1+EJ2E_{J}=E_{J1}+E_{J2}.

To realize the analog of the Kapitza pendulum, we set the external flux to ϕext(t)=ωt\phi_{\text{ext}}(t)=\omega t. The idling dynamics are described by a Floquet Hamiltonian

H^0(t)=4ECn^2EJcosωtcosϕ^\hat{H}_{0}(t)=4E_{C}\hat{n}^{2}-E_{J}\cos\omega t\cos\hat{\phi} (3)

with a time period of T=2π/ωT=2\pi/\omega. Throughout this paper, we choose Kapitzonium parameters EJ/2π=100GHzE_{J}/2\pi=100~{}\text{GHz}, ω/2π=10GHz\omega/2\pi=10~{}\text{GHz} and EC/2π=0.01GHzE_{C}/2\pi=0.01~{}\text{GHz} unless otherwise specified. To understand the emergence of the degenerate “ground” states, we derive the effective Hamiltonian Eckardt (2017); Venkatraman et al. (2022) of H^0(t)\hat{H}_{0}(t) by integrating over the fast modulation, and find

H^eff=4ECn^2E~Jcos2ϕ^,\hat{H}_{\text{eff}}=4E_{C}\hat{n}^{2}-\tilde{E}_{J}\cos 2\hat{\phi}, (4)

where E~J=ECEJ2/ω2\tilde{E}_{J}=E_{C}E_{J}^{2}/\omega^{2}. Here we only keep terms up to the second order in the effective Hamiltonian, while the third order term is 0 and the fourth order term scales as EC3EJ2/ω4E~JE_{C}^{3}E_{J}^{2}/\omega^{4}\ll\tilde{E}_{J}, which can be neglected Eckardt (2017); Venkatraman et al. (2022).

The static effective Hamiltonian H^eff\hat{H}_{\text{eff}} possesses some of the key features of a protected qubit: disjoint wavefunctions and energy degeneracy. Since ϕ^\hat{\phi} is 2π2\pi-periodic, H^eff\hat{H}_{\text{eff}} has two near-degenerate ground states (|0±|π)/2(\left|0\right\rangle\pm\left|\pi\right\rangle)/\sqrt{2} in the deep transmon regime of E~J/EC=100\tilde{E}_{J}/E_{C}=100. Here |0\left|0\right\rangle and |π\left|\pi\right\rangle are localized at the two minima of the effective potential Veff(ϕ)=E~Jcos2ϕV_{\text{eff}}(\phi)=-\tilde{E}_{J}\cos 2\phi (Fig. 1(c)) with exponentially small overlap.

Refer to caption
Figure 2: (a) Floquet eigenstates of H^0(t)\hat{H}_{0}(t). (b) Floquet eigenenergies H^0(t)\hat{H}_{0}(t) compared with the spectrum of H^eff\hat{H}_{\text{eff}}. (c) Time evolution of |ψ0\left|\psi_{0}\right\rangle and |ψ1\left|\psi_{1}\right\rangle from 0 to TT. Here we plot the probability distribution of the ϕ\phi space wavefunction. (d) The effective potential when performing different gates.

II.2 Floquet eigenstates

The effective Hamiltonian provides a useful approximate picture of the potential and the “ground” states of the Kapitzonium – note that the states (|0±|π)/2(\left|0\right\rangle\pm\left|\pi\right\rangle)/\sqrt{2} are only the ground states of the effective Hamiltonian. To better understand the states, gates, and noise, we will need to move beyond the effective description and consider the Floquet eigenstates of the qubit. The Floquet eigenstates |Ψα(t)\left|\Psi_{\alpha}(t)\right\rangle of the idling Kapitzonium Hamiltonian H^0(t)\hat{H}_{0}(t) satisfy

H^0(t)|Ψα(t)=iddt|Ψα(t),\hat{H}_{0}(t)\left|\Psi_{\alpha}(t)\right\rangle=i\frac{d}{dt}\left|\Psi_{\alpha}(t)\right\rangle, (5)

and have the form of

|Ψα(t)=eiεαt|Φα(t).\left|\Psi_{\alpha}(t)\right\rangle=e^{-i\varepsilon_{\alpha}t}\left|\Phi_{\alpha}(t)\right\rangle. (6)

Here |Φα(t)=|Φα(t+T)\left|\Phi_{\alpha}(t)\right\rangle=\left|\Phi_{\alpha}(t+T)\right\rangle are the periodic Floquet modes and εα\varepsilon_{\alpha} are the Floquet eigenenergies.

In Fig. 2(a), we plot the even and odd superpositions of the Floquet eigenstates defined as

|ψ2k=12(|Ψ2k(t=0)+|Ψ2k+1(t=0))|ψ2k+1=12(|Ψ2k(t=0)|Ψ2k+1(t=0))\begin{split}\left|\psi_{2k}\right\rangle=&\frac{1}{\sqrt{2}}\left(\left|\Psi_{2k}(t=0)\right\rangle+\left|\Psi_{2k+1}(t=0)\right\rangle\right)\\ \left|\psi_{2k+1}\right\rangle=&\frac{1}{\sqrt{2}}\left(\left|\Psi_{2k}(t=0)\right\rangle-\left|\Psi_{2k+1}(t=0)\right\rangle\right)\end{split} (7)

for k=0,1,2k=0,1,2. The phases of |Ψα(t=0)\left|\Psi_{\alpha}(t=0)\right\rangle are chosen such that |ψ2k\left|\psi_{2k}\right\rangle (|ψ2k+1\left|\psi_{2k+1}\right\rangle) are mostly localized around ϕ=0\phi=0 (ϕ=π\phi=\pi). In Fig. 2(b), we compare the exact Floquet eigenenergies εα\varepsilon_{\alpha} with the spectrum of H^eff\hat{H}_{\text{eff}} and find a close match. Here the index α\alpha are sorted based on the overlaps between |Ψα(t=0)\left|\Psi_{\alpha}(t=0)\right\rangle and the nnth eigenstates of H^eff\hat{H}_{\text{eff}}. The disjoint wavefunctions of {|ψ2k,|ψ2k+1}\{\left|\psi_{2k}\right\rangle,\left|\psi_{2k+1}\right\rangle\} and the near degeneracy ε2kε2k+1\varepsilon_{2k}\approx\varepsilon_{2k+1} for k=0,1,2k=0,1,2 confirm the effective double well potential Veff(ϕ)V_{\text{eff}}(\phi). Finally, Floquet eigenstates are not stationary within one period. For example, the time evolution of |ψ0\left|\psi_{0}\right\rangle and |ψ1\left|\psi_{1}\right\rangle from 0 to TT shows varying ϕ\phi space wavefunctions (Fig. 2(c)).

A single qubit state (c0,c1)T(c_{0},c_{1})^{T} can be encoded in Kapitzonium as

|ψ(t)=c0|Ψ0(t)+c1|Ψ1(t),\left|\psi(t)\right\rangle=c_{0}\left|\Psi_{0}(t)\right\rangle+c_{1}\left|\Psi_{1}(t)\right\rangle, (8)

where c0c_{0} and c1c_{1} are invariant under H^0(t)\hat{H}_{0}(t). For (ε1ε0)/2π4.7(\varepsilon_{1}-\varepsilon_{0})/2\pi\approx 4.7 kHz, 1/(ε1ε0)1/(\varepsilon_{1}-\varepsilon_{0}) is much longer than the time scale of any relevant Kapitzonium operations. We thus assume ε0=ε1\varepsilon_{0}=\varepsilon_{1} and only consider the system at t=nTt=nT to simplify our discussions. In this case, the qubit basis states |Ψ0\left|\Psi_{0}\right\rangle and |Ψ1\left|\Psi_{1}\right\rangle become static, and are related to |0\left|0\right\rangle and |π\left|\pi\right\rangle by a Hadamard transformation

|Ψ0=12(|0+|π)|Ψ1=12(|0|π).\begin{split}\left|\Psi_{0}\right\rangle=&\frac{1}{\sqrt{2}}\left(\left|0\right\rangle+\left|\pi\right\rangle\right)\\ \left|\Psi_{1}\right\rangle=&\frac{1}{\sqrt{2}}\left(\left|0\right\rangle-\left|\pi\right\rangle\right).\end{split} (9)

II.3 Kapitzonium gates

The encoded qubit state can be manipulated by engineering the flux drive ϕext(t)\phi_{\text{ext}}(t). The resulting effective potentials (Fig. 2(d)) provide the intuition for the Kapitzonium gates. Here we focus on the gate Hamiltonians and details on the required flux drive are discussed in Sec. IV.

X rotation. The X gate Hamiltonian H^x(t)=H^0(t)+αxcosϕ^\hat{H}_{x}(t)=\hat{H}_{0}(t)+\alpha_{x}\cos\hat{\phi} generates the rotation along X axis. The effective potential now becomes an asymmetric double well Veff(x)(ϕ)=E~Jcos2ϕ+αxcosϕV_{\text{eff}}^{(x)}(\phi)=-\tilde{E}_{J}\cos 2\phi+\alpha_{x}\cos\phi, which lifts the degeneracy between |0\left|0\right\rangle and |π\left|\pi\right\rangle. Therefore, in the {|0,|π}\{\left|0\right\rangle,\left|\pi\right\rangle\} basis H^x(t)αxσ^z\hat{H}_{x}(t)\approx\alpha_{x}\hat{\sigma}_{z} and in the {|Ψ0,|Ψ1}\{\left|\Psi_{0}\right\rangle,\left|\Psi_{1}\right\rangle\} basis, H^x(t)αxσ^x\hat{H}_{x}(t)\approx\alpha_{x}\hat{\sigma}_{x}, leading to Rabi oscillation between |Ψ0\left|\Psi_{0}\right\rangle and |Ψ1\left|\Psi_{1}\right\rangle.

