Current address: ]Pritzker School of Molecular Engineering, The University of Chicago, Chicago, Illinois 60637, USA
Quantum control and noise protection of a Floquet qubit
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 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 and the other at . In the qubit that we propose here, these correspond to the qubit basis states. We thus name this Floquet 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 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 across each Josephson junction and the conjugate charge variable satisfy for . The circuit Hamiltonian is
(1) |
where are the Josephson energies and are the charging energies with being the junction capacitances.

Flux quantization sets the constraint of . In the symmetric case of and , Eq. (1) reduces to the Kapitzonium Hamiltonian
(2) |
where . The flux variable corresponds to the pendulum rotation angle, and is the conjugate charge variable satisfying . The charging energy is and the Josephson energy is .
To realize the analog of the Kapitza pendulum, we set the external flux to . The idling dynamics are described by a Floquet Hamiltonian
(3) |
with a time period of . Throughout this paper, we choose Kapitzonium parameters , and unless otherwise specified. To understand the emergence of the degenerate “ground” states, we derive the effective Hamiltonian Eckardt (2017); Venkatraman et al. (2022) of by integrating over the fast modulation, and find
(4) |
where . 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 , which can be neglected Eckardt (2017); Venkatraman et al. (2022).
The static effective Hamiltonian possesses some of the key features of a protected qubit: disjoint wavefunctions and energy degeneracy. Since is -periodic, has two near-degenerate ground states in the deep transmon regime of . Here and are localized at the two minima of the effective potential (Fig. 1(c)) with exponentially small overlap.

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 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 of the idling Kapitzonium Hamiltonian satisfy
(5) |
and have the form of
(6) |
Here are the periodic Floquet modes and are the Floquet eigenenergies.
In Fig. 2(a), we plot the even and odd superpositions of the Floquet eigenstates defined as
(7) |
for . The phases of are chosen such that () are mostly localized around (). In Fig. 2(b), we compare the exact Floquet eigenenergies with the spectrum of and find a close match. Here the index are sorted based on the overlaps between and the th eigenstates of . The disjoint wavefunctions of and the near degeneracy for confirm the effective double well potential . Finally, Floquet eigenstates are not stationary within one period. For example, the time evolution of and from to shows varying space wavefunctions (Fig. 2(c)).
A single qubit state can be encoded in Kapitzonium as
(8) |
where and are invariant under . For kHz, is much longer than the time scale of any relevant Kapitzonium operations. We thus assume and only consider the system at to simplify our discussions. In this case, the qubit basis states and become static, and are related to and by a Hadamard transformation
(9) |
II.3 Kapitzonium gates
The encoded qubit state can be manipulated by engineering the flux drive . 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 generates the rotation along X axis. The effective potential now becomes an asymmetric double well , which lifts the degeneracy between and . Therefore, in the basis and in the basis, , leading to Rabi oscillation between and .
Z rotation. The depth of the effective double well potential can be controlled dynamically with the flux driving frequency . Increasing reduces and induces stronger coupling between and . We choose GHz for the Z gate, which lifts the degeneracy between and with a splitting MHz. This implements a phase gate for and , 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 . Replacing the Josephson junction with a SQUID makes the coupling tunable. The joint effective potential is , which lifts the degeneracy between and . In the basis , and in the basis , 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 , with the opposing transition 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 and transitions are allowed in a Kapitzonium interacting with a zero-temperature bath. Therefore in contrast to the static qubit Brooks et al. (2013), the Floquet and can still decay out of the qubit subspace through what looks like a heating process.
Consider a generic Floquet system coupled to some bath degrees of freedom via the system operator . The system-bath Hamiltonian in the rotating frame of the bath is
(10) |
where
(11) |
Here is the frequency of the th bath mode and is the coupling between the th mode and the system.
The emission spectrum of the Floquet system can be calculated in the interaction picture of , which we call the Floquet frame. More specifically, the unitary generated by is given by
(12) |
where are the Floquet eigenstates of , and satisfies the Schrödinger equation
(13) |
Now we could perform the unitary transformation to enter the Floquet frame, where the system-bath Hamiltonian becomes
(14) |
and
(15) |
with . We set without loss of generality.
