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

Current address: ]Nord Quantique, Sherbrooke, Québec, J1K 0A5, Canada

Fast parametric two-qubit gates with suppressed residual interaction
using a parity-violated superconducting qubit

Atsushi Noguchi1,2,3 u-atsushi@g.ecc.u-tokyo.ac.jp    Alto Osada1,3    Shumpei Masuda4    Shingo Kono2,5    Kentaro Heya2    Samuel Piotr Wolski2    Hiroki Takahashi2    Takanori Sugiyama2    Dany Lachance-Quirion2 [    Yasunobu Nakamura2,5 1Komaba Institute for Science (KIS), The University of Tokyo, Meguro-ku, Tokyo, 153-8902, Japan,
2Research Center for Advanced Science and Technology (RCAST), The University of Tokyo, Meguro-ku, Tokyo, 153-8904, Japan,
3PRESTO, Japan Science and Technology Agency, Kawaguchi-shi, Saitama 332-0012, Japan,
4College of Liberal Arts and Sciences, Tokyo Medical and Dental University, Ichikawa, Chiba 272-0827, Japan,
5Center for Emergent Matter Science (CEMS), RIKEN, Wako-shi, Saitama 351-0198, Japan
Abstract

We demonstrate fast two-qubit gates using a parity-violated superconducting qubit consisting of a capacitively-shunted asymmetric Josephson-junction loop under a finite magnetic flux bias. The second-order nonlinearity manifesting in the qubit enables the interaction with a neighboring single-junction transmon qubit via first-order inter-qubit sideband transitions with Rabi frequencies up to 30 MHz. Simultaneously, the unwanted static longitudinal (ZZ) interaction is eliminated with ac Stark shifts induced by a continuous microwave drive near-resonant to the sideband transitions. The average fidelities of the two-qubit gates are evaluated with randomized benchmarking as 0.967, 0.951, 0.956 for CZ, iSWAP and SWAP gates, respectively.

Quantum information processing with superconducting qubits has been intensively studied recently. High-fidelity quantum manipulations and projective measurements have been achieved in multi-qubit systems martinis2014 ; ibm2016 ; rigetti ; martinis2019 ; johnson2020 , and basic quantum error-correction protocols have been demonstrated schoelkopf2012 ; martinis2015 ; schoelkopf2016 ; correct2019 ; wallraf2019 . For fault-tolerant quantum computing, however, the gate and readout fidelity should be further improved by a few orders of magnitude fowler ; terhal .

To this end, a variety of two-qubit gates have been proposed and demonstrated. These can be classified into two groups, based on their use of either a coupling between (near-)degenerate qubits martinis2014 ; martinis2019 or a microwave-induced parametric coupling nakamura2007 ; ibm2011 ; ibm2016 ; rigetti ; johnson2020 ; screview . For the gate operation, the former usually requires fast frequency tuning of the qubits and/or a coupler through a flux bias, while the latter only uses microwave pulses for the dynamical control. For the parametric gates, qubits are usually far off-resonant from each other in order to suppress residual couplings between them. On the other hand, a large detuning slows down the parametric gate, causing a trade-off that hinders the improvement of the gate fidelity. Recent works have addressed this issue by introducing various types of coupler circuits to eliminate the residual coupling without sacrificing the gate speed significantly DiVincenzo2017 ; steele2018 ; houck2019 ; duan2020 . A simpler scheme combining two qubits with opposite signs of anharmonicity also allows a residual-coupling-free two-qubit cross-resonance gate plourde2020 .

In this letter, we propose and demonstrate fast parametric two-qubit gates using sideband transitions between an ordinary transmon qubit and a parity-violated superconducting qubit, which we call a cubic transmon. The parity violation originates in a cubic component of the inductive potential of a Josephson-junction circuit under a finite magnetic flux bias. This circuit, known as a superconducting nonlinear asymmetric inductive element (SNAIL), was recently proposed SNAIL and utilized in parametric amplifiers SNAIL2 , bosonic-mode qubits SNAIL3 , and hybrid quantum systems noguchi2018 . The parity symmetry breaking is essential for a physical system to acquire a second-order nonlinearity, which allows three-wave-mixing-type first-order sideband transitions and thus the parametric interactions with a neighboring qubit nakamura2015 ; sideband ; DiVincenzo2016 . The large capacitive coupling strength between the qubits introduces strong parametric interactions, but also a large residual static interaction. We eliminate the latter by using ac Stark shifts induced by a continuous near-resonant drive of the sideband transitions and solve the trade-off. This approach of microwave-assisted elimination of the static interactions brings in more tuning knobs, i.e. amplitudes and frequencies of multiple drives, which can be applied to the cases with multiple qubits and higher-order residual couplings.

Refer to caption
Figure 1: (a) Photographs of the device. Most of the structures are made from Nb electrodes (light gray) on a Si substrate (dark gray), and the Josephson junctions (at the three crosses in the bottom picture) are made of Al/AlOx/Al junctions evaporated together with Al electrodes (white). Air-bridges across the coplanar resonators and transmission lines suppress spurious modes on the chip. (b) Circuit diagram of the device. (c) Eigenstates |ij~\widetilde{|ij\rangle} (i,j{g,e,f}i,j\in\{g,e,f\}) of the two qubit system. The vertical axis indicates the eigenfrequency of the states.
Refer to caption
Figure 2: Elimination of the static ZZ interaction with a continuous-wave (CW) drive. The black and red dots are respectively the frequencies of the |gg~|ge~\widetilde{|gg\rangle}\leftrightarrow\widetilde{|ge\rangle} and |eg~|ee~\widetilde{|eg\rangle}\leftrightarrow\widetilde{|ee\rangle} transitions, determined by spectroscopy. The inset shows the energy diagram. The black and red arrows respectively indicate the corresponding transitions, and the green arrows represent the CW drive, which simultaneously couples to all the sideband transitions. The blue arrows show the directions of the ac Stark shifts of the eigenstates. The transition frequencies become identical at the CW-drive power of 10\sim 10 nW, and the static ZZ interaction is eliminated.

Figures 1(a) and 1(b) present optical micrographs and a circuit diagram of the device, which contains two superconducting qubits and two resonators for the dispersive readout of each qubit. The qubit on the right-hand side is a conventional transmon, which is composed of a capacitively-shunted single Josephson junction koch2007 . The other qubit is a cubic transmon, which is a capacitively-shunted SNAIL circuit. The SNAIL is a Josephson-junction loop formed by a parallel circuit of a single small Josephson-junction and two large Josephson junctions. The SNAIL loop is threaded by a flux Φ\Phi. The Hamiltonian of the two-qubit system reads

H^/=\displaystyle\hat{H}/\hbar=\!\!\!\!\! ωc0a^a^+βc0(a^a^a^+a^a^a^)+αc02a^a^a^a^\displaystyle\omega_{\mathrm{c0}}\hat{a}^{\dagger}\hat{a}+\beta_{\mathrm{c0}}(\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}+\hat{a}^{\dagger}\hat{a}\hat{a})+\frac{\alpha_{\mathrm{c0}}}{2}\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}\hat{a} (1)
+ωt0b^b^+αt02b^b^b^b^+g0(a^b^+a^b^),\displaystyle+\omega_{\mathrm{t0}}\hat{b}^{\dagger}\hat{b}+\frac{\alpha_{\mathrm{t0}}}{2}\hat{b}^{\dagger}\hat{b}^{\dagger}\hat{b}\hat{b}+g_{0}(\hat{a}^{\dagger}\hat{b}+\hat{a}\hat{b}^{\dagger}),

where =h/2π\hbar=h/2\pi is the reduced Planck constant, ωc0\omega_{\mathrm{c0}} and ωt0\omega_{\mathrm{t0}} are the bare eigenmode frequencies, and a^\hat{a} and b^\hat{b} are the annihilation operators for the cubic transmon and conventional transmon, respectively. The coefficient βc0\beta_{\mathrm{c0}} is the second-order nonlinearity of the cubic transmon, αc0\alpha_{\mathrm{c0}} and αt0\alpha_{\mathrm{t0}} are the third-order nonlinearities of each qubit, and g0g_{0} is the capacitive coupling strength between the two qubits.