Z rotation. The depth E~J\tilde{E}_{J} of the effective double well potential can be controlled dynamically with the flux driving frequency ω\omega. Increasing ω\omega reduces E~J/EC=(EJ/ω)2\tilde{E}_{J}/E_{C}=(E_{J}/\omega)^{2} and induces stronger coupling between |0\left|0\right\rangle and |π\left|\pi\right\rangle. We choose ωz/2π=20\omega_{z}/2\pi=20 GHz for the Z gate, which lifts the degeneracy between |Ψ0\left|\Psi_{0}\right\rangle and |Ψ1\left|\Psi_{1}\right\rangle with a splitting (ε1ε0)/2π1.8(\varepsilon_{1}-\varepsilon_{0})/2\pi\approx 1.8 MHz. This implements a phase gate for |Ψ0\left|\Psi_{0}\right\rangle and |Ψ1\left|\Psi_{1}\right\rangle, i.e., the rotation along Z axis.

Two qubit XX rotation. We couple two Kapitzonium with a Josephson junction to realize the XX gate Hamiltonian H^xx(t)=H^0(t)+αxxcos(ϕ^1ϕ^2)\hat{H}_{xx}(t)=\hat{H}_{0}(t)+\alpha_{xx}\cos(\hat{\phi}_{1}-\hat{\phi}_{2}). Replacing the Josephson junction with a SQUID makes the coupling tunable. The joint effective potential is Veff(xx)(ϕ1,ϕ2)=E~Jcos2ϕ1E~Jcos2ϕ2+αxxcos(ϕ1ϕ2)V_{\text{eff}}^{(xx)}(\phi_{1},\phi_{2})=-\tilde{E}_{J}\cos 2\phi_{1}-\tilde{E}_{J}\cos 2\phi_{2}+\alpha_{xx}\cos(\phi_{1}-\phi_{2}), which lifts the degeneracy between {|00,|ππ}\{\left|00\right\rangle,\left|\pi\pi\right\rangle\} and {|0π,|π0}\{\left|0\pi\right\rangle,\left|\pi 0\right\rangle\}. In the {|0,|π}\{\left|0\right\rangle,\left|\pi\right\rangle\} basis H^xx(t)αxxσ^zσ^z\hat{H}_{xx}(t)\approx\alpha_{xx}\hat{\sigma}_{z}\otimes\hat{\sigma}_{z}, and in the {|Ψ0,|Ψ1}\{\left|\Psi_{0}\right\rangle,\left|\Psi_{1}\right\rangle\} basis H^xx(t)αxxσ^xσ^x\hat{H}_{xx}(t)\approx\alpha_{xx}\hat{\sigma}_{x}\otimes\hat{\sigma}_{x}, generating the XX rotation.

III Open system dynamics of Kapitzonium

III.1 Heating problem

In an open quantum system, a coupling between the system and bath allows the bath to induce transitions between different eigenstates of the system. The form of the coupling and the temperature of the bath both go into determining the transitions and their rates. For simplicity, we only consider a zero-temperature bath in this paper. For time-independent systems, any transition from lower to higher energy will require energy to be absorbed from the bath. Because of this, static systems coupled to zero-temperature baths eventually decay to their ground states.

Considering the level structure in Fig. 2(a), we would naively expect the zero-temperature bath to induce only the downward transition |Ψ2|Ψ0\left|\Psi_{2}\right\rangle\rightarrow\left|\Psi_{0}\right\rangle, with the opposing transition |Ψ0|Ψ2\left|\Psi_{0}\right\rangle\rightarrow\left|\Psi_{2}\right\rangle being suppressed as that would require absorption of energy from the bath. This intuition is incorrect because the Kapitzonium is a Floquet system, which is constantly exchanging energy with the Floquet drive. As a result, the bath induced transitions happen in both directions between the Floquet eigenstates: both |Ψ2|Ψ0\left|\Psi_{2}\right\rangle\rightarrow\left|\Psi_{0}\right\rangle and |Ψ0|Ψ2\left|\Psi_{0}\right\rangle\rightarrow\left|\Psi_{2}\right\rangle transitions are allowed in a Kapitzonium interacting with a zero-temperature bath. Therefore in contrast to the static 0π0-\pi qubit Brooks et al. (2013), the Floquet |0\left|0\right\rangle and |π\left|\pi\right\rangle can still decay out of the qubit subspace through what looks like a heating process.

Consider a generic Floquet system H^0(t)=H^0(t+T)\hat{H}_{0}(t)=\hat{H}_{0}(t+T) coupled to some bath degrees of freedom B^(t)\hat{B}(t) via the system operator O^\hat{O}. The system-bath Hamiltonian in the rotating frame of the bath is

H^SB(t)=H^0(t)+O^B^(t),\hat{H}_{\text{SB}}(t)=\hat{H}_{0}(t)+\hat{O}\hat{B}(t), (10)

where

B^(t)=kgk(b^keiωkt+b^keiωkt).\hat{B}(t)=\sum_{k}g_{k}\left(\hat{b}_{k}e^{-i\omega_{k}t}+\hat{b}_{k}^{\dagger}e^{i\omega_{k}t}\right). (11)

Here ωk0\omega_{k}\geq 0 is the frequency of the kkth bath mode and gkg_{k} is the coupling between the kkth mode and the system.

The emission spectrum of the Floquet system can be calculated in the interaction picture of H^0(t)\hat{H}_{0}(t), which we call the Floquet frame. More specifically, the unitary U^0(t)\hat{U}_{0}(t) generated by H^0(t)\hat{H}_{0}(t) is given by

U^0(t,t0)=α|Ψα(t)Ψα(t0)|,\hat{U}_{0}(t,t_{0})=\sum_{\alpha}\left|\Psi_{\alpha}(t)\right\rangle\left\langle\Psi_{\alpha}(t_{0})\right|, (12)

where |Ψα(t)\left|\Psi_{\alpha}(t)\right\rangle are the Floquet eigenstates of H^0(t)\hat{H}_{0}(t), and U^0(t,t0)\hat{U}_{0}(t,t_{0}) satisfies the Schrödinger equation

iddtU^0(t,t0)=H^0(t)U^0(t,t0).i\frac{d}{dt}\hat{U}_{0}(t,t_{0})=\hat{H}_{0}(t)\hat{U}_{0}(t,t_{0}). (13)

Now we could perform the unitary transformation U^0(t,t0)\hat{U}_{0}(t,t_{0}) to enter the Floquet frame, where the system-bath Hamiltonian becomes

H~SB(t)=O^(t)B^(t),\tilde{H}_{\text{SB}}(t)=\hat{O}(t)\hat{B}(t), (14)

and

O^(t)=U^0(t,t0)O^U^0(t,t0)=αβOαβ(t)|Ψα(t0)Ψβ(t0)|\hat{O}(t)=\hat{U}_{0}^{\dagger}(t,t_{0})\hat{O}\hat{U}_{0}(t,t_{0})=\sum_{\alpha\beta}O_{\alpha\beta}(t)\left|\Psi_{\alpha}(t_{0})\right\rangle\left\langle\Psi_{\beta}(t_{0})\right| (15)

with Oαβ(t)=Ψα(t)|O^|Ψβ(t)O_{\alpha\beta}(t)=\left\langle\Psi_{\alpha}(t)\middle|\hat{O}\middle|\Psi_{\beta}(t)\right\rangle. We set t0=0t_{0}=0 without loss of generality.

Since the Floquet modes |Φα(t)\left|\Phi_{\alpha}(t)\right\rangle are periodic in time, we Fourier expand Oαβ(t)O_{\alpha\beta}(t):

Oαβ(t)=ei(εαεβ)tΦα(t)|O^|Φβ(t)=ei(εαεβ)tn=Oαβneinωt,\begin{split}O_{\alpha\beta}(t)=&e^{i(\varepsilon_{\alpha}-\varepsilon_{\beta})t}\left\langle\Phi_{\alpha}(t)\middle|\hat{O}\middle|\Phi_{\beta}(t)\right\rangle\\ =&e^{i(\varepsilon_{\alpha}-\varepsilon_{\beta})t}\sum_{n=-\infty}^{\infty}O_{\alpha\beta n}e^{in\omega t},\end{split} (16)

where OαβnO_{\alpha\beta n} together with {εα}\{\varepsilon_{\alpha}\} gives the emission spectrum. Since all b^k\hat{b}_{k} modes are in vacuum, transition αβ\alpha\rightarrow\beta from |Ψα(t0)\left|\Psi_{\alpha}(t_{0})\right\rangle to |Ψβ(t0)\left|\Psi_{\beta}(t_{0})\right\rangle is only possible if εαεβ+nω>0\varepsilon_{\alpha}-\varepsilon_{\beta}+n\omega>0, with the transition rate determined by |Oαβn|2|O_{\alpha\beta n}|^{2} and the bath spectral density at frequency εαεβ+nω\varepsilon_{\alpha}-\varepsilon_{\beta}+n\omega. Furthermore, transition αβ\alpha\rightarrow\beta could emit photons at multiple frequencies, and both αβ\alpha\rightarrow\beta and βα\beta\rightarrow\alpha could occur, which is different from the relaxation of static systems.