Since the Floquet modes are periodic in time, we Fourier expand :
(16) |
where together with gives the emission spectrum. Since all modes are in vacuum, transition from to is only possible if , with the transition rate determined by and the bath spectral density at frequency . Furthermore, transition could emit photons at multiple frequencies, and both and could occur, which is different from the relaxation of static systems.
In Fig. 3(a), we plot for various transitions of the Kapitzonium under charge noise with . The dominant heating processes (red cross) and (red dot) occur at near degenerate frequency around . The reverse cooling processes (blue cross) and (blue dot) are also allowed at a different frequency around .
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 (see Appendix A), where . Here is the heating rate due to intrinsic loss and
(17) |
where
(18) |
Without the Floquet driving, is approximately the amplitude damping rate of a transmon corresponding to its .

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 . Furthermore, the environment cannot distinguish whether the emitted photon comes from the or transition since . Therefore the phase coherence between and 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 Putterman et al. (2022). The first filter mode also couples capacitively to the Kapitzonium via the interaction . The full Hamiltonian is (Fig. 3(b))
(19) |
where is the center frequency of the filter and is the coupling rate between two adjacent filter modes. The last filter mode decays into a zero temperature bath at rate , which is described by the Lindblad dissipator .
The qubit subspace is autonomously protected when the engineered cooling rate is sufficiently larger than the intrinsic heating rate of Kapitzonium. Following Ref. Putterman et al. (2022), we choose , and , such that the filter bandwidth is and the filter modes are only weakly excited. The cooling rate is about after adiabatically eliminating all filter modes Reiter and Sørensen (2012); Chamberland et al. (2022). For MHz we have 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 . 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 between the dominant heating and cooling transitions. limits the filter bandwidth and thus the maximal cooling rate . The other quantity is the degeneracy of the two heating (cooling) transitions measured by . In the degenerate regime where , such as idling with MHz, the qubit subspace is protected (Fig. 3(c)). In the non-degenerate regime where , the environment could distinguish which cooling transition the emitted photon comes from and dephase the qubit (Fig. 3(c)). Therefore the qubit basis states and are protected but not their coherent superpositions.
For X gate with relatively small MHz, the emission spectrum is similar to the idling emission spectrum. Therefore the idling filter also protects the X gate, with MHz well within the degenerate regime.
For Z gate with GHz, we have MHz and MHz. A different filter is required with center frequency at about 20 GHz. Z gate is unprotected, since the largest possible cooling rate MHz is comparable with .
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 on the order of a few MHz, since its and are similar to the X gate. However, we didn’t simulate the XX gate due to the high computational cost.

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 or (Fig. 4(a)). We could apply the X gate but with a much larger to implement the X measurement (Fig. 4(b)). Larger lifts the degeneracy between and , and are the measurement basis. On the other hand, the unprotected Z gate dephases the qubit and naturally perform the Z measurement with measurement basis (Fig. 4(b)).
The measurement rate is proportional to the occupation of the excited states and . We could increase the measurement rate by capacitively driving and at the heating transition frequency (Fig. 4(a)). The charge operator in the Floquet frame is
(20) |
where we only include terms from Eq. (15) and the dominant Fourier coefficients . We apply a charge drive to the Kapitzonium where , and . This leads to the coupling after the rotating wave approximation (RWA). Furthermore, are only weakly excited since , which decay back to 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 , we monitor the output field from the filter with a heterodyne measurement for , and then calculate the power spectrum for each measurement record . Depending on whether the system is projected into or , show a peak at or (Fig. 4(a)). We therefore define the signal as where is the integrated power within a narrow frequency window around . or 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 , the occupation of during the measurement for , and the spectrum .
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)
(21) |
where is the total junction capacitance and . To symmetrize the flux allocation inside the two cosines, we move to another reference frame by performing the unitary transformation
(22) |
where the Hamiltonian becomes
(23) |
We will drop the unwanted term for now, since in principle it can be compensated for with the gate voltage.