In the dispersive coupling regime |Δ0||ωc0ωt0|g0|\Delta_{0}|\equiv|\omega_{\mathrm{c0}}-\omega_{\mathrm{t0}}|\gg g_{0}, the effective Hamiltonian can be written as

H^eff/=\displaystyle\hat{H}_{\mathrm{eff}}/\hbar=\!\!\!\!\! [ωc+g(b^+b^)]a^a^+αc2a^a^a^a^\displaystyle\left[\omega_{\mathrm{c}}+g(\hat{b}^{\dagger}+\hat{b})\right]\hat{a}^{\dagger}\hat{a}+\frac{\alpha_{\mathrm{c}}}{2}\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}\hat{a} (2)
+ωtb^b^+αt2b^b^b^b^+JZZa^a^b^b^,\displaystyle+\omega_{\mathrm{t}}\hat{b}^{\dagger}\hat{b}+\frac{\alpha_{\mathrm{t}}}{2}\hat{b}^{\dagger}\hat{b}^{\dagger}\hat{b}\hat{b}+J_{\mathrm{ZZ}}\hat{a}^{\dagger}\hat{a}\hat{b}^{\dagger}\hat{b},

where gg(βc0\propto\beta_{\mathrm{c0}}) is the effective coupling strength, and ωc\omega_{\mathrm{c}}, ωt\omega_{\mathrm{t}}, αc\alpha_{\mathrm{c}}, and αt\alpha_{\mathrm{t}} are the eigenmode frequencies and self-Kerr nonlinearities of the qubits after the perturbative treatment of the coupling term in Eq. (1), respectively. The term with a coefficient gg arises from the second-order nonlinearity in the parity-violated cubic transmon and gives the interaction in the same form as the radiation pressure in optomechanics optomechanics ; noguchi2018 and the state-dependent force in trapped ions blatt2008 ; ion . There is also a static longitudinal (ZZ) interaction between the qubits, whose amplitude is JZZJ_{\mathrm{ZZ}}. The detailed derivations and expressions of the parameters in Eqs. (1) and (2) are presented in the Supplementary Material supple .

Figure 1(c) illustrates the eigenstates |ij~\widetilde{|ij\rangle} (i,j{g,e,f})(i,j\in\{g,e,f\}) and their frequencies. The coupling between qubits hybridizes the bare qubit states, |ic|jt|i\rangle_{\mathrm{c}}|j\rangle_{\mathrm{t}}, and forms the eigenstates |ij~\widetilde{|ij\rangle} supple . When a drive field at frequency ωd=Δωtωc\omega_{\mathrm{d}}=\Delta\equiv\omega_{\mathrm{t}}-\omega_{\mathrm{c}} is applied to the cubic transmon, the two qubits resonate with each other in the rotating frame. Under the rotating-wave approximation, the parametric coupling follows

H^p/\displaystyle\hat{H}_{\mathrm{p}}/\hbar\!\! =\displaystyle= ηΩ(eiωdt+iθa^b^+eiωdtiθa^b^),\displaystyle\!\!\eta\Omega(e^{i\omega_{\mathrm{d}}t+i\theta}\hat{a}^{\dagger}\hat{b}+e^{-i\omega_{\mathrm{d}}t-i\theta}\hat{a}\hat{b}^{\dagger}), (3)
η\displaystyle\eta\!\! \displaystyle\equiv 2g0βc0(2ωc02αc0Δ0+2αc0ωc0)Δ0(Δ0ωc0)(αc0+ωc0)(αc0+ωc0+Δ0),\displaystyle\!\!\frac{-2g_{0}\beta_{\mathrm{c0}}(2\omega_{\mathrm{c0}}^{2}-\alpha_{\mathrm{c0}}\Delta_{0}+2\alpha_{\mathrm{c0}}\omega_{\mathrm{c0}})}{\Delta_{0}(\Delta_{0}-\omega_{\mathrm{c0}})(\alpha_{\mathrm{c0}}+\omega_{\mathrm{c0}})(\alpha_{\mathrm{c0}}+\omega_{\mathrm{c0}}+\Delta_{0})},~{} (4)

where Ω\Omega and θ\theta are the amplitude and phase of the drive field, respectively. Note that η\eta is proportional to the second-order nonlinearity βc0\beta_{\mathrm{c0}} for the three-wave-mixing process occurring under the Hamiltonian H^p\hat{H}_{\mathrm{p}} in Eq. (3).

Refer to caption
Figure 3: Implementation of the two-qubit gates. (a) Composite pulses for CZ, iSWAP and SWAP gates, applied to the cubic transmon. The frequencies of the blue and orange pulses are resonant to the |ge~|eg~\widetilde{|ge\rangle}\leftrightarrow\widetilde{|eg\rangle} and |ee~|gf~\widetilde{|ee\rangle}\leftrightarrow\widetilde{|gf\rangle} transitions, respectively. The relative phases of the second segment of the orange pulses are fine-tuned to optimize the amount of the conditional phase. (b) [(c)] Pulse sequence for characterizing the two-qubit gates, identity and CZ [iSWAP and SWAP], and corresponding transitions. Xτ(ϕ)X_{\tau}^{(\phi)} (Xτ(θ)X_{\tau}^{(\theta)}) represents a single-qubit rotation with the rotation angle τ\tau and phase ϕ\phi (θ\theta). The right panels in (b) and (c) are energy-level diagrams illustrating the Ramsey interferometries involving a two-qubit gate. The black solid and red dashed arrows indicate pairs of interfering eigenstates. The blue and orange arrows indicate the sideband transitions. (d) [(e)] Ramsey-type interference, conditioned on the state of the cubic transmon, with the two-qubit gates such as identity and CZ (iSWAP and SWAP) gates. The excitation probability of the transmon, σ¯z,t\bar{\sigma}_{z,\mathrm{t}} (the cubic transmon, σ¯z,c\bar{\sigma}_{z,\mathrm{c}}) is determined by the average readout in the time-ensemble measurement. The black (red) dots show the experimental results without (with) the initial XπX_{\pi}-pulse [red dashed rectangles in (b) and (c)] for the cubic transmon. The black solid and red dashed curves represent the functions for the ideal gates.