In Fig. 3(a), we plot |Oαβn|2|O_{\alpha\beta n}|^{2} for various transitions of the Kapitzonium under charge noise with O^=n^\hat{O}=\hat{n}. The dominant heating processes 020\rightarrow 2 (red cross) and 131\rightarrow 3 (red dot) occur at near degenerate frequency around ω02/2πω13/2π9.5GHz\omega_{02}/2\pi\approx\omega_{13}/2\pi\approx 9.5~{}\text{GHz}. The reverse cooling processes 202\rightarrow 0 (blue cross) and 313\rightarrow 1 (blue dot) are also allowed at a different frequency around ω20/2πω31/2π10.5GHz\omega_{20}/2\pi\approx\omega_{31}/2\pi\approx 10.5~{}\text{GHz}.

Assuming the bath spectral density is flat, we trace out the bath degree of freedoms and derive a master equation for the Kapitzonium. In the Floquet frame, the master equation is ρ^˙=κhD[O^(t)](ρ^)\dot{\hat{\rho}}=\kappa_{h}D[\hat{O}_{-}(t)](\hat{\rho}) (see Appendix A), where D[A^](ρ^)=A^ρ^A^12{A^A^,ρ^}D[\hat{A}](\hat{\rho})=\hat{A}\hat{\rho}\hat{A}^{\dagger}-\frac{1}{2}\{\hat{A}^{\dagger}\hat{A},\hat{\rho}\}. Here κh\kappa_{h} is the heating rate due to intrinsic loss and

O^(t)=αβOαβ(t)|Ψα(t0)Ψβ(t0)|,\hat{O}_{-}(t)=\sum_{\alpha\beta}O_{\alpha\beta}^{-}(t)\left|\Psi_{\alpha}(t_{0})\right\rangle\left\langle\Psi_{\beta}(t_{0})\right|, (17)

where

Oαβ(t)=ei(εαεβ)tεαεβ+nω<0Oαβneinωt.O_{\alpha\beta}^{-}(t)=e^{i(\varepsilon_{\alpha}-\varepsilon_{\beta})t}\sum_{\varepsilon_{\alpha}-\varepsilon_{\beta}+n\omega<0}O_{\alpha\beta n}e^{in\omega t}. (18)

Without the Floquet driving, κh\kappa_{h} is approximately the amplitude damping rate of a transmon corresponding to its 1/T11/T_{1}.

Refer to caption
Figure 3: (a) Emission spectrum of Kapitzonium under charge noise. The upper (lower) mnm\rightarrow n label within each rounded rectangle corresponds to the cross (dot). (b) Schematic for Kapitzonium coupled to a bandpass filter. (c) Two different regimes of the emission spectrum, which determines whether the qubit can be protected or not. (d) Idling and gate fidelity for different T1T_{1} time 1/κh1/\kappa_{h}, with (cross and diamond) and without (dot) the filter.

III.2 Enhanced cooling with filter

The frequency dependence of the cavity emission suggests that we can enhance a specific transition rate by increasing the bath spectral density at the transition frequency Murch et al. (2012); Putterman et al. (2022). More concretely, we propose to capacitively couple the Kapitzonium to a bandpass filter around 10.5 GHz with 800 MHz bandwidth (Fig. 3(a) grey shade region). The bandpass filter enhances the cooling processes without causing extra heating, which preserves the qubit basis states {|Ψ0,|Ψ1}\{\left|\Psi_{0}\right\rangle,\left|\Psi_{1}\right\rangle\}. Furthermore, the environment cannot distinguish whether the emitted photon comes from the 202\rightarrow 0 or 313\rightarrow 1 transition since ω20ω31\omega_{20}\approx\omega_{31}. Therefore the phase coherence between |Ψ0\left|\Psi_{0}\right\rangle and |Ψ1\left|\Psi_{1}\right\rangle is also preserved by the cooling processes, enabling fully autonomous protection of the qubit subspace.

The bandpass filter can be modeled as a chain of linearly coupled harmonic ocsillators a^1,,a^N\hat{a}_{1},...,\hat{a}_{N} Putterman et al. (2022). The first filter mode a^1\hat{a}_{1} also couples capacitively to the Kapitzonium via the interaction gn^(a^1+a^1)g\hat{n}\left(\hat{a}_{1}+\hat{a}_{1}^{\dagger}\right). The full Hamiltonian is (Fig. 3(b))

H^(t)=H^0(t)+ωfk=1Na^ka^k+gn^(a^1+a^1)+Jk=1N1(a^ka^k+1+a^ka^k+1),\begin{split}\hat{H}(t)=&\hat{H}_{0}(t)+\omega_{f}\sum_{k=1}^{N}\hat{a}_{k}^{\dagger}\hat{a}_{k}+g\hat{n}\left(\hat{a}_{1}+\hat{a}_{1}^{\dagger}\right)\\ &+J\sum_{k=1}^{N-1}\left(\hat{a}_{k}\hat{a}_{k+1}^{\dagger}+\hat{a}_{k}^{\dagger}\hat{a}_{k+1}\right),\end{split} (19)

where ωf\omega_{f} is the center frequency of the filter and JJ is the coupling rate between two adjacent filter modes. The last filter mode a^N\hat{a}_{N} decays into a zero temperature bath at rate κf\kappa_{f}, which is described by the Lindblad dissipator κfD[a^N]\kappa_{f}D[\hat{a}_{N}].

The qubit subspace is autonomously protected when the engineered cooling rate is sufficiently larger than the intrinsic heating rate κh\kappa_{h} of Kapitzonium. Following Ref. Putterman et al. (2022), we choose N=3N=3, κf=2J\kappa_{f}=2J and g=κf/5g=\kappa_{f}/5, such that the filter bandwidth is 2κf2\kappa_{f} and the filter modes are only weakly excited. The cooling rate is about κc=4g2/κf=4κf/25\kappa_{c}=4g^{2}/\kappa_{f}=4\kappa_{f}/25 after adiabatically eliminating all filter modes Reiter and Sørensen (2012); Chamberland et al. (2022). For κf/2π=400\kappa_{f}/2\pi=400 MHz we have κc/2π=64\kappa_{c}/2\pi=64 MHz. In Fig. 3(d), we plot the average fidelity of Kapitzonium idling for 50 ns with (blue cross) and without (blue dot) the filter for different values of 1/κh1/\kappa_{h}. The cooling enhanced by the filter has reduced the idling infidelity by more than 2 orders of magnitude.

III.3 Gate protection

The filter that protects idling may not protect the Kapitzonium gates since the emission spectrum could be different during the gates. The filter performance is determined by two quantities of the emission spectrum. One quantity is the frequency spacing |(ω20+ω31)(ω02+ω13)|/2\mathcal{B}\equiv|(\omega_{20}+\omega_{31})-(\omega_{02}+\omega_{13})|/2 between the dominant heating and cooling transitions. \mathcal{B} limits the filter bandwidth κf\kappa_{f} and thus the maximal cooling rate κc\kappa_{c}. The other quantity is the degeneracy of the two heating (cooling) transitions measured by Δ|ω02ω13|=|ω20ω31|\Delta\equiv|\omega_{02}-\omega_{13}|=|\omega_{20}-\omega_{31}|. In the degenerate regime where Δκc\Delta\ll\kappa_{c}, such as idling with Δ/2π0.2\Delta/2\pi\approx 0.2 MHz, the qubit subspace is protected (Fig. 3(c)). In the non-degenerate regime where Δκc\Delta\gtrsim\kappa_{c}, the environment could distinguish which cooling transition the emitted photon comes from and dephase the qubit (Fig. 3(c)). Therefore the qubit basis states |Ψ0\left|\Psi_{0}\right\rangle and |Ψ1\left|\Psi_{1}\right\rangle are protected but not their coherent superpositions.

For X gate with relatively small αx/2π5.2\alpha_{x}/2\pi\approx 5.2 MHz, the emission spectrum is similar to the idling emission spectrum. Therefore the idling filter also protects the X gate, with Δ/2π0.8\Delta/2\pi\approx 0.8 MHz well within the degenerate regime.

For Z gate with ω/2π=20\omega/2\pi=20 GHz, we have /2π468\mathcal{B}/2\pi\approx 468 MHz and Δ/2π27\Delta/2\pi\approx 27 MHz. A different filter is required with center frequency at about 20 GHz. Z gate is unprotected, since the largest possible cooling rate 4/25754\mathcal{B}/25\approx 75 MHz is comparable with Δ\Delta.

The full protection of the X gate can be verified numerically, and the gate infidelity is reduced by about 2 orders of magnitude with the filter (Fig. 3(d) orange diamond and dot). The partial protection of the qubit basis states during the Z gate is verified in Appendix B. In principle, the XX gate should also be fully protected for αxx\alpha_{xx} on the order of a few MHz, since its Δ\Delta and \mathcal{B} are similar to the X gate. However, we didn’t simulate the XX gate due to the high computational cost.

Refer to caption
Figure 4: (a) Kapitzonium measurement by measuring the frequency of the emitted photon. (b) The measurement basis depends on how to lift the degeneracy between ω20\omega_{20} and ω31\omega_{31}. (c) Quantum trajectory simulation of the Kapitzonium measurement. (d) A single trajectory corresponding to the red dashed line in (c), where S(ω)S(\omega) is plotted in the rotating frame of ωf\omega_{f}. (e) Charge sensitivity of the Floquet eigenenergies and the emission spectrum.