For disorder in junction energies, we define . With the idling flux drive , the effective Hamiltonian is
(24) |
Therefore disorder in junction energies reduces the effective , and for small 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 (idling) and (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
(25) |
where . However, this flux choice poses inconveniences for implementing the X gate and X measurement.
Alternatively, we could set with
(26) |
where we have applied the Jacobi–Anger expansion and is the -th Bessel function of the first kind. The effective Hamiltonian for Eq. (2) now becomes
(27) |
Therefore we could choose different such that for Kapitzonium idling, Z gate and Z measurement, and 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 and phase-flip rate of the Kapitzonium without cooling are 0, since at any time
(28) |
for . This is because the Kapitzonium Hamiltonian (Eq. (3)) is symmetric under the transformation of , and both and 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 (see Appendix D). Due to the Floquet driving, the dominant error is the leakage from the qubit subspace, e.g., from to . This leakage rate is proportional to .
The filter implemented in Sec. III suppresses the leakage rate by so . Such a cooling process induces dephasing error due to the frequency difference in the emitted photons, with a rate of . In the deep transmon regime , we have and the dephasing rate is exponentially small in .
In summary, the bit and phase flip rates are exponentially suppressed, while the leakage rate is reduced by the filter factor . 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 , which typically has a spectral density and thus a small spectral weight at high frequencies. The effects of depend on how we choose to drive the flux. For , 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 (Sec. IV.2) more attractive. In this case, flux noise leads to an imposed flux of where 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 qubit Groszkowski et al. (2018); Gyenis et al. (2021b). Nonetheless, as in the case of the soft 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 , the effective Hamiltonian is also insensitive to first order to around , and leading order terms scale as . Therefore operating at the sweet spot 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 basis induced by quasiparticles are proportional to , where is the index of the junction, and . Therefore for , and on average don’t accumulate extra phase and should be robust to quasiparticle tunneling . This analysis may not hold when . 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 with in the Kapitzonium Hamiltonian where is the random offset charge.
The Floquet eigenenergies weakly depend on during idling (Fig. 4(e), left figure). The weak dependence also holds when adding 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 and the emitted photon frequency depend on (Fig. 4(e), middle and right figures). For example, at we have and . Therefore both Z gate and Z measurement cannot be performed at .
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 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 in the interaction picture generates the dynamics
(29) |
Now we make the Born approximation , where is the system density matrix and is the stationary bath density matrix. In addition, we also make the standard assumption that . The system dynamics becomes
(30) |
where we define the two-point correlation functions
(31) |
To proceed, we assume that the bath is stationary with a very short correlation decay time. In other words
(32) |
where is much shorter than any time scale we are interested in. Assuming weak system-bath coupling, we have and which is second order in the coupling strength. Therefore we could make the Markov approximation and replace with in Eq. (30). Since only contribute significantly to the integration, the upper limit of the integration can be extended to :
(33) |
For Floquet systems, we have with since is Hermitian. Therefore
(34) |
where
(35) |
We could decompose into its real and imaginary parts as , where
(36) |
Physically the real part represents the decay rate while the imaginary part can be absorbed into the system Hamiltonian which we will ignore for now.
Consider a zero temperature bath with a flat spectral density function
(37) |
as an example. We could decompose into positive and negative frequency parts where
(38) |
and . Therefore
(39) |
and Eq. (33) gives the Floquet master equation
(40) |
where we apply RWA to drop 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 Nielsen (2002); Wang et al. (2022). Intuitively, describes how well the qubit subspace is preserved under some quantum process. More specifically, the average fidelity of a quantum channel over all states in a -dimensional qubit subspace is
(41) |
where
(42) |
Therefore the average fidelity can be simplified as
(43) |
To compare with some target unitary , we could compare with identity instead, where .
B.2 Unitary case
The gates are designed such that the system Hamiltonian adiabatically evolve from the idling to the gate Hamiltonian , and then adiabatically evolve back to after certain amount of gate time. The average gate fidelity is calculated for the single-qubit subspace with or the two-qubit subspace with .
We choose the pulse shape
(44) |
where is the total gate duration and is the adiabatic ramping time.
X gate. The Hamiltonian is
(45) |
where ns, ns and MHz. The X gate implements the mapping of with an infidelity of .