Under the resonant condition ωd=Δ\omega_{\mathrm{d}}=\Delta, the drive exchanges the excitation of the two qubits, and thus the iSWAP and SWAP gates can be implemented. Another type of two-qubit gate, controlled-phase (CZ) gate, is similarly achieved with a parametric drive. When the drive frequency is equal to Δ+αt\Delta+\alpha_{\mathrm{t}}, the transition |ee~|gf~\widetilde{|ee\rangle}\leftrightarrow\widetilde{|gf\rangle} takes place. A 2π2\pi-rotation of the transition induces a geometric phase factor of 1-1 only to the |ee~\widetilde{|ee\rangle} state in the computational subspace.

In parallel with the dynamically-induced coupling, there remains the spurious static ZZ interaction, the last term in Eq. (2), between the capacitively coupled qubits with higher energy levels ibm2019 . Remarkably, the residual interaction can be eliminated also with a parametric drive. We irradiate the cubic transmon with a continuous-wave (CW) microwave field, whose frequency is slightly detuned from the transition of |ee~|fg~\widetilde{|ee\rangle}\leftrightarrow\widetilde{|fg\rangle}. The ac Stark effect by the CW drive shifts the eigenfrequencies in the two-qubit subspace. These shifts give rise to a tunable ZZ interaction and allow compensation for the unwanted interaction.

In the experiment, we use the device shown in Fig. 1. The parameters at the operating flux bias, Φ=0.34Φ0\Phi=0.34\Phi_{0}, where Φ0h/2e\Phi_{0}\equiv h/2e, are the followings: The eigenfrequencies of the cubic transmon and the transmon are ωc/2π=3.633GHz\omega_{\mathrm{c}}/2\pi=3.633\mathrm{~{}GHz} and ωt/2π=4.479GHz\omega_{\mathrm{t}}/2\pi=4.479\mathrm{~{}GHz}, respectively. The third-order nonlinearities of the qubits are αc/2π=132MHz\alpha_{\mathrm{c}}/2\pi=-132~{}\mathrm{MHz} and αt/2π=168MHz\alpha_{\mathrm{t}}/2\pi=-168~{}\mathrm{MHz}, and the bare coupling strength between the qubits is g0/2π=75MHzg_{0}/2\pi=75\mathrm{~{}MHz}, which are determined by spectroscopic measurements. The details of the sample characterization are described in the Supplementary Material supple . Using these values, we estimate the second-order nonlinearity βc/2π=195MHz\beta_{\mathrm{c}}/2\pi=-195\mathrm{~{}MHz}, the effective coupling strength g/2π=14MHzg/2\pi=-14\mathrm{~{}MHz}, and the coupling coefficient of the parametric drive, η=0.022\eta=0.022. The energy-relaxation and Ramsey-dephasing times of the qubits are T1=3.9μsT_{1}=3.9\mathrm{~{}\mu s} and T2=0.6μsT_{2}^{*}=0.6\mathrm{~{}\mu s} for the cubic transmon, and T1=4.0μsT_{1}=4.0\mathrm{~{}\mu s} and T2=2.3μsT_{2}^{*}=2.3\mathrm{~{}\mu s} for the transmon, respectively. The dephasing time of the cubic transmon is improved to T2E=1.5μsT_{2}^{\mathrm{E}}=1.5\mathrm{~{}\mu s} with an echo pulse, while no change is seen for the transmon.

We first eliminate the residual ZZ interaction by the CW drive (Fig. 2). The drive frequency is 930 MHz, and the detuning from the |ge~|eg~\widetilde{|ge\rangle}\leftrightarrow\widetilde{|eg\rangle} transition is 84 MHz. The inset of Fig. 2 shows the shifts of eigenstates induced by the CW drive. The CW drive is red detuned from the |ee~|fg~\widetilde{|ee\rangle}\leftrightarrow\widetilde{|fg\rangle} transition and blue detuned from the |ge~|eg~\widetilde{|ge\rangle}\leftrightarrow\widetilde{|eg\rangle} and |ee~|gf~\widetilde{|ee\rangle}\leftrightarrow\widetilde{|gf\rangle} transitions. Thus, the sign of the frequency shift is different from each other. The amplitude of the ZZ interaction corresponds to the frequency difference between the |gg~|ge~\widetilde{|gg\rangle}\leftrightarrow\widetilde{|ge\rangle} and |eg~|ee~\widetilde{|eg\rangle}\leftrightarrow\widetilde{|ee\rangle} transitions, which amounts to 55 MHz in the absence of the CW drive. The frequency difference vanishes at a certain power of the drive. It is also found that the CW drive does not degrade the coherence of the qubits (data not shown).

In the presence of the CW drive, we implement the two-qubit Clifford gate set, i.e. CZ, iSWAP and SWAP gates, using parametric couplings induced by additional microwave pulses. These gates are within the family of the Fermionic Simulation gate set, characterized by two parameters, the swap angle θsw\theta_{\mathrm{sw}} and the conditional phase θcp\theta_{\mathrm{cp}} fermi2018 ; martinis2020 . The CZ, iSWAP, and SWAP gates have the parameters (θsw,θcp)=(0,π)(\theta_{\mathrm{sw}},\theta_{\mathrm{cp}})=(0,\pi), (π,0)(\pi,0) and (π,π)(\pi,\pi), respectively. In our setup, θsw\theta_{\mathrm{sw}} and θcp\theta_{\mathrm{cp}} are independently and simultaneously controlled via parametric couplings.

Figure 3(a) illustrates the waveforms of the synthesized two-tone pulses for these gates. The total gate time is 50 ns for each. The swap pulse (blue) is resonant to the |ge~|eg~\widetilde{|ge\rangle}\leftrightarrow\widetilde{|eg\rangle} transition and is used to control θsw\theta_{\mathrm{sw}}. The control-phase pulse (orange) is resonant to the |ee~|gf~\widetilde{|ee\rangle}\leftrightarrow\widetilde{|gf\rangle} transition and controls θcp\theta_{\mathrm{cp}} through the relative phase between two serial segments, applying the conditional phase θcp\theta_{\mathrm{cp}} as a geometrical phase only to the |ee~\widetilde{|ee\rangle} state. Because the swap pulses also generate a small conditional phase due to the Stark shift, we simultaneously apply a control-phase pulse for iSWAP and SWAP gate to eliminate the unwanted phase.

Figure 3(b) [3(c)] shows the pulse sequence of the Ramsey interferometry for characterizing the identity and CZ gates [iSWAP and SWAP gates]. For the iSWAP and SWAP gates, which exchange an excitation between the qubits, we apply the second π\pi/2-pulse to the cubic transmon instead of the transmon to form an interferometric sequence [Fig. 3(c)], in contrast to the standard Ramsey experiments. Figures 3(d) and 3(e) show the experimental data of Ramsey oscillations, conditioned on the state of the cubic transmon, revealing the amount of the conditional phase θcp\theta_{\mathrm{cp}} of each two-qubit gate. The phase difference between the Ramsey oscillations, with and without an initial π\pi-rotation of the cubic transmon, corresponds to the conditional phase shift. The experimental data have good agreements with the ideal behaviors in Figs. 3(d) and 3(e).

Refer to caption
Figure 4: Randomized benchmarking (RB) for two-qubit gates. The vertical axes show the normalized average quadrature amplitude σ¯z,t\bar{\sigma}_{z,\mathrm{t}} of the transmon readout signal RB2 . The horizontal axes show the number of Clifford gates in the randomized sequence. The red dots show the result of standard RB with two-qubit gates, and the blue dots express those of interleaved RB for (a) CZ, (b) iSWAP and (c) SWAP gates, respectively.