III.4 Fluorescence-based state measurement

Kapitzonium measurements can be performed in the unprotected regime. By measuring the frequency of the emitted photon, we learn about which transition it comes from and randomly project the system to either |Ψ0\left|\Psi_{0}\right\rangle or |Ψ1\left|\Psi_{1}\right\rangle (Fig. 4(a)). We could apply the X gate but with a much larger αx\alpha_{x} to implement the X measurement (Fig. 4(b)). Larger αx\alpha_{x} lifts the degeneracy between ω20\omega_{20} and ω31\omega_{31}, and {|0,|π}\{\left|0\right\rangle,\left|\pi\right\rangle\} are the measurement basis. On the other hand, the unprotected Z gate dephases the qubit and naturally perform the Z measurement with measurement basis {|Ψ0,|Ψ1}\{\left|\Psi_{0}\right\rangle,\left|\Psi_{1}\right\rangle\} (Fig. 4(b)).

The measurement rate is proportional to the occupation of the excited states |Ψ2\left|\Psi_{2}\right\rangle and |Ψ3\left|\Psi_{3}\right\rangle. We could increase the measurement rate by capacitively driving |Ψ0|Ψ2\left|\Psi_{0}\right\rangle\leftrightarrow\left|\Psi_{2}\right\rangle and |Ψ1|Ψ3\left|\Psi_{1}\right\rangle\leftrightarrow\left|\Psi_{3}\right\rangle at the heating transition frequency (Fig. 4(a)). The charge operator in the Floquet frame is

n^(t)=|Ψ0Ψ2|(n02eiω02t+n20eiω20t+)+|Ψ1Ψ3|(n13eiω13t+n31eiω31t+)+h.c.,\begin{split}\hat{n}(t)=&\left|\Psi_{0}\right\rangle\left\langle\Psi_{2}\right|\left(n_{02}e^{i\omega_{02}t}+n_{20}e^{-i\omega_{20}t}+...\right)\\ &+\left|\Psi_{1}\right\rangle\left\langle\Psi_{3}\right|\left(n_{13}e^{i\omega_{13}t}+n_{31}e^{-i\omega_{31}t}+...\right)+\text{h.c.},\end{split} (20)

where we only include |Ψ03\left|\Psi_{0\sim 3}\right\rangle terms from Eq. (15) and the dominant Fourier coefficients n02n20n13n311n_{02}\approx n_{20}\approx n_{13}\approx n_{31}\approx 1. We apply a charge drive 2Ω(cos(ωd1t)+cos(ωd2t))n^(t)2\Omega(\cos(\omega_{d1}t)+\cos(\omega_{d2}t))\hat{n}(t) to the Kapitzonium where ωd1=ω02\omega_{d1}=\omega_{02}, ωd2=ω13\omega_{d2}=\omega_{13} and ΩΔ,κc\Omega\ll\Delta,\kappa_{c}. This leads to the coupling Ω(n02|Ψ0Ψ2|+n13|Ψ1Ψ3|+h.c.)\Omega(n_{02}\left|\Psi_{0}\right\rangle\left\langle\Psi_{2}\right|+n_{13}\left|\Psi_{1}\right\rangle\left\langle\Psi_{3}\right|+\text{h.c.}) after the rotating wave approximation (RWA). Furthermore, {|Ψ2,|Ψ3}\{\left|\Psi_{2}\right\rangle,\left|\Psi_{3}\right\rangle\} are only weakly excited since Ωκc\Omega\ll\kappa_{c}, which decay back to {|Ψ0,|Ψ1}\{\left|\Psi_{0}\right\rangle,\left|\Psi_{1}\right\rangle\} after the measurement with the charge drive turned off.

We simulate the Kapitzonium measurements with quantum trajectory methods Wiseman and Milburn (2009) (see Appendix C). Starting from an initial state of (|Ψ0+|Ψ1)/2(\left|\Psi_{0}\right\rangle+\left|\Psi_{1}\right\rangle)/\sqrt{2}, we monitor the output field from the filter with a heterodyne measurement for 2μs2~{}\mu\text{s}, and then calculate the power spectrum S(ω)S(\omega) for each measurement record S(t)S(t). Depending on whether the system is projected into |Ψ0\left|\Psi_{0}\right\rangle or |Ψ1\left|\Psi_{1}\right\rangle, S(ω)S(\omega) show a peak at ω20\omega_{20} or ω31\omega_{31} (Fig. 4(a)). We therefore define the signal as S=S31S20S=S_{31}-S_{20} where SijS_{ij} is the integrated power within a narrow frequency window around ωij\omega_{ij}. S<0S<0 or S>0S>0 represents a measurement result of 0 or 1. The measurement fidelity is about 99.4%, estimated from a total of 3000 trajectories (Fig. 4(c)). We plot a single trajectory in Fig. 4(d) showing the measurement result S(t)S(t), the occupation pip_{i} of |Ψi\left|\Psi_{i}\right\rangle during the measurement for i=0,1i=0,1, and the spectrum S(ω)S(\omega).

IV Towards experimental implementation

IV.1 Parameter disorder

Ideally, the SQUID in Fig. 1(b) should be symmetric to engineer the Kapitzonium Hamiltonian Eq. (2). However in actual experiments, there always exists some amount of disorder which requires a more general treatment of the circuit.

The circuit Hamiltonian in presence of parameter disorder is You et al. (2019); Riwar and DiVincenzo (2022)

H^(t)=4ECn^2EJ1cos(ϕ^C2CΣΦext(t))EJ2cos(ϕ^+C1CΣΦext(t)),\begin{split}\hat{H}(t)=4E_{C}\hat{n}^{2}&-E_{J1}\cos\left(\hat{\phi}-\frac{C_{2}}{C_{\Sigma}}\Phi_{\text{ext}}(t)\right)\\ &-E_{J2}\cos\left(\hat{\phi}+\frac{C_{1}}{C_{\Sigma}}\Phi_{\text{ext}}(t)\right),\end{split} (21)

where CΣ=C1+C2C_{\Sigma}=C_{1}+C_{2} is the total junction capacitance and EC=e2/2CΣE_{C}=e^{2}/2C_{\Sigma}. To symmetrize the flux allocation inside the two cosines, we move to another reference frame by performing the unitary transformation

U^=exp(in^C2C12CΣΦext(t)),\hat{U}=\exp\left(i\hat{n}\frac{C_{2}-C_{1}}{2C_{\Sigma}}\Phi_{\text{ext}}(t)\right), (22)

where the Hamiltonian becomes

H^(t)=4ECn^2EJ1cos(ϕ^12Φext(t))EJ2cos(ϕ^+12Φext(t))C2C12CΣΦ˙ext(t)n^.\begin{split}\hat{H}(t)=&4E_{C}\hat{n}^{2}-E_{J1}\cos\left(\hat{\phi}-\frac{1}{2}\Phi_{\text{ext}}(t)\right)\\ &-E_{J2}\cos\left(\hat{\phi}+\frac{1}{2}\Phi_{\text{ext}}(t)\right)-\frac{C_{2}-C_{1}}{2C_{\Sigma}}\dot{\Phi}_{\text{ext}}(t)\hat{n}.\end{split} (23)

We will drop the unwanted term Φ˙ext(t)n^\propto\dot{\Phi}_{\text{ext}}(t)\hat{n} for now, since in principle it can be compensated for with the gate voltage.

For disorder in junction energies, we define δe=(EJ1EJ2)/(EJ1+EJ2)\delta_{e}=(E_{J1}-E_{J2})/(E_{J1}+E_{J2}). With the idling flux drive ϕext(t)=ωt\phi_{\text{ext}}(t)=\omega t, the effective Hamiltonian is

H^eff=4ECn^2E~J(1δe2)cos2ϕ^.\hat{H}_{\text{eff}}=4E_{C}\hat{n}^{2}-\tilde{E}_{J}(1-\delta_{e}^{2})\cos 2\hat{\phi}. (24)

Therefore disorder in junction energies reduces the effective E~J\tilde{E}_{J}, and for small δe\delta_{e} on the order of 0.05 this reduction is likely negligible.

IV.2 Flux control

Here we discuss the flux control for implementing the Kapitzonium gates and measurements. The flux drives ϕext(t)=ωt\phi_{\text{ext}}(t)=\omega t (idling) and ϕext(t)=ωzt\phi_{\text{ext}}(t)=\omega_{z}t (Z gate and Z measurement) are not feasible experimentally, since the required bias current grows linearly with time. One natural solution is to use a triangle waveform instead

ϕ~ext(t)={u(t)0u(t)<2π4πu(t)2πu(t)<4π,\tilde{\phi}_{\text{ext}}(t)=\left\{\begin{array}[]{ll}u(t)&0\leq u(t)<2\pi\\ 4\pi-u(t)&2\pi\leq u(t)<4\pi\end{array}\right., (25)

where u(t)=ϕext(t)(mod4π)u(t)=\phi_{\text{ext}}(t)\pmod{4\pi}. However, this flux choice poses inconveniences for implementing the X gate and X measurement.

Alternatively, we could set ϕext(t)=αcosωt\phi_{\text{ext}}(t)=\alpha\cos\omega t with

cosϕext(t)=J0(α)+2n=1(1)nJ2n(α)cos2nωt,\cos\phi_{\text{ext}}(t)=J_{0}(\alpha)+2\sum_{n=1}^{\infty}(-1)^{n}J_{2n}(\alpha)\cos 2n\omega t, (26)

where we have applied the Jacobi–Anger expansion and Jn(x)J_{n}(x) is the nn-th Bessel function of the first kind. The effective Hamiltonian for Eq. (2) now becomes

H^eff=4ECn^2EJJ0(α)cosϕ^E~Jn=1(J2n(α)n)2cos2ϕ^.\begin{split}\hat{H}_{\text{eff}}=&4E_{C}\hat{n}^{2}-E_{J}J_{0}(\alpha)\cos\hat{\phi}\\ &-\tilde{E}_{J}\sum_{n=1}^{\infty}\left(\frac{J_{2n}(\alpha)}{n}\right)^{2}\cos 2\hat{\phi}.\end{split} (27)

Therefore we could choose different α\alpha such that J0(α)=0J_{0}(\alpha)=0 for Kapitzonium idling, Z gate and Z measurement, and J0(α)0J_{0}(\alpha)\neq 0 for X gate and X measurement.