Z gate. The Hamiltonian is
(46) |
where
(47) |
We choose ns, ns and GHz. The Z gate implements the mapping of with an infidelity of .
Another Floquet drive that seems reasonable at first is to have frequency modulation instead of amplitude modulation:
(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
(49) |
where is the idling Hamiltonian for each qubit with . We choose ns, ns and MHz. The XX gate implements the mapping of with an infidelity of .
State initialization. To initialize the system state into , we could start from the ground state of the static transmon Hamiltonian , and adiabatically apply the Floquet drive:
(50) |
where 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 we add loss to the system. More specifically, the simulation is performed in the Floquet frame with Hamiltonian 0 and a single Lindblad dissipator . Notice that the Floquet frame here is defined with respect to instead .
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, is no longer the Floquet eigenstates of full Hamiltonian (Eq. (19) in the main text) where is the ground state of the filter. Therefore we work with the dressed Floquet eigenstates of Eq. (19) instead. The qubit basis states are chosen based on their overlap with .
The Floquet drive parameters requires a slight fine tuning due to this hybridization. During the gate time , the simulation is performed in the Floquet frame defined by with Hamiltonian 0 and two Lindblad dissipators and . Here is calculated similarly to Eq. (17) with . 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 and are the dressed Floquet eigenstates with GHz and are their even and odd superpositions. We choose GHz, MHz and 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.

Appendix C Measurement simulation
The Hamiltonian for the measurement simulation is
(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 in Eq. (20) represented by a matrix.
For X measurement, the system Hamiltonian is with MHz. can be calculated from the emission spectrum of . The charge drive frequencies are GHz and GHz. The emitted photon frequencies are GHz and GHz. We choose MHz and the filter parameters 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 GHz and GHz. The emitted photon frequencies are GHz and GHz. We choose MHz and the filter parameters are the same as the Z gate filter. Starting from the initial state , we simulate 1000 trajectories for the Z measurement with a measurement time of s (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 , , which shifts the emitted photon frequencies as well. Second of all, the charge drive frequency should be outside the filter passband. In principle setting , 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 to be comparable or larger than will cause measurement error due to crosstalk between the drives. On the other hand, a very small reduces the measurement rate and requires a long measurement time.

Appendix D Kapitzonium lifetime estimation
The Kapitzonium Hamiltonian in presence of offset charge is
(52) |
For , we have where the parity operator is and are the charge eigenstates. The Floquet eigenstates are parity eigenstates, and it turns out that both and have even parity. Therefore
(53) |
for .
For , don’t have the symmetry under 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 results to be exponentially small in . We define the total transition rate from to as
(54) |
where and . In Fig. 6, we plot the transition rates for different values of . The results indeed show an exponential suppression of both the bit-flip and phase flip rates with .
References
- Blais et al. (2021) A. Blais, A. L. Grimsmo, S. M. Girvin, and A. Wallraff, Reviews of Modern Physics 93, 025005 (2021).
- Kjaergaard et al. (2020) M. Kjaergaard, M. E. Schwartz, J. Braumüller, P. Krantz, J. I.-J. Wang, S. Gustavsson, and W. D. Oliver, Annual Review of Condensed Matter Physics 11, 369 (2020).
- Koch et al. (2007) J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Physical Review A 76, 042319 (2007).
- Manucharyan et al. (2009) V. E. Manucharyan, J. Koch, L. I. Glazman, and M. H. Devoret, Science 326, 113 (2009).
- Brooks et al. (2013) P. Brooks, A. Kitaev, and J. Preskill, Physical Review A 87, 052306 (2013).
- Gyenis et al. (2021a) A. Gyenis, P. S. Mundada, A. Di Paolo, T. M. Hazard, X. You, D. I. Schuster, J. Koch, A. Blais, and A. A. Houck, PRX Quantum 2, 010339 (2021a).
- Kalashnikov et al. (2020) K. Kalashnikov, W. T. Hsieh, W. Zhang, W.-S. Lu, P. Kamenov, A. Di Paolo, A. Blais, M. E. Gershenson, and M. Bell, PRX Quantum 1, 010307 (2020).