Finally, we characterize the average two-qubit gate fidelities with the randomized-benchmarking (RB) protocols RBth1 ; RBth2 ; RB1 . The gate time is uniformly set to 50 ns for the CZ, iSWAP, SWAP gates and all the single-qubit gates. Figure 4 shows the experimental results of the two-qubit RB. From the standard RB, the average gate fidelity of the two-qubit Clifford gates is determined to be 0.950±0.0010.950\pm 0.001. Using this value and those from the interleaved RB, we estimate the average gate fidelity of each two-qubit gate: 0.971±0.0020.971\pm 0.002, 0.958±0.0010.958\pm 0.001 and 0.962±0.0010.962\pm 0.001 for CZ, iSWAP and SWAP gates, respectively. The achieved fidelities of the two-qubit gates are comparable to those of the single-qubit gates supple and mostly limited by the energy relaxation time of the qubits. The coherence limits are approximately 0.97 according to the gate pulse widths, which are close to the observed fidelities. As the cubic transmon has basically the same layout as conventional transmons, we expect improvement of the relaxation time through optimizations of the design and fabrication.

In conclusion, we have demonstrated a parity-violated qubit called a cubic transmon, and realized microwave-controlled fast two-qubit gates between a cubic transmon and a conventional transmon. As the gates originate from the second-order nonlinearity of the circuit, the coupling strength scales inversely proportional to the detuning between the qubits, not to the square of it, and is thus sufficiently large for a wide detuning range of the qubits. This is advantageous for a multi-qubit system, which often suffers from a frequency-crowding problem. The residual static ZZ interaction is eliminated by applying a continuous microwave field, which will allow us to increase the bare coupling strength further and make the two-qubit gates as fast as 20 ns with optimal device parameters. This scheme for the suppression of the residual coupling can also be extended to multi-qubit systems as well as to higher-order interactions by using multiple drives.

The authors acknowledge Y. Sunada, K. Nittoh and K. Kusuyama for the help in sample fabrication and W. Oliver for providing a TWPA. This work was partly supported by JSPS KAKENHI (Grant Number 26220601, 18K03486), JST PRESTO (Grant Number JPMJPR1429), JST ERATO (Grant Number JPMJER1601), and Q-LEAP (Grant Number JPMXS0118068682).

References

  • (1) R. Barends et al., Nature 508, 500 (2014).
  • (2) F. Arute et al., Nature 574, 505 (2019).
  • (3) S. S. Hong et al., Phys. Rev. A 101, 012302 (2020).
  • (4) S. A. Caldwell et al., Phys. Rev. Appl. 10, 034050 (2018).
  • (5) S. Sheldon, E. Magesan, J. M. Chow, and J. M. Gambetta, Phys. Rev. A 93, 060302(R) (2016).
  • (6) M. D. Reed et al., Nature 482, 382 (2012).
  • (7) J. Kelly et al., Nature 519, 66 (2015).
  • (8) N. Ofek et al., Nature 536, 441 (2016).
  • (9) L. Hu et al., Nat. Phys. 15, 503 (2019).
  • (10) C. K. Andersen et al., npj Quantum Inf. 5, 69 (2019).
  • (11) A. G. Fowler, M. Mariantoni, J. M. Martinis, and A. N. Cleland, Phys. Rev. A 86, 032324 (2012).
  • (12) B. M. Terhal et al., arXiv:2002.11008 (2002).
  • (13) A. O. Niskanen et al., Science 316, 723 (2007).
  • (14) J. M. Chow et al., Phys. Rev. Lett. 107, 080502 (2011).
  • (15) P. Krantz et al., Appl. Phys. Rev. 6, 021318 (2019).
  • (16) S. Richer, N. Maleeva, S. T. Skacel, I. M. Pop, D. DiVincenzo, Phys. Rev. B 96, 174520 (2017).
  • (17) M. Kounalakis, C. Dickel, A. Bruno, N. K. Langford, and G. A. Steele, npj Quantum Information 4, 38 (2018).
  • (18) P. Mundada, G. Zhang, T. Hazard, and A. Houck, Phys. Rev. Appl. 12, 054023 (2019).
  • (19) X. Han, T. Cai, X. Li, Y. Wu, Y. Ma, J. Wang, H. Zhang, Y. Song, and L. Duan, arXiv:2003.08542 (2020).
  • (20) J. Ku et al., arXiv:2003.02775 (2020).
  • (21) N. E. Frattini, U. Vool, S. Shankar, A. Narla, K. M. Sliwa and M. H. Devoret, Appl. Phys. Lett. 110, 222603 (2017).
  • (22) N. E. Frattini, V. V. Sivak, A. Lingenfelter, S. Shankar, and M. H. Devoret, Phys. Rev. Applied 10, 054020 (2018).
  • (23) V. V. Sivak, N. E. Frattini, V. R. Joshi, A. Lingenfelter, S. Shankar, and M. H. Devoret, Phys. Rev. Applied 11, 054060 (2019).
  • (24) A. Noguchi, R. Yamazaki, Y. Tabuchi, and Y. Nakamura, Nat. Commun. 11, 1183 (2020).
  • (25) P.-M. Billangeon, J. S. Tsai, and Y. Nakamura, Phys. Rev. B 91, 094517 (2015).
  • (26) S. Richer and D. DiVincenzo, Phys. Rev. B 93, 134501 (2016).
  • (27) F. Beaudoin, M. P. daSilva, Z. Dutton, A. Blais, et al., Phys. Rev. A 86, 022305 (2012).
  • (28) J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Phys. Rev A, 76, 042319 (2007).
  • (29) M. Aspelmeyer, T. J. Kippenberg and F. Marquardt, Rev. Mod. Phys. 86, 1391 (2014).
  • (30) H. Häffner, C. F. Roos, and R. Blatt, Physics Reports 469, 155 (2008).
  • (31) F. Mintert and C. Wunderlich, Phys. Rev. Lett. 87, 257904 (2001).
  • (32) Supplementary Material
  • (33) M. Ware et al., arXiv:1905.11480 (2019).
  • (34) I. D. Kivlichan, J. McClean, N. Wiebe, C. Gidney, A. Aspuru-Guzik, Garnet Kin-Lic Chan, and R. Babbush, Phys. Rev. Lett. 120, 110501 (2018).
  • (35) B. Foxen et al., arXiv:2001.08343 (2020).
  • (36) E. Magesan, J. M. Gambetta, J. Emerson, Phys. Rev. Lett. 106, 180504 (2011).
  • (37) E. Magesan et al., Phys. Rev. Lett. 109, 080505 (2012).
  • (38) A. D. Córcoles, J. M. Gambetta, J. M. Chow, J. A. Smolin, M. Ware, J. Strand, B. L. T. Plourde, M. Steffen, Phys. Rev. A 87, 030301(R) (2013).
  • (39) A. D. Córcoles et al., Nat. Commun. 6, 6979 (2015).

Supplementary Material for
“Fast parametric two-qubit gates with suppressed residual interaction
using a parity-violated superconducting qubit”

Refer to caption
Figure S1: Wiring scheme of the experimental setup. (a) Connections from the sample chip to the ports at room temperature. (b) Pulse generating system for the qubit control. (c) Readout system. All the local oscillators (LOs) in (b) and (c) are phase-locked with a 10-MHz reference clock.