IV.3 Coherence properties

We first estimate the coherence properties of the Kapitzonium, while neglecting any nonidealities due to flux, quasiparticle and offset charge noise. In this idealized model, both bit-flip rate 1/T11/T_{1} and phase-flip rate 1/T21/T_{2} of the Kapitzonium without cooling are 0, since at any time tt

Ψα(t)|n^|Ψβ(t)=0\left\langle\Psi_{\alpha}(t)\middle|\hat{n}\middle|\Psi_{\beta}(t)\right\rangle=0 (28)

for α,β{0,1}\alpha,\beta\in\{0,1\}. This is because the Kapitzonium Hamiltonian (Eq. (3)) is symmetric under the transformation of n^n^\hat{n}\rightarrow-\hat{n}, and both |Ψ0(t)\left|\Psi_{0}(t)\right\rangle and |Ψ1(t)\left|\Psi_{1}(t)\right\rangle have even parity under this transformation. More generally with a static offset charge, numerical results suggest that both bit-flip and phase-flip rates are exponentially small in EJ/ωE_{J}/\omega (see Appendix D). Due to the Floquet driving, the dominant error is the leakage from the qubit subspace, e.g., from |0(|π)\left|0\right\rangle(\left|\pi\right\rangle) to |ψ2(|ψ3)\left|\psi_{2}\right\rangle(\left|\psi_{3}\right\rangle). This leakage rate 1/Tl1/T_{l} is proportional to κh\kappa_{h}.

The filter implemented in Sec. III suppresses the leakage rate by κh/κc\kappa_{h}/\kappa_{c} so 1/Tlκh2/κc1/T_{l}\propto\kappa_{h}^{2}/\kappa_{c}. Such a cooling process induces dephasing error due to the frequency difference Δ\Delta in the emitted photons, with a rate of 1/T2κh2(Δκc)21/T_{2}\propto\frac{\kappa_{h}}{2}\left(\frac{\Delta}{\kappa_{c}}\right)^{2}. In the deep transmon regime EJωE_{J}\gg\omega, we have Δexp(π2EJ82ω)\Delta\propto\exp(-\frac{\pi^{2}E_{J}}{8\sqrt{2}\omega}) and the dephasing rate is exponentially small in EJ/ωE_{J}/\omega.

In summary, the bit and phase flip rates are exponentially suppressed, while the leakage rate is reduced by the filter factor κh/κc\kappa_{h}/\kappa_{c}. Below, we give a preliminary discussion of how other noise processes that affect superconducting circuits further reduce the coherence of the Kapitzonium.

Flux noise. The flux imposed on the loop will be invariably accompanied by additional flux noise δϕ\delta\phi, which typically has a fαf^{-\alpha} spectral density and thus a small spectral weight at high frequencies. The effects of δϕ\delta\phi depend on how we choose to drive the flux. For ϕext(t)=ωt+δϕ(mod4π)\phi_{\text{ext}}(t)=\omega t+\delta\phi\pmod{4\pi}, δϕ\delta\phi corresponds to a shift in time which does not change the Floquet eigenenergies. Since the idling, Z gate and Z measurement can be implemented in this way (Sec. IV.2), they are robust to flux noise. However, experimental constraints in implementing a flux drive, as well as the proposed approach to X gate and X measurement, make other driving schemes such as ϕext(t)=αcosωt\phi_{\text{ext}}(t)=\alpha\cos\omega t (Sec. IV.2) more attractive. In this case, flux noise leads to an imposed flux of ϕext(t)=αcosωt+δϕ\phi_{\text{ext}}(t)=\alpha\cos\omega t+\delta\phi where δϕ\delta\phi has a nontrivial effect on the effective Hamiltonian.

Flux sensitivity at arbitrary flux bias occurs in other protected qubits subject to experimental constraints. For example it is also present in the static soft 0π0-\pi qubit Groszkowski et al. (2018); Gyenis et al. (2021b). Nonetheless, as in the case of the soft 0π0-\pi qubit Gyenis et al. (2021b), it is possible to operate at flux sweet-spot where the dephasing rate is protected to first order against flux noise. For a Kapitzonium with a noisy periodic flux drive ϕext(t)=αcosωt+δϕ\phi_{\text{ext}}(t)=\alpha\cos\omega t+\delta\phi, the effective Hamiltonian is also insensitive to first order to δϕ\delta\phi around δϕ=0\delta\phi=0, and leading order terms scale as δϕ2\delta\phi^{2}. Therefore operating at the sweet spot δϕ=0\delta\phi=0 can significantly reduce errors of X gate and X measurement.

Quasiparticle noise. Quasiparticle tunneling changes the charge parity. This may induces both energy relaxation and dephasing Lutchyn et al. (2005, 2006); Koch et al. (2007); Martinis et al. (2009); Catelani et al. (2011); Serniak et al. (2018) of the Kapitzonium. Here we note that in a highly simplified model of quasiparticle dynamics, the frequency shifts in the {|0,|π}\{\left|0\right\rangle,\left|\pi\right\rangle\} basis induced by quasiparticles are proportional to 0|sinφ^k2|0(1)k1sinϕext(t)2\left\langle 0\middle|\sin\frac{\hat{\varphi}_{k}}{2}\middle|0\right\rangle\approx(-1)^{k-1}\sin\frac{\phi_{\text{ext}}(t)}{2}, where k=1,2k=1,2 is the index of the junction, and π|sinφ^k2|πcosϕext(t)2\left\langle\pi\middle|\sin\frac{\hat{\varphi}_{k}}{2}\middle|\pi\right\rangle\approx\cos\frac{\phi_{\text{ext}}(t)}{2}. Therefore for ϕext(t)=ωt\phi_{\text{ext}}(t)=\omega t, |0\left|0\right\rangle and |π\left|\pi\right\rangle on average don’t accumulate extra phase and should be robust to quasiparticle tunneling . This analysis may not hold when ϕext(t)=αcosωt\phi_{\text{ext}}(t)=\alpha\cos\omega t. A more accurate model of quasiparticle tunneling is also likely to be needed to better understand the effects of quasiparticles in the Kapitzonium. In particular, the modulation drive itself can cause photon-assisted effects Houzet et al. (2019). We leave such a detailed study to future work.

Offset charge noise. In addition to quasiparticle tunneling which flips the parity of gate charge, the Kapitzonium, like the charge qubit from which it is derived, suffers from the continuous variations in the gate charge offset parameter due to uncontrolled fluctuations in the background electric fields. This can be modeled by replacing n^\hat{n} with n^ng\hat{n}-n_{g} in the Kapitzonium Hamiltonian where ngn_{g} is the random offset charge.

The Floquet eigenenergies weakly depend on ngn_{g} during idling (Fig. 4(e), left figure). The weak dependence also holds when adding αxcosϕ^\alpha_{x}\cos\hat{\phi} to the Hamiltonian. Therefore idling, X gate and X measurement are all insensitive to charge noise. On the other hand, Z gate and Z measurement are not robust to charge noise since ε1ε0\varepsilon_{1}-\varepsilon_{0} and the emitted photon frequency ω20,ω31\omega_{20},\omega_{31} depend on ngn_{g} (Fig. 4(e), middle and right figures). For example, at ng=0.5n_{g}=0.5 we have ε1=ε0\varepsilon_{1}=\varepsilon_{0} and ω20=ω31\omega_{20}=\omega_{31}. Therefore both Z gate and Z measurement cannot be performed at ng=0.5n_{g}=0.5.

Finding ways to work around these challenges as well as other issues which we may learn about from experiments (such as photon-assisted tunneling and unforeseen effects due to the drive), will be the subject of future work.

V Discussion

We have proposed a quantum Kapitza pendulum in superconducting circuit as a Floquet 0π0-\pi qubit. We identify how single- and two-qubit gates can be implemented, and propose a cooling scheme to protect the Kapitzonium against charge noise. Remarkably, we find that this exceedingly simple Floquet superconducting circuit, a flux-modulated capacitively-shunted SQUID loop, can support a protected qubit subspace. Our work reveals some of the subtle features of Floquet qubits – we elucidate the challenges associated with noise-induced heating, and how they can be overcome using filter cavities, and even used to our advantage to realize a fluorescence-based method for qubit state measurement. Our work lays the groundwork to study new Floquet systems for quantum information processing with superconducting circuits, and outlines a path towards experiments with such devices.

Acknowledgements.
We thank Yudan Guo, Taha Rajabzadeh, Nathan Lee, Takuma Makihara, Qile Su, Jayameenakshi Venkatraman and Jeremy Boaz Kline for helpful discussions. This work was supported by the U.S. government through the Office of Naval Research (ONR) under grant No. N00014-20-1-2422 and the National Science Foundation CAREER award No. ECCS-1941826, and by Amazon Web Services Inc. A.H.S.-N. acknowledges support from the Sloan fellowship.