- Pechenezhskiy et al. (2020) I. V. Pechenezhskiy, R. A. Mencia, L. B. Nguyen, Y.-H. Lin, and V. E. Manucharyan, Nature 585, 368 (2020).
- Ofek et al. (2016) N. Ofek, A. Petrenko, R. Heeres, P. Reinhold, Z. Leghtas, B. Vlastakis, Y. Liu, L. Frunzio, S. M. Girvin, L. Jiang, M. Mirrahimi, M. H. Devoret, and R. J. Schoelkopf, Nature 536, 441 (2016).
- Hu et al. (2019) L. Hu, Y. Ma, W. Cai, X. Mu, Y. Xu, W. Wang, Y. Wu, H. Wang, Y. P. Song, C.-L. Zou, S. M. Girvin, L.-M. Duan, and L. Sun, Nature Physics 15, 503 (2019).
- Campagne-Ibarcq et al. (2020) P. Campagne-Ibarcq, A. Eickbusch, S. Touzard, E. Zalys-Geller, N. E. Frattini, V. V. Sivak, P. Reinhold, S. Puri, S. Shankar, R. J. Schoelkopf, L. Frunzio, M. Mirrahimi, and M. H. Devoret, Nature 584, 368 (2020).
- Groszkowski et al. (2018) P. Groszkowski, A. D. Paolo, A. L. Grimsmo, A. Blais, D. I. Schuster, A. A. Houck, and J. Koch, New Journal of Physics 20, 043053 (2018).
- Gyenis et al. (2021b) A. Gyenis, A. Di Paolo, J. Koch, A. Blais, A. A. Houck, and D. I. Schuster, PRX Quantum 2, 030101 (2021b).
- Mirrahimi et al. (2014) M. Mirrahimi, Z. Leghtas, V. V. Albert, S. Touzard, R. J. Schoelkopf, L. Jiang, and M. H. Devoret, 16, 045014 (2014).
- Leghtas et al. (2015) Z. Leghtas, S. Touzard, I. M. Pop, A. Kou, B. Vlastakis, A. Petrenko, K. M. Sliwa, A. Narla, S. Shankar, M. J. Hatridge, M. Reagor, L. Frunzio, R. J. Schoelkopf, M. Mirrahimi, and M. H. Devoret, Science 347, 853 (2015).
- Lescanne et al. (2020) R. Lescanne, M. Villiers, T. Peronnin, A. Sarlette, M. Delbecq, B. Huard, T. Kontos, M. Mirrahimi, and Z. Leghtas, Nature Physics 16, 509 (2020).
- Puri et al. (2017) S. Puri, S. Boutin, and A. Blais, npj Quantum Information 3, 1 (2017).
- Grimm et al. (2020) A. Grimm, N. E. Frattini, S. Puri, S. O. Mundhada, S. Touzard, M. Mirrahimi, S. M. Girvin, S. Shankar, and M. H. Devoret, Nature 584, 205 (2020).
- Kapit (2016) E. Kapit, Physical Review Letters 116, 150501 (2016).
- Li et al. (2023) Z. Li, T. Roy, D. R. Perez, K.-H. Lee, E. Kapit, and D. I. Schuster, “Autonomous error correction of a single logical qubit using two transmons,” (2023), arXiv:2302.06707 [quant-ph] .
- Eckardt (2017) A. Eckardt, Reviews of Modern Physics 89, 011004 (2017).
- Venkatraman et al. (2022) J. Venkatraman, X. Xiao, R. G. Cortiñas, A. Eickbusch, and M. H. Devoret, Physical Review Letters 129, 100601 (2022).
- Didier et al. (2019) N. Didier, E. A. Sete, J. Combes, and M. P. da Silva, Physical Review Applied 12, 054015 (2019).
- Huang et al. (2021) Z. Huang, P. S. Mundada, A. Gyenis, D. I. Schuster, A. A. Houck, and J. Koch, Physical Review Applied 15, 034065 (2021).
- Gandon et al. (2022) A. Gandon, C. Le Calonnec, R. Shillito, A. Petrescu, and A. Blais, Physical Review Applied 17, 064006 (2022).