I Measurement setup

Figure S1 illustrates the wiring scheme for the gate experiments. The sample chip is connected to one readout port (C) and two drive ports (A, B). We apply microwave pulses generated by modulating the local oscillator signals. The qubits are simultaneously read out with the dispersive technique. The resonance frequencies and total decay rates of the readout resonators are 6.767 GHz and 0.8 MHz for the cubic transmon and 6.509 GHz and 1.0 MHz for the transmon, respectively. The reflection signals of the readout resonators are amplified by a Josephson traveling wave parametric amplifier (TWPA) and two low-noise amplifiers and demodulated for the readout. We apply a magnetic flux into the SNAIL loop through an external superconducting coil.

II SINGLE-PHASE APPROXIMATION OF CUBIC TRANSMON

Figure S2 shows the full-circuit model of a cubic transmon. Each of the two isolated superconducting islands has two degrees of freedom of the phase and charge. The full Hamiltonian is written as

H\displaystyle H =\displaystyle= KEJ1cosϕ1EJ2cosϕ2EJ3cos(ϕ+ϕ1ϕ2),\displaystyle K-E_{\mathrm{J1}}\cos\phi_{1}-E_{\mathrm{J2}}\cos\phi_{2}-E_{\mathrm{J3}}\cos(\phi+\phi_{1}-\phi_{2}), (S1)
K\displaystyle K =\displaystyle= 4EC(n1,n2)C(n1n2),\displaystyle 4E_{\mathrm{C}}\,\left(n_{1},n_{2}\right)\textbf{C}\left(\begin{array}[]{c}n_{1}\\ n_{2}\end{array}\right), (S4)
C =\displaystyle= (1+k1+k3k3k3k2+k3)1,\displaystyle\left(\begin{array}[]{cc}1+k_{1}+k_{3}&-k_{3}\\ -k_{3}&k_{2}+k_{3}\end{array}\right)^{\!-1}, (S7)

where n1n_{1} and n2n_{2} are the numbers of excess Cooper pairs on each island, ϕ1\phi_{1} and ϕ2\phi_{2} are the superconducting phases across each junction connected to the ground, ϕ=2πΦ/Φ0\phi=2\pi\Phi/\Phi_{0} is the reduced magnetic flux, Φ\Phi is the flux threading the loop, Φ0=h/2e\Phi_{0}=h/2e is the flux quantum, and kik_{i} (i=1,2,3i=1,2,3) are the scaling factors depending on the junction size. ECE_{\mathrm{C}} is the single-electron charging energy of the shunt capacitance. We assume that each Josephson energy EJiE_{\mathrm{J}i} scales as EJi=kiEJ0E_{\mathrm{J}i}=k_{i}E_{\mathrm{J0}}, where EJ0E_{\mathrm{J0}} is an independent parameter to be determined.

Refer to caption
Figure S2: Full-circuit model of a cubic transmon.
Refer to caption
Figure S3: Wave functions of the cubic-transmon eigenstates, obtained by diagonalizing the Hamiltonian [Eq. (S1)] with the parameters in the second line of Table S1, where a condition k2=k3k_{2}=k_{3} is assumed. Real parts of the wave functions for the ground, first-excited and second-excited states are plotted from left to right, respectively. The dashed lines represent the constraint 2ϕ2=ϕ+ϕ12\phi_{2}=\phi+\phi_{1}, which is used in the single-phase approximation.

By diagonalizing the Hamiltonian, we obtain wave functions of the eigenstates of the cubic transmon in the phase representation (Fig. S3). Under the condition of k2=k3k_{2}=k_{3}, the fringes of the excited states lie on the dashed lines indicating the relation 2ϕ2=ϕ+ϕ12\phi_{2}=\phi+\phi_{1}. The confinement of the wave functions along the dashed line suggests an approximation, 2ϕ2ϕ+ϕ12\phi_{2}\approx\phi+\phi_{1}, which we call the single-phase approximation. Using this relation, we can write the inductive energy of the SNAIL,

U(φ)\displaystyle U(\varphi) =\displaystyle= k1EJ0cosφ2k2EJ0cos(ϕφ2)\displaystyle-k_{1}E_{\mathrm{J0}}\cos\varphi-2k_{2}E_{\mathrm{J0}}\cos\left(\frac{\phi-\varphi}{2}\right) (S8)
=\displaystyle= D2δ2+D3δ3+D4δ4+O(δ5).\displaystyle D_{2}\delta^{2}+D_{3}\delta^{3}+D_{4}\delta^{4}+O(\delta^{5}).

This gives an effective model with a single phase degree of freedom, φ(ϕ1)\varphi\,(\equiv\phi_{1}). The Josephson energies in the main text are defined as EJ=k1EJ0E_{\mathrm{J}}^{\prime}=k_{1}E_{\mathrm{J0}} and EJ=k2EJ0E_{\mathrm{J}}=k_{2}E_{\mathrm{J0}}. The second formula in Eq. (S8) is the Taylor expansion around the phase φ0\varphi_{0} at a minimum of the inductive energy, where δφφ0\delta\equiv\varphi-\varphi_{0} is the relative phase variable for the expansion and DiD_{i} (i=2,3,4)(i=2,3,4) are the expansion coefficients. The parity symmetry δδ\delta\leftrightarrow-\delta is broken as seen in the existence of the δ3\delta^{3} term in the presence of a finite magnetic flux penetrating through the SNAIL loop.

Under the approximation, the Hamiltonian of the cubic transmon up to the third-order nonlinearity is, as in the main text,

H^/=\displaystyle\hat{H}/\hbar=\!\!\!\!\! ωc0a^a^+βc0(a^a^a^+a^a^a^)+αc02a^a^a^a^,\displaystyle\omega_{\mathrm{c0}}\hat{a}^{\dagger}\hat{a}+\beta_{\mathrm{c0}}(\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}+\hat{a}^{\dagger}\hat{a}\hat{a})+\frac{\alpha_{\mathrm{c0}}}{2}\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}\hat{a}, (S9)

where

ωc0\displaystyle\hbar\omega_{\mathrm{c0}} =\displaystyle= 16D2EC+12D4ECD2\displaystyle\sqrt{16D_{2}E_{\mathrm{C}}^{\prime}}+\frac{12D_{4}E_{\mathrm{C}}^{\prime}}{D_{2}} (S10)
βc0\displaystyle\hbar\beta_{\mathrm{c0}} =\displaystyle= 3(ECD2)3/4D3\displaystyle 3\left(\frac{E_{\mathrm{C}}^{\prime}}{D_{2}}\right)^{3/4}D_{3} (S11)
αc0\displaystyle\hbar\alpha_{\mathrm{c0}} =\displaystyle= 6D4ECD2.\displaystyle\frac{6D_{4}E_{\mathrm{C}}^{\prime}}{D_{2}}. (S12)