Appendix A Floquet master equation

In this section, we derive the Floquet master equation Grifoni and Hänggi (1998) and the Kapitzonium dissipator. The system-bath Hamiltonian H^(t)=O^(t)B^(t)\hat{H}(t)=\hat{O}(t)\otimes\hat{B}(t) in the interaction picture generates the dynamics

ddtρ^SB(t)=i[H^(t),ρ^SB(t)]=i[H^(t),ρ^SB(0)i0t[H^(τ),ρ^SB(τ)]dτ]=i[H^(t),ρ^SB(0)][H^(t),0t[H^(τ),ρ^SB(τ)]dτ].\begin{split}&\frac{\mathrm{d}}{\mathrm{d}t}\hat{\rho}_{SB}(t)=-i[\hat{H}(t),\hat{\rho}_{SB}(t)]\\ =&-i\left[\hat{H}(t),\hat{\rho}_{SB}(0)-i\int_{0}^{t}[\hat{H}(\tau),\hat{\rho}_{SB}(\tau)]\text{d}\tau\right]\\ =&-i[\hat{H}(t),\hat{\rho}_{SB}(0)]-\left[\hat{H}(t),\int_{0}^{t}[\hat{H}(\tau),\hat{\rho}_{SB}(\tau)]\text{d}\tau\right].\end{split} (29)

Now we make the Born approximation ρ^SB(t)=ρ^(t)ρ^B\hat{\rho}_{SB}(t)=\hat{\rho}(t)\otimes\hat{\rho}_{B}, where ρ^(t)\hat{\rho}(t) is the system density matrix and ρ^B\hat{\rho}_{B} is the stationary bath density matrix. In addition, we also make the standard assumption that Tr[B^(t)ρ^B]=0\text{Tr}[\hat{B}(t)\hat{\rho}_{B}]=0. The system dynamics becomes

ddtρ^(t)=TrB[H^(t),0t[H^(tτ),ρ^(tτ)ρ^B]dτ]=0t(O^(tτ)ρ^(tτ)O^(t)O^(t)O^(tτ)ρ^(tτ))CB(t,tτ)dτ+h.c.,\begin{split}&\frac{\mathrm{d}}{\mathrm{d}t}\hat{\rho}(t)=-\text{Tr}_{B}\left[\hat{H}(t),\int_{0}^{t}[\hat{H}(t-\tau),\hat{\rho}(t-\tau)\otimes\hat{\rho}_{B}]\text{d}\tau\right]\\ =&\int_{0}^{t}\left(\hat{O}(t-\tau)\hat{\rho}(t-\tau)\hat{O}(t)-\hat{O}(t)\hat{O}(t-\tau)\hat{\rho}(t-\tau)\right)\\ &\qquad C_{B}(t,t-\tau)\text{d}\tau+\text{h.c.},\end{split} (30)

where we define the two-point correlation functions

CB(t,tτ)=B^(t)B^(tτ)=TrB[B^(t)B^(tτ)ρ^B]=CB(tτ,t).\begin{split}&C_{B}(t,t-\tau)=\left\langle\hat{B}(t)\hat{B}(t-\tau)\right\rangle\\ =&\text{Tr}_{B}[\hat{B}(t)\hat{B}(t-\tau)\hat{\rho}_{B}]=C_{B}(t-\tau,t)^{*}.\end{split} (31)

To proceed, we assume that the bath is stationary with a very short correlation decay time. In other words

CB(t,tτ)=CB(τ,0)CB(τ)eτ/τBC_{B}(t,t-\tau)=C_{B}(\tau,0)\equiv C_{B}(\tau)\sim e^{-\tau/\tau_{B}} (32)

where τB\tau_{B} is much shorter than any time scale we are interested in. Assuming weak system-bath coupling, we have |O^(t)O^(tτB)||O^||\hat{O}(t)-\hat{O}(t-\tau_{B})|\sim|\hat{O}| and |ρ^(t)ρ^(tτB)||O^|2|\hat{\rho}(t)-\hat{\rho}(t-\tau_{B})|\sim|\hat{O}|^{2} which is second order in the coupling strength. Therefore we could make the Markov approximation and replace ρ^(tτ)\hat{\rho}(t-\tau) with ρ^(t)\hat{\rho}(t) in Eq. (30). Since only τ0\tau\approx 0 contribute significantly to the integration, the upper limit of the integration can be extended to \infty:

ddtρ^(t)=0(O^(tτ)ρ^(t)O^(t)O^(t)O^(tτ)ρ^(t))CB(τ)dτ+h.c..\begin{split}\frac{\mathrm{d}}{\mathrm{d}t}\hat{\rho}(t)=&\int_{0}^{\infty}\left(\hat{O}(t-\tau)\hat{\rho}(t)\hat{O}(t)-\hat{O}(t)\hat{O}(t-\tau)\hat{\rho}(t)\right)\\ &\qquad C_{B}(\tau)\text{d}\tau+\text{h.c.}.\end{split} (33)

For Floquet systems, we have O^(t)=ωO^(ω)eiωt\hat{O}(t)=\sum_{\omega}\hat{O}(\omega)e^{-i\omega t} with O^(ω)=O^(ω)\hat{O}(\omega)=\hat{O}^{\dagger}(-\omega) since O^(t)\hat{O}(t) is Hermitian. Therefore

0O^(tτ)CB(τ)dτ=ωO^(ω)eiωtΓ(ω),\int_{0}^{\infty}\hat{O}(t-\tau)C_{B}(\tau)\text{d}\tau=\sum_{\omega}\hat{O}(\omega)e^{-i\omega t}\Gamma(\omega), (34)

where

Γ(ω)=0eiωτCB(τ)dτ.\Gamma(\omega)=\int_{0}^{\infty}e^{i\omega\tau}C_{B}(\tau)\text{d}\tau. (35)

We could decompose Γ(ω)\Gamma(\omega) into its real and imaginary parts as Γ(ω)=12γ(ω)+iS(ω)\Gamma(\omega)=\frac{1}{2}\gamma(\omega)+iS(\omega), where

γ(ω)=eiωτCB(τ)dτS(ω)=dω2πγ(ω)𝒫(1ωω).\begin{split}\gamma(\omega)=&\int_{-\infty}^{\infty}e^{i\omega\tau}C_{B}(\tau)\text{d}\tau\\ S(\omega)=&\int_{-\infty}^{\infty}\frac{\text{d}\omega^{\prime}}{2\pi}\gamma(\omega^{\prime})\mathcal{P}\left(\frac{1}{\omega-\omega^{\prime}}\right).\end{split} (36)

Physically the real part γ(ω)\gamma(\omega) represents the decay rate while the imaginary part S(ω)S(\omega) can be absorbed into the system Hamiltonian which we will ignore for now.

Consider a zero temperature bath with a flat spectral density function

γ(ω)={γω>00ω0\gamma(\omega)=\left\{\begin{array}[]{ll}\gamma&\quad\omega>0\\ 0&\quad\omega\leq 0\end{array}\right. (37)

as an example. We could decompose O^(t)\hat{O}(t) into positive and negative frequency parts where

O^+(t)=ω<0O^(ω)eiωtO^(t)=ω>0O^(ω)eiωt.\hat{O}_{+}(t)=\sum_{\omega<0}\hat{O}(\omega)e^{-i\omega t}\qquad\hat{O}_{-}(t)=\sum_{\omega>0}\hat{O}(\omega)e^{-i\omega t}. (38)

and O^+(t)=O^(t)\hat{O}_{+}(t)=\hat{O}^{\dagger}_{-}(t). Therefore

0O^(tτ)CB(τ)dτ=γ2ω>0O^(ω)eiωt=γ2O^(t),\int_{0}^{\infty}\hat{O}(t-\tau)C_{B}(\tau)\text{d}\tau=\frac{\gamma}{2}\sum_{\omega>0}\hat{O}(\omega)e^{-i\omega t}=\frac{\gamma}{2}\hat{O}_{-}(t), (39)

and Eq. (33) gives the Floquet master equation

ddtρ^(t)=γ2O^(t)ρ^(t)O^(t)γ2O^(t)O^(t)ρ^(t)+h.c.=γD[O^(t)]ρ^(t),\begin{split}\frac{\mathrm{d}}{\mathrm{d}t}\hat{\rho}(t)=&\frac{\gamma}{2}\hat{O}_{-}(t)\hat{\rho}(t)\hat{O}(t)-\frac{\gamma}{2}\hat{O}(t)\hat{O}_{-}(t)\hat{\rho}(t)+\text{h.c.}\\ =&\gamma D[\hat{O}_{-}(t)]\hat{\rho}(t),\end{split} (40)

where we apply RWA to drop O^±2(t)\hat{O}^{2}_{\pm}(t) terms. This justifies the Kapitzonium dissipator Eq. (17) in the main text.

