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

Universal singlet-triplet qubits implemented near the transverse sweet spot

Wen-Xin Xie Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, and School of Physics
and Telecommunication Engineering, South China Normal University, Guangzhou 510006, China
   Chengxian Zhang [email protected] Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, and School of Physics
and Telecommunication Engineering, South China Normal University, Guangzhou 510006, China
Guangdong-Hong Kong Joint Laboratory of Quantum Matter and Frontier Research Institute for Physics,
South China Normal University, Guangzhou 510006, China
   Zheng-Yuan Xue [email protected] Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, and School of Physics
and Telecommunication Engineering, South China Normal University, Guangzhou 510006, China
Guangdong-Hong Kong Joint Laboratory of Quantum Matter and Frontier Research Institute for Physics,
South China Normal University, Guangzhou 510006, China
Abstract

The key to realizing fault-tolerant quantum computation for singlet-triplet (ST) qubits in semiconductor double quantum dot (DQD) is to operate both the single- and two-qubit gates with high fidelity. The feasible way includes operating the qubit near the transverse sweet spot (TSS) to reduce the leading order of the noise, as well as adopting the proper pulse sequences which are immune to noise. The single-qubit gates can be achieved by introducing an AC drive on the detuning near the TSS. The large dipole moment of the DQDs at the TSS has enabled strong coupling between the qubits and the cavity resonator, which leads to a two-qubit entangling gates. When operating in the proper region and applying modest pulse sequences, both single- and two-qubit gates are having fidelity higher than 99%. Our results suggest that taking advantage of the appropriate pulse sequences near the TSS can be effective to obtain high-fidelity ST qubits.

I Introduction

Quantum computing using the spin states of the electrons confined in the semiconductor quantum dots Loss and DiVincenzo (1998); Petta et al. (2005); Hanson et al. (2007); Zimmerman et al. (2014); Bermeister et al. (2014); Veldhorst et al. (2014, 2015); Takeda et al. (2016); Schreiber and Bluhm (2018); Watson et al. (2018); Yoneda et al. (2018); Chan et al. (2018); Sigillito et al. (2019); Crippa et al. (2019); Emerson (2019); Huang et al. (2019) is promising due to the long coherence time and the possibility for scalability Hanson et al. (2007). Singlet-triplet (ST) qubits in double quantum dot (DQD) Petta et al. (2005); Foletti et al. (2009); Wang et al. (2014); Shulman et al. (2014); Nichol et al. (2017); Zhang et al. (2017); Takeda et al. (2020); Cerfontaine et al. (2020) is particularly standing out of various proposals that are encoded using the spin states because of its all-electrical operation and strong exchange interaction. However, in the operation for the ST qubits, both single- and two-qubit gates are susceptible to the charge fluctuation (charge noise). Although recent experiments have shown that the single-qubit gate fidelity for the ST qubits can be higher than 99% Nichol et al. (2017); Cerfontaine et al. (2020); Takeda et al. (2020), the fidelity for the two-qubit gate reported until now has been lower than the requirement for the fault-tolerant quantum computing Shulman et al. (2012); Nichol et al. (2017). This motivates us to further search for useful methods to suppress the charge noise.

Recently, much works have focused on the symmetric operation of the sweet spot (