The effective charging energy ECE_{\mathrm{C}}^{\prime} is expressed as

ECECk2+k3k2+k3+k1(k2+k3)+k2k3.E_{\mathrm{C}}^{\prime}\equiv E_{\mathrm{C}}\frac{k_{2}+k_{3}}{k_{2}+k_{3}+k_{1}(k_{2}+k_{3})+k_{2}k_{3}}. (S13)
Refer to caption
Figure S4: Accuracy of the single-phase approximation. (a) Eigenfrequency ωc0\omega_{\mathrm{c0}}, (b) third-order nonlinearity αc0\alpha_{\mathrm{c0}} and (c) second-order nonlinearity |βc0||\beta_{\mathrm{c0}}| of a cubic transmon as a function of Φ\Phi. Black curves show the calculations based on the single-phase approximation with the parameters determined by the fittings in Figs. S5(a) and S5(b). Red curves are the simulation from the full-circuit model with the same parameters. Blue dashed curves are from the full-circuit model with adjusted parameters. The parameters are listed in Table S1.
Table S1: Parameters used for the calculations in Figs. S3 and S4. The values for the single-phase approximation are determined from the fittings in Figs. S5(a) and S5(b). Those for the full-circuit model are adjusted to obtain the blue dashed curves in Fig. S4, which closely reproduce the calculations based on the single-phase approximation.
EC/h(GHz)E_{\mathrm{C}}/h\mathrm{~{}(GHz)} EJ0/h(GHz)E_{\mathrm{J0}}/h\mathrm{~{}(GHz)} k1k_{1} k2(=k3)k_{2}(=k_{3})
Single-phase approximation 0.21 84 0.070 0.20
Full-circuit model 0.18 103 0.070 0.20

We quantitatively compare the single-phase approximation with the full-circuit model. We calculate the eigenmode frequency of the first excited state, ωc0\omega_{\mathrm{c0}}, third-order nonlinearity αc0\alpha_{\mathrm{c0}} and second-order nonlinearity βc0\beta_{\mathrm{c0}} of the cubic transmon based on each model (Fig. S4). For the calculation with the single-phase approximation, we use the parameters obtained by the fittings in Figs. S5(a) and (b) below. Next, we use the same parameters in the full-circuit model and compare the results (red lines). For the full-circuit model, the second-order nonlinearity βc0\beta_{\mathrm{c0}} is evaluated from the transition moment between the ground and second-excited states. The transition moment is defined as

Aij=|i|ni|j|,A_{ij}=|\langle i|n_{i}|j\rangle|, (S14)

where i,j{g,e,f}i,j\in\{g,e,f\}. We also obtain from the perturbative approach

Age\displaystyle A_{ge} \displaystyle\propto Ω0\displaystyle\Omega_{0} (S15)
Agf\displaystyle A_{gf} \displaystyle\propto 2βc0Ω0ωc0+αc0,\displaystyle\frac{2\beta_{\mathrm{c0}}\Omega_{0}}{\omega_{\mathrm{c0}}+\alpha_{\mathrm{c0}}}, (S16)

where Ω0\Omega_{0} is the external drive amplitude for these transition. Using these, we calculate the absolute value of βc0\beta_{\mathrm{c0}} as

|βc0|=ωc0+αc02AgfAge.|\beta_{\mathrm{c0}}|=\frac{\omega_{\mathrm{c0}}+\alpha_{\mathrm{c0}}}{2}\frac{A_{gf}}{A_{ge}}. (S17)

The calculations based on these models qualitatively agree with each other [Fig. S4(a)–(c)] and demonstrate the validity and accuracy of the single-phase approximation. There is a small quantitative deviation between the two models, which is not surprising as the wave functions of the eigenstates (Fig. S3) are not completely localized along the dashed line. This means that 2ϕ2=ϕ+ϕ12\phi_{2}=\phi+\phi_{1} is not strictly satisfied because of the quantum fluctuation of ϕ2\phi_{2}, and we cannot construct an exact single-phase model. Blue dashed curves in Fig. S4 show calculations based on the full-circuit model with adjusted parameters to reproduce the results of the single-phase approximation. For the region with small reduced magnetic flux, these calculations have a good agreement with each other.

III COUPLED QUBITS

Refer to caption
Figure S5: Calibration experiments of the cubic-transmon–transmon coupled system. (a) Eigenfrequency of the first excited state, ωc\omega_{\mathrm{c}}, (b) third-order nonlinearity αc\alpha_{\mathrm{c}}, and (c) strength of the residual ZZ interaction JZZJ_{\mathrm{ZZ}} as a function of the flux bias Φ\Phi. Red dots are the experimental data. (d) Coupling coefficient of the parametric drive, η\eta, calculated with Eq. (S36). Black curves in (a) and (b) are the fitting results using Eqs. (S25) and (S27). Black curves in (c) and (d) are calculated based on Eqs. (S31) and (S36), respectively, with the parameters determined by the fittings.

As described in the main text, the total Hamiltonian H^\hat{H} of the cubit-transmon–transmon coupled system is given as

H^\displaystyle\hat{H} =\displaystyle= H^0+V^,\displaystyle\hat{H}_{0}+\hat{V}, (S18)
H^0/\displaystyle\hat{H}_{0}/\hbar =\displaystyle= ωc0a^a^+αc02a^a^a^a^+ωt0b^b^+αt02b^b^b^b^,\displaystyle\omega_{\mathrm{c0}}\hat{a}^{\dagger}\hat{a}+\frac{\alpha_{\mathrm{c0}}}{2}\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}\hat{a}+\omega_{\mathrm{t0}}\hat{b}^{\dagger}\hat{b}+\frac{\alpha_{\mathrm{t0}}}{2}\hat{b}^{\dagger}\hat{b}^{\dagger}\hat{b}\hat{b}, (S19)
V^/\displaystyle\hat{V}/\hbar =\displaystyle= βc0(a^a^a^+a^a^a^)+g0(a^b^+a^b^).\displaystyle\beta_{\mathrm{c0}}(\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}+\hat{a}^{\dagger}\hat{a}\hat{a})+g_{0}(\hat{a}^{\dagger}\hat{b}+\hat{a}\hat{b}^{\dagger}). (S20)

The parameters are defined in the main text. We treat the off-diagonal part V^\hat{V} as a perturbative term and obtain the effective Hamiltonian via Schrieffer-Wolff transformation.

H^eS^H^eS^H^+[S^,H^]+12[S^,[S^,H^]].\hat{H}^{\prime}\equiv e^{\hat{S}}\hat{H}e^{-\hat{S}}\sim\hat{H}+[\hat{S},\hat{H}]+\frac{1}{2}[\hat{S},[\hat{S},\hat{H}]]. (S21)

We introduce S^1\hat{S}_{1} which fulfills

V^=[S^1,H^0].\hat{V}=-[\hat{S}_{1},\hat{H}_{0}]. (S22)

Then, the effective Hamiltonian in the second order reads

H^eff(2)=H^0+[S^1,V^]+12[S^1,[S^1,H^0]].\hat{H}_{\mathrm{eff}}^{(2)}=\hat{H}_{0}+[\hat{S}_{1},\hat{V}]+\frac{1}{2}[\hat{S}_{1},[\hat{S}_{1},\hat{H}_{0}]]. (S23)

We calculate the effective Hamiltonian by ignoring states with more than four excitation quanta in each qubit and truncating it into a matrix with 16×1616\times 16 elements for the two-qubit system. The calculation is valid when g0|ωc0ωt0|g_{0}\ll|\omega_{\mathrm{c0}}-\omega_{\mathrm{t0}}| and |βc0|ωc0|\beta_{\mathrm{c0}}|\ll\omega_{\mathrm{c0}} are satisfied. The effective Hamiltonian reads [Eq. (2) in the main text]