Appendix B Gate simulation

B.1 Average gate fidelity

We benchmark the Kapitzonium gates with the measure of average fidelity F¯\bar{F} Nielsen (2002); Wang et al. (2022). Intuitively, F¯\bar{F} describes how well the qubit subspace is preserved under some quantum process. More specifically, the average fidelity of a quantum channel \mathcal{M} over all states |ψ=n=1Ncn|n\left|\psi\right\rangle=\sum_{n=1}^{N}c_{n}\left|n\right\rangle in a NN-dimensional qubit subspace is

F¯=𝑑ψTr[(|ψψ|)|ψψ|]=mnTr[(|mn|)ρmn],\begin{split}\bar{F}=&\int d\psi\text{Tr}\left[\mathcal{M}(\left|\psi\right\rangle\left\langle\psi\right|)\left|\psi\right\rangle\left\langle\psi\right|\right]\\ =&\sum_{mn}\text{Tr}\left[\mathcal{M}(\left|m\right\rangle\left\langle n\right|)\rho_{mn}\right],\end{split} (41)

where

ρmn=𝑑ψcmcn|ψψ|=kl|kl|𝑑ψcmcnckcl=δmn[2N(N+1)|nn|+kn1N(N+1)|kk|]+(1δmn)1N(N+1)|nm|.\begin{split}\rho_{mn}=&\int d\psi c_{m}c_{n}^{*}\left|\psi\right\rangle\left\langle\psi\right|=\sum_{kl}\left|k\right\rangle\left\langle l\right|\int d\psi c_{m}c_{n}^{*}c_{k}c_{l}^{*}\\ =&\delta_{mn}\left[\frac{2}{N(N+1)}\left|n\right\rangle\left\langle n\right|+\sum_{k\neq n}\frac{1}{N(N+1)}\left|k\right\rangle\left\langle k\right|\right]\\ &+(1-\delta_{mn})\frac{1}{N(N+1)}\left|n\right\rangle\left\langle m\right|.\end{split} (42)

Therefore the average fidelity can be simplified as

F¯=1N(N+1)nTr[(|nn|)k(1+δnk)|kk|]+2N(N+1)m<nRe{Tr[(|mn|)|nm|]}.\begin{split}\bar{F}=&\frac{1}{N(N+1)}\sum_{n}\text{Tr}\left[\mathcal{M}(\left|n\right\rangle\left\langle n\right|)\sum_{k}(1+\delta_{nk})\left|k\right\rangle\left\langle k\right|\right]\\ &+\frac{2}{N(N+1)}\sum_{m<n}\text{Re}\left\{\text{Tr}\left[\mathcal{M}(\left|m\right\rangle\left\langle n\right|)\left|n\right\rangle\left\langle m\right|\right]\right\}.\end{split} (43)

To compare \mathcal{M} with some target unitary U^\hat{U}, we could compare \mathcal{M}^{\prime} with identity instead, where (ρ^)U^(ρ^)U^\mathcal{M}^{\prime}(\hat{\rho})\equiv\hat{U}^{\dagger}\mathcal{M}(\hat{\rho})\hat{U}.

B.2 Unitary case

The gates are designed such that the system Hamiltonian adiabatically evolve from the idling H^0(t)\hat{H}_{0}(t) to the gate Hamiltonian H^gate(t)\hat{H}_{\text{gate}}(t), and then adiabatically evolve back to H^0(t)\hat{H}_{0}(t) after certain amount of gate time. The average gate fidelity is calculated for the single-qubit subspace with N=2N=2 or the two-qubit subspace with N=4N=4.

We choose the pulse shape

α(t,tgate,τ)={sin(πt2τ)20t<τ1τttgateτsin(π(tgatet)2τ)2tgateτ<ttgate,\alpha(t,t_{\text{gate}},\tau)=\left\{\begin{array}[]{ll}\sin\left(\frac{\pi t}{2\tau}\right)^{2}&0\leq t<\tau\\ 1&\tau\leq t\leq t_{\text{gate}}-\tau\\ \sin\left(\frac{\pi(t_{\text{gate}}-t)}{2\tau}\right)^{2}&t_{\text{gate}}-\tau<t\leq t_{\text{gate}}\end{array}\right., (44)

where tgatet_{\text{gate}} is the total gate duration and τ\tau is the adiabatic ramping time.

X gate. The Hamiltonian is

H^x(t)=H^0(t)+αxα(t,tx,τx)cosϕ^,\hat{H}_{x}(t)=\hat{H}_{0}(t)+\alpha_{x}\alpha(t,t_{x},\tau_{x})\cos\hat{\phi}, (45)

where tx=60t_{x}=60~{}ns, τx=10\tau_{x}=10~{}ns and αx/2π5.2\alpha_{x}/2\pi\approx 5.2~{}MHz. The X gate implements the mapping of |Ψ0|Ψ1,|Ψ1|Ψ0\left|\Psi_{0}\right\rangle\rightarrow\left|\Psi_{1}\right\rangle,\left|\Psi_{1}\right\rangle\rightarrow\left|\Psi_{0}\right\rangle with an infidelity of 1.7×1071.7\times 10^{-7}.

Z gate. The Hamiltonian is

H^z(t)=4ECn^2EJαz(t)cosϕ^,\hat{H}_{z}(t)=4E_{C}\hat{n}^{2}-E_{J}\alpha_{z}(t)\cos\hat{\phi}, (46)

where

αz(t)=(1α(t,tz,τz))cosωt+α(t,tz,τz)cosωzt.\alpha_{z}(t)=(1-\alpha(t,t_{z},\tau_{z}))\cos\omega t+\alpha(t,t_{z},\tau_{z})\cos\omega_{z}t. (47)

We choose tz=296.2t_{z}=296.2~{}ns, τz=20\tau_{z}=20~{}ns and ωz/2π=20\omega_{z}/2\pi=20 GHz. The Z gate implements the mapping of |Ψ0|Ψ0,|Ψ1|Ψ1\left|\Psi_{0}\right\rangle\rightarrow\left|\Psi_{0}\right\rangle,\left|\Psi_{1}\right\rangle\rightarrow-\left|\Psi_{1}\right\rangle with an infidelity of 4.6×1064.6\times 10^{-6}.

Another Floquet drive that seems reasonable at first is to have frequency modulation instead of amplitude modulation:

αz(t)=cos[((1α(t,tz,τz))ω+α(t,tz,τz)ωz)t].\alpha_{z}(t)=\cos[((1-\alpha(t,t_{z},\tau_{z}))\omega+\alpha(t,t_{z},\tau_{z})\omega_{z})t]. (48)

However, this Floquet drive always leads to unstable dynamics which heats up the Kapitzonium even with very slow ramping.

XX gate. The Hamiltonian is

H^xx(t)=H^0(1)(t)+H^0(2)(t)+αxxα(t,txx,τxx)cos(ϕ^1ϕ^2),\hat{H}_{xx}(t)=\hat{H}_{0}^{(1)}(t)+\hat{H}_{0}^{(2)}(t)+\alpha_{xx}\alpha(t,t_{xx},\tau_{xx})\cos(\hat{\phi}_{1}-\hat{\phi}_{2}), (49)

where H^0(i)(t)=4ECn^i2EJcosωtcosϕ^i\hat{H}_{0}^{(i)}(t)=4E_{C}\hat{n}_{i}^{2}-E_{J}\cos\omega t\cos\hat{\phi}_{i} is the idling Hamiltonian for each qubit with i=1,2i=1,2. We choose txx=39t_{xx}=39~{}ns, τxx=12\tau_{xx}=12~{}ns and αxx/2π=10\alpha_{xx}/2\pi=10 MHz. The XX gate implements the mapping of |Ψ0Ψ0|Ψ1Ψ1,|Ψ0Ψ1|Ψ1Ψ0,|Ψ1Ψ0|Ψ0Ψ1,|Ψ1Ψ1|Ψ0Ψ0\left|\Psi_{0}\Psi_{0}\right\rangle\rightarrow\left|\Psi_{1}\Psi_{1}\right\rangle,\left|\Psi_{0}\Psi_{1}\right\rangle\rightarrow\left|\Psi_{1}\Psi_{0}\right\rangle,\left|\Psi_{1}\Psi_{0}\right\rangle\rightarrow\left|\Psi_{0}\Psi_{1}\right\rangle,\left|\Psi_{1}\Psi_{1}\right\rangle\rightarrow\left|\Psi_{0}\Psi_{0}\right\rangle with an infidelity of 2.4×1072.4\times 10^{-7}.

State initialization. To initialize the system state into |0\left|0\right\rangle, we could start from the ground state of the static transmon Hamiltonian H^=4ECn^2EJcosϕ^\hat{H}=4E_{C}\hat{n}^{2}-E_{J}\cos\hat{\phi}, and adiabatically apply the Floquet drive:

H^(t)=4ECn^2EJ(α(t)cosωt+(1α(t)))cosϕ^,\hat{H}(t)=4E_{C}\hat{n}^{2}-E_{J}\left(\alpha(t)\cos\omega t+(1-\alpha(t))\right)\cos\hat{\phi}, (50)

where α(t)\alpha(t) increases from 0 to 1.

B.3 Open system without filter

In the open system simulation without filter, the Floquet drives are the identical to the unitary case. The only difference is that during t[τ,tgateτ]t\in[\tau,t_{\text{gate}}-\tau] we add loss to the system. More specifically, the simulation is performed in the Floquet frame with Hamiltonian 0 and a single Lindblad dissipator κhD[O^(t)]\kappa_{h}D[\hat{O}_{-}(t)]. Notice that the Floquet frame here is defined with respect to H^gate(t)\hat{H}_{\text{gate}}(t) instead H^0(t)\hat{H}_{0}(t).

During the ramping parts of Floquet drive, the Hamiltonian is not strictly time-periodic which makes it difficult to calculate the time-dependent dissipator. Therefore the ramping parts are always assumed to be unitary where the simulation is done in the lab frame.

B.4 Open system with filter

Due to the hybridization between the Kapitzonium and the filter, |Ψα(t)|0f\left|\Psi_{\alpha}(t)\right\rangle\otimes\left|0_{f}\right\rangle is no longer the Floquet eigenstates of full Hamiltonian (Eq. (19) in the main text) where |0f\left|0_{f}\right\rangle is the ground state of the filter. Therefore we work with the dressed Floquet eigenstates |Ψα(t)~\widetilde{\left|\Psi_{\alpha}(t)\right\rangle} of Eq. (19) instead. The qubit basis states {|Ψ0(t)~,|Ψ1(t)~}\{\widetilde{\left|\Psi_{0}(t)\right\rangle},\widetilde{\left|\Psi_{1}(t)\right\rangle}\} are chosen based on their overlap with {|Ψ0(t)|0f,|Ψ1(t)|0f}\{\left|\Psi_{0}(t)\right\rangle\otimes\left|0_{f}\right\rangle,\left|\Psi_{1}(t)\right\rangle\otimes\left|0_{f}\right\rangle\}.