- Landau and Lifshits (1976) L. D. Landau and E. M. Lifshits, Mechanics, 3rd ed., Course of Theoretical Physics No. v. 1 (Pergamon Press, Oxford New York, 1976).
- Vool and Devoret (2017) U. Vool and M. Devoret, International Journal of Circuit Theory and Applications 45, 897 (2017).
- You et al. (2019) X. You, J. A. Sauls, and J. Koch, Physical Review B 99, 174512 (2019).
- Rajabzadeh et al. (2022) T. Rajabzadeh, Z. Wang, N. Lee, T. Makihara, Y. Guo, and A. H. Safavi-Naeini, “Analysis of arbitrary superconducting quantum circuits accompanied by a Python package: SQcircuit,” (2022), arXiv:2206.08319 [cond-mat, physics:physics, physics:quant-ph] .
- Chitta et al. (2022) S. P. Chitta, T. Zhao, Z. Huang, I. Mondragon-Shem, and J. Koch, (2022), 10.48550/arXiv.2206.08320.
- Murch et al. (2012) K. W. Murch, U. Vool, D. Zhou, S. J. Weber, S. M. Girvin, and I. Siddiqi, Physical Review Letters 109, 183602 (2012).
- Putterman et al. (2022) H. Putterman, J. Iverson, Q. Xu, L. Jiang, O. Painter, F. G. S. L. Brandão, and K. Noh, Physical Review Letters 128, 110502 (2022).
- Reiter and Sørensen (2012) F. Reiter and A. S. Sørensen, Physical Review A 85, 032111 (2012).
- Chamberland et al. (2022) C. Chamberland, K. Noh, P. Arrangoiz-Arriola, E. T. Campbell, C. T. Hann, J. Iverson, H. Putterman, T. C. Bohdanowicz, S. T. Flammia, A. Keller, G. Refael, J. Preskill, L. Jiang, A. H. Safavi-Naeini, O. Painter, and F. G. Brandão, PRX Quantum 3, 010329 (2022).
- Wiseman and Milburn (2009) H. M. Wiseman and G. J. Milburn, Quantum Measurement and Control, 1st ed. (Cambridge University Press, 2009).
- Riwar and DiVincenzo (2022) R.-P. Riwar and D. P. DiVincenzo, npj Quantum Information 8, 1 (2022).
- Lutchyn et al. (2005) R. Lutchyn, L. Glazman, and A. Larkin, Physical Review B 72, 014517 (2005).
- Lutchyn et al. (2006) R. M. Lutchyn, L. I. Glazman, and A. I. Larkin, Physical Review B 74, 064515 (2006).
- Martinis et al. (2009) J. M. Martinis, M. Ansmann, and J. Aumentado, Physical Review Letters 103, 097002 (2009).
- Catelani et al. (2011) G. Catelani, J. Koch, L. Frunzio, R. J. Schoelkopf, M. H. Devoret, and L. I. Glazman, Physical Review Letters 106, 077002 (2011).
- Serniak et al. (2018) K. Serniak, M. Hays, G. de Lange, S. Diamond, S. Shankar, L. D. Burkhart, L. Frunzio, M. Houzet, and M. H. Devoret, Physical Review Letters 121, 157701 (2018).
- Houzet et al. (2019) M. Houzet, K. Serniak, G. Catelani, M. H. Devoret, and L. I. Glazman, Physical Review Letters 123, 107704 (2019).
- Grifoni and Hänggi (1998) M. Grifoni and P. Hänggi, Physics Reports 304, 229 (1998).
- Nielsen (2002) M. A. Nielsen, Physics Letters A 303, 249 (2002).
- Wang et al. (2022) Z. Wang, T. Rajabzadeh, N. Lee, and A. H. Safavi-Naeini, PRX Quantum 3, 020302 (2022).
- Kapit (2018) E. Kapit, Physical Review Letters 120, 050503 (2018).
- Johansson et al. (2012) J. Johansson, P. Nation, and F. Nori, Computer Physics Communications 183, 1760 (2012).
- Johansson et al. (2013) J. R. Johansson, P. D. Nation, and F. Nori, Computer Physics Communications 184, 1234 (2013).