H^eff(2)/=[ωc+g(b^+b^)]a^a^+αc2a^a^a^a^+ωtb^b^+αt2b^b^b^b^,\displaystyle\hat{H}_{\mathrm{eff}}^{(2)}/\hbar=\left[\omega_{\mathrm{c}}+g(\hat{b}^{\dagger}+\hat{b})\right]\hat{a}^{\dagger}\hat{a}+\frac{\alpha_{\mathrm{c}}}{2}\hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}\hat{a}+\omega_{\mathrm{t}}\hat{b}^{\dagger}\hat{b}+\frac{\alpha_{\mathrm{t}}}{2}\hat{b}^{\dagger}\hat{b}^{\dagger}\hat{b}\hat{b}, (S24)

where ωc\omega_{\mathrm{c}}, ωt\omega_{\mathrm{t}}, αc\alpha_{\mathrm{c}}, and αt\alpha_{\mathrm{t}} are the eigenmode frequencies and self-Kerr nonlinearities of the qubits after the perturbative treatment of the coupling term. They are expressed as follows:

ωc\displaystyle\omega_{\mathrm{c}}\!\!\! =\displaystyle= ωc02βc02ωc0+αc0+g02Δ0,\displaystyle\!\!\!\omega_{\mathrm{c0}}-\frac{2\beta_{\mathrm{c0}}^{2}}{\omega_{\mathrm{c0}}+\alpha_{\mathrm{c0}}}+\frac{g_{0}^{2}}{\Delta_{0}}, (S25)
ωt\displaystyle\omega_{\mathrm{t}}\!\!\! =\displaystyle= ωt0g02Δ0,\displaystyle\!\!\!\omega_{\mathrm{t0}}-\frac{g_{0}^{2}}{\Delta_{0}}, (S26)
αc\displaystyle\alpha_{\mathrm{c}}\!\!\! =\displaystyle= αc06βc02ωc0(ωc0+αc0)(2αc0+ωc0)2g02αc0Δ0(αc0+Δ0),\displaystyle\!\!\!\alpha_{\mathrm{c0}}-\frac{6\beta_{\mathrm{c0}}^{2}\omega_{\mathrm{c0}}}{(\omega_{\mathrm{c0}}+\alpha_{\mathrm{c0}})(2\alpha_{\mathrm{c0}}+\omega_{\mathrm{c0}})}-\frac{2g_{0}^{2}\alpha_{\mathrm{c0}}}{\Delta_{0}(\alpha_{\mathrm{c0}}+\Delta_{0})}, (S27)
αt\displaystyle\alpha_{\mathrm{t}}\!\!\! =\displaystyle= αt0+2g02αt0Δ0(αt0Δ0),\displaystyle\!\!\!\alpha_{\mathrm{t0}}+\frac{2g_{0}^{2}\alpha_{\mathrm{t0}}}{\Delta_{0}(\alpha_{\mathrm{t0}}-\Delta_{0})}, (S28)

where Δ0ωt0ωc0\Delta_{0}\equiv\omega_{\mathrm{t0}}-\omega_{\mathrm{c0}} is the detuning between the qubit bare frequencies. We use these expressions to fit the experimental data, as shown in Figs. S5(a) and S5(b). The fitting parameters are listed in Table S1.

The term ga^a^(b^+b^)g\hat{a}^{\dagger}\hat{a}(\hat{b}^{\dagger}+\hat{b}) in the effective Hamiltonian [Eq. (S24)] is the parity-violating term, and the effective couping strength gg is expressed as

g=g0βc0(Δ0+ωc0+2αc0)(αc0+Δ0)(αc0+ωc0).g=-\frac{g_{0}\beta_{\mathrm{c0}}(\Delta_{0}+\omega_{\mathrm{c0}}+2\alpha_{\mathrm{c0}})}{(\alpha_{\mathrm{c0}}+\Delta_{0})(\alpha_{\mathrm{c0}}+\omega_{\mathrm{c0}})}. (S29)

The amplitude of the residual ZZ interaction, JZZJ_{\mathrm{ZZ}}, between the qubits is derived through Schrieffer-Wolff transformation up to the fourth order of g0g_{0},

H^ZZ\displaystyle\hat{H}_{\mathrm{ZZ}} =\displaystyle= JZZa^a^b^b^,\displaystyle J_{\mathrm{ZZ}}\hat{a}^{\dagger}\hat{a}\hat{b}^{\dagger}\hat{b}, (S30)
JZZ\displaystyle J_{\mathrm{ZZ}} =\displaystyle= 2g02(αc0+αt0)(αc0+Δ0)(αt0+Δ0).\displaystyle\frac{2g_{0}^{2}(\alpha_{\mathrm{c0}}+\alpha_{\mathrm{t0}})}{(\alpha_{\mathrm{c0}}+\Delta_{0})(-\alpha_{\mathrm{t0}}+\Delta_{0})}. (S31)

In Fig. S5(c), we plot JZZJ_{\mathrm{ZZ}} obtained with the parameters that are determined from the fittings in Figs. S5(a) and S5(b). In the dispersive regime of the two qubits, i.e., for Δ0g0\Delta_{0}\gg g_{0}, the experimental data in Fig. S5(c) has a good agreement with the theoretically expected values.

For the calculation of the parametric coupling, we continue this procedure one more step. We set S^2\hat{S}_{2} such that

V^2=[S^2,H^0],\hat{V}_{2}=-[\hat{S}_{2},\hat{H}_{0}], (S32)

where V^2\hat{V}_{2} is the off-diagonal part of the effective Hamiltonian H^eff(2)\hat{H}_{\mathrm{eff}}^{(2)}. We drive this system at the frequency ωd\omega_{\mathrm{d}} with a phase θ\theta, such that

H^d=Ω(eiωdtiθa^+eiωdt+iθa^),\hat{H}_{\mathrm{d}}=\Omega(e^{-i\omega_{\mathrm{d}}t-i\theta}\hat{a}+e^{i\omega_{\mathrm{d}}t+i\theta}\hat{a}^{\dagger}), (S33)

and transform the drive Hamiltonian as

H^dp=H^d+[S^1,H^d]+12[S^1,[S^1,H^d]]+[S^2,H^d],\hat{H}_{\mathrm{dp}}=\hat{H}_{\mathrm{d}}+[\hat{S}_{1},\hat{H}_{\mathrm{d}}]+\frac{1}{2}[\hat{S}_{1},[\hat{S}_{1},\hat{H}_{\mathrm{d}}]]+[\hat{S}_{2},\hat{H}_{\mathrm{d}}], (S34)

to obtain the parametric coupling of the SWAP interaction for ωd=Δωtωc\omega_{\mathrm{d}}=\Delta\equiv\omega_{\mathrm{t}}-\omega_{\mathrm{c}} [Eqs. (3) and (4) in the main text],