The Floquet drive parameters requires a slight fine tuning due to this hybridization. During the gate time t[τ,tgateτ]t\in[\tau,t_{\text{gate}}-\tau], the simulation is performed in the Floquet frame defined by H^gate(t)\hat{H}_{\text{gate}}(t) with Hamiltonian 0 and two Lindblad dissipators κhD[O^(t)]\kappa_{h}D[\hat{O}_{-}(t)] and κfD[a^N(t)]\kappa_{f}D[\hat{a}_{N}(t)]. Here a^N(t)\hat{a}_{N}(t) is calculated similarly to Eq. (17) with O^=a^N+a^N\hat{O}=\hat{a}_{N}+\hat{a}_{N}^{\dagger}. To remove any transient effects at the beginning of the cooling Kapit (2018), we prepare the initial states for benchmarking the gates by evolving the qubit basis states for 50 ns idling until the system reaches equilibrium.

In Fig. 5(a), we simulate the unprotected Z gate for different initial states. Here |Ψ0\left|\Psi_{0}\right\rangle and |Ψ1\left|\Psi_{1}\right\rangle are the dressed Floquet eigenstates with ω/2π=20\omega/2\pi=20 GHz and |±\left|\pm\right\rangle are their even and odd superpositions. We choose ωf/2π20.234\omega_{f}/2\pi\approx 20.234~{}GHz, κf/2π=200\kappa_{f}/2\pi=200~{}MHz J=κf/2J=\kappa_{f}/2 and g=κf/5g=\kappa_{f}/5 for the Z gate filter.

We use QuTiP Johansson et al. (2012, 2013) for all the simulations, and modify the built-in mesolve function to speed up the open system simulation with time dependent dissipators.

Refer to caption
Figure 5: (a) Z gate simulation with filter for different initial states. (b) Z measurement results.

Appendix C Measurement simulation

The Hamiltonian for the measurement simulation is

H^(t)=2Ω(cos(ωd1t)+cos(ωd2t))n^(t)+gn^(t)(a^1eiωft+a^1eiωft)+Jk=1N1(a^ka^k+1+a^ka^k+1),\begin{split}\hat{H}(t)=&2\Omega(\cos(\omega_{d1}t)+\cos(\omega_{d2}t))\hat{n}(t)\\ &+g\hat{n}(t)(\hat{a}_{1}e^{-i\omega_{f}t}+\hat{a}_{1}^{\dagger}e^{i\omega_{f}t})\\ &+J\sum_{k=1}^{N-1}\left(\hat{a}_{k}\hat{a}_{k+1}^{\dagger}+\hat{a}_{k}^{\dagger}\hat{a}_{k+1}\right),\end{split} (51)

which is in the Floquet frame of the Kapitzonium and the rotating frame of the filter modes. Notice that the simulation only includes the first 4 Floquet eigenstates with n^(t)\hat{n}(t) in Eq. (20) represented by a 4×44\times 4 matrix.

For X measurement, the system Hamiltonian is H^x=H^0(t)+αxcosϕ^\hat{H}_{x}=\hat{H}_{0}(t)+\alpha_{x}\cos\hat{\phi} with αx/2π=500\alpha_{x}/2\pi=500 MHz. n^(t)\hat{n}(t) can be calculated from the emission spectrum of H^x\hat{H}_{x}. The charge drive frequencies are ω02/2π9.44\omega_{02}/2\pi\approx 9.44~{}GHz and ω13/2π9.52\omega_{13}/2\pi\approx 9.52~{}GHz. The emitted photon frequencies are ω20/2π10.56\omega_{20}/2\pi\approx 10.56~{}GHz and ω31/2π10.48\omega_{31}/2\pi\approx 10.48~{}GHz. We choose Ω/2π=6.4\Omega/2\pi=6.4~{}MHz and the filter parameters g,J,ωfg,J,\omega_{f} are the same as the idling filter.

For Z measurement, the system Hamiltonian is the same as the Z gate with an additional charge drive. The charge drive frequencies are ω02/2π19.78\omega_{02}/2\pi\approx 19.78~{}GHz and ω13/2π19.75\omega_{13}/2\pi\approx 19.75~{}GHz. The emitted photon frequencies are ω20/2π20.22\omega_{20}/2\pi\approx 20.22~{}GHz and ω31/2π20.25\omega_{31}/2\pi\approx 20.25~{}GHz. We choose Ω/2π=2.3\Omega/2\pi=2.3~{}MHz and the filter parameters are the same as the Z gate filter. Starting from the initial state (|Ψ0+|Ψ1)/2(\left|\Psi_{0}\right\rangle+\left|\Psi_{1}\right\rangle)/\sqrt{2}, we simulate 1000 trajectories for the Z measurement with a measurement time of 10μ10~{}\mus (Fig. 5(b)) and the measurement fidelity is about 99.8%.

We would like to make a few comments on Kapitzonium measurement. First of all, the charge drive could be slightly off-resonant from the heating transitions with ωd1ω02\omega_{d1}\approx\omega_{02}, ωd2ω13\omega_{d2}\approx\omega_{13}, which shifts the emitted photon frequencies as well. Second of all, the charge drive frequency should be outside the filter passband. In principle setting ωd1ω20\omega_{d1}\approx\omega_{20}, ωd2ω13\omega_{d2}\approx\omega_{13} also drives the Rabi oscillation via the cooling transitions. However, this could cause measurement error if there is any direct leakage from the charge drive to the output of the filter. Finally, choosing Ω\Omega to be comparable or larger than Δ\Delta will cause measurement error due to crosstalk between the drives. On the other hand, a very small Ω\Omega reduces the measurement rate and requires a long measurement time.

Refer to caption
Figure 6: Kapitzonium transition rates for ng=0.3n_{g}=0.3. The junction energies are EJ/2π=60,80,100,120,140,160E_{J}/2\pi=60,80,100,120,140,160 GHz, with ω/2π=10GHz\omega/2\pi=10~{}\text{GHz} and EC/2π=0.01GHzE_{C}/2\pi=0.01~{}\text{GHz} fixed.

Appendix D Kapitzonium lifetime estimation

The Kapitzonium Hamiltonian in presence of offset charge ngn_{g} is

H^0(t)=4EC(n^ng)2EJcosωtcosϕ^.\hat{H}_{0}(t)=4E_{C}(\hat{n}-n_{g})^{2}-E_{J}\cos\omega t\cos\hat{\phi}. (52)

For ng=0n_{g}=0, we have Π^H^0(t)=H^0(t)Π^\hat{\Pi}\hat{H}_{0}(t)=\hat{H}_{0}(t)\hat{\Pi} where the parity operator is Π^=n|nn|\hat{\Pi}=\sum_{n}\left|-n\right\rangle\left\langle n\right| and {|n}\{\left|n\right\rangle\} are the charge eigenstates. The Floquet eigenstates |Ψα(t)\left|\Psi_{\alpha}(t)\right\rangle are parity eigenstates, and it turns out that both |Ψ0(t)\left|\Psi_{0}(t)\right\rangle and |Ψ1(t)\left|\Psi_{1}(t)\right\rangle have even parity. Therefore

Ψα(t)|n^|Ψβ(t)=Ψα(t)|Π^n^Π^|Ψβ(t)=Ψα(t)|n^|Ψβ(t)=0,\begin{split}&\left\langle\Psi_{\alpha}(t)\middle|\hat{n}\middle|\Psi_{\beta}(t)\right\rangle=\left\langle\Psi_{\alpha}(t)\middle|\hat{\Pi}\hat{n}\hat{\Pi}\middle|\Psi_{\beta}(t)\right\rangle\\ =&-\left\langle\Psi_{\alpha}(t)\middle|\hat{n}\middle|\Psi_{\beta}(t)\right\rangle=0,\end{split} (53)

for α,β{0,1}\alpha,\beta\in\{0,1\}.

For ng0n_{g}\neq 0, H^0(t)\hat{H}_{0}(t) don’t have the symmetry under Π^\hat{\Pi} and the bit-flip and phase-flip rates are no longer exactly 0. However, since Kapitzonium is in the deep transmon regime, we still expect the differences from the ng=0n_{g}=0 results to be exponentially small in EJ/ωE_{J}/\omega. We define the total transition rate from |Ψα\left|\Psi_{\alpha}\right\rangle to |Ψβ\left|\Psi_{\beta}\right\rangle as

Γαβεαεβ+nω>0|Oαβn|2,\Gamma_{\alpha\rightarrow\beta}\equiv\sum_{\varepsilon_{\alpha}-\varepsilon_{\beta}+n\omega>0}|O_{\alpha\beta n}|^{2}, (54)

where 1/T1Γ101/T_{1}\sim\Gamma_{1\rightarrow 0} and 1/T2Γ00,Γ111/T_{2}\sim\Gamma_{0\rightarrow 0},\Gamma_{1\rightarrow 1}. In Fig. 6, we plot the transition rates for different values of EJ/ωE_{J}/\omega. The results indeed show an exponential suppression of both the bit-flip and phase flip rates with EJ/ωE_{J}/\omega.

References