H^p/\displaystyle\hat{H}_{\mathrm{p}}/\hbar =\displaystyle= ηΩ(eiωdt+iθa^b^+eiωdtiθa^b^),\displaystyle\eta\Omega(e^{i\omega_{\mathrm{d}}t+i\theta}\hat{a}^{\dagger}\hat{b}+e^{-i\omega_{\mathrm{d}}t-i\theta}\hat{a}\hat{b}^{\dagger}), (S35)
η\displaystyle\eta \displaystyle\equiv 2g0βc0(2ωc02αc0Δ0+2αc0ωc0)Δ0(Δ0ωc0)(αc0+ωc0)(αc0+ωc0+Δ0).\displaystyle\frac{-2g_{0}\beta_{\mathrm{c0}}(2\omega_{\mathrm{c0}}^{2}-\alpha_{\mathrm{c0}}\Delta_{0}+2\alpha_{\mathrm{c0}}\omega_{\mathrm{c0}})}{\Delta_{0}(\Delta_{0}-\omega_{\mathrm{c0}})(\alpha_{\mathrm{c0}}+\omega_{\mathrm{c0}})(\alpha_{\mathrm{c0}}+\omega_{\mathrm{c0}}+\Delta_{0})}. (S36)

For the CZ gate, we use the transition at ωd=Δ+αt\omega_{\mathrm{d}}=\Delta+\alpha_{\mathrm{t}} involving the second-excited state of the transmon, whose amplitude is similarly obtained as

ηCZ\displaystyle\eta_{\mathrm{CZ}} \displaystyle\equiv 22g0βc0(2ωc02αc0Δ0+2αc0ωc0+αc0αt0)(Δ0αt0)(αt0Δ0+ωc0)(αc0+ωc0)(αc0αt0+ωc0+Δ0).\displaystyle\frac{2\sqrt{2}g_{0}\beta_{\mathrm{c0}}(2\omega_{\mathrm{c0}}^{2}-\alpha_{\mathrm{c0}}\Delta_{0}+2\alpha_{\mathrm{c0}}\omega_{\mathrm{c0}}+\alpha_{\mathrm{c0}}\alpha_{\mathrm{t0}})}{(\Delta_{0}-\alpha_{\mathrm{t0}})(\alpha_{\mathrm{t0}}-\Delta_{0}+\omega_{\mathrm{c0}})(\alpha_{\mathrm{c0}}+\omega_{\mathrm{c0}})(\alpha_{\mathrm{c0}}-\alpha_{\mathrm{t0}}+\omega_{\mathrm{c0}}+\Delta_{0})}. (S37)

IV PARAMETRICALLY-INDUCED transition

Refer to caption
Figure S6: Rabi oscillation of the |ge~|eg~\widetilde{|ge\rangle}\leftrightarrow\widetilde{|eg\rangle} transition. The vertical axis is the frequency of the parametric drive, ωd\omega_{\mathrm{d}}, and the horizontal axis is the interaction time. The color shows the normalized average quadrature amplitude σ¯z,t\bar{\sigma}_{z,\mathrm{t}} of the transmon readout signal.

Figure S6 shows the experimental data of the parametrically-induced |ge~|eg~\widetilde{|ge\rangle}\leftrightarrow\widetilde{|eg\rangle} transition. We prepare the |eg~\widetilde{|eg\rangle} state with a π\pi-pulse to the cubic transmon and apply the parametric drive to the cubic transmon. The excitation is swapped between the two states by the parametric transition. The resonance frequency, 846 MHz, is the frequency difference between eigenfrequencies of the cubic transmon and transmon. The Rabi frequency is proportional to the amplitude of the drive, and the maximum Rabi frequency of 30 MHz is obtained.

V Randomized Benchmarking

Refer to caption
Figure S7: Pulse sequences for RB. (a) and (b) Pulse sequences for single-qubit RB with the cubic transmon and transmon, respectively. An array of single-qubit random Clifford gates, C1,C2,,Cn1C_{1},C_{2},\ldots,C_{n-1}, is applied, and U1U^{-1} is the inverse of the preceding sequence. (c) Pulse sequence for two-qubit RB. An array of two-qubit random Clifford gates, B1,B2,,Bn1B_{1},B_{2},\ldots,B_{n-1}, is applied, followed by the inverse U1U^{-1}. (d) Decompositions of two-qubit Clifford gates.
Refer to caption
Figure S8: Randomized benchmarking of the single-qubit gates. The vertical axes show the normalized average quadrature amplitudes of the readout signals for (a) the cubic transmon and (b) the transmon. The horizontal axes show the number of Clifford gates applied in the randomized sequence. Curves are the fittings to the depolarization model.

Figure S7 shows the gate sequences for the single-qubit and two-qubit randomized benchmarking (RB). We drive the cubic transmon with a CW field to eliminate the static ZZ interaction (not shown). The pulse shapes for the single-qubit gates are Gaussian with a full width at half maximum of 18.6 ns. The swap pulse and each segment of the control-phase pulse [Fig. 3(a) in the main text] have rising and falling edges of a Gaussian shape with the half width at half maximum of 3.0 ns and 1.5 ns, respectively. The length of the flat-top region is 32 ns for the swap pulse and 16 ns for each segment of the control-phase pulse.

The tails for all pulse are truncated when the amplitudes become 10310^{-3} times smaller than the maximum. For the interleaved RB, we add a target gate (CZ, iSWAP and SWAP) between each Clifford gates. We repeat the sequences 5000 times with 100 (50) different random patterns for the protocol in Figs. S7(a) and S7(b) [Fig. S7(c)]. We measure the average value of the σz\sigma_{z}-component of the cubic transmon in the protocol in Fig. S7(a) and that of the transmon in Figs. S7(b) and S7(c).

Figure S8 shows the results of standard RB for the single-qubit gates. The average gate fidelities of the single-qubit gates are evaluated to be 0.963±0.0010.963\pm 0.001 for the cubic transmon and 0.977±0.0010.977\pm 0.001 for the transmon.

VI Raman transition through a CONTINUOUS MICROWAVE field

Refer to caption
Figure S9: Raman transition assisted by the CW drive. The vertical axis shows the normalized average quadrature amplitude of the readout signal for the cubic transmon. Horizontal axis is the frequency of the probe pulse, ωp\omega_{\mathrm{p}}, applied to the transmon, subtracted by the eigenfrequency of the transmon, ωt\omega_{\mathrm{t}}. Black dots and blue circles are the experimental data with and without the CW drive. The black curve fits the data with a Gaussian function, whose spectral width is one of the fitting parameters and consistent with the temporal shape of the probe pulse. Inset shows the energy-level diagram. Red and black arrows represent the CW-drive and probe-microwave frequencies, respectively.

In the main text, we irradiate the cubic transmon with a continuous microwave (CW) drive to eliminate the unwanted static ZZ interaction between the two qubits. However, this CW drive also induces an unwanted Raman transition which is mediated by the transmon excitation. Figure S9 shows the experimental data regarding the transition with pulsed spectroscopy. We sweep the frequency of the probe microwave pulse, ωp\omega_{\mathrm{p}}, around the transmon excitation frequency and measure the state of the cubic transmon. The pulse has a Gaussian shape with the full width at half maximum of 60 ns. The peak observed in Fig. S9 corresponds to the Raman transition process depicted in the inset. In accordance with the CW-drive detuning of 84 MHz from the |eg~|ge~\widetilde{|eg\rangle}\leftrightarrow\widetilde{|ge\rangle} transition, the Raman transition appears at (ωt/2π+84)(\omega_{\mathrm{t}}/2\pi+84) MHz. This transition is close to the transmon resonance and can be an error source for single-qubit gates with a short pulse. This error can be suppressed by the use of DRAG pulses DRAG .

References

  • (1) F. Motzoi, J. M. Gambetta, P. Rebentrost, and F. K. Wilhelm, Phys. Rev. Lett. 103, 110501 (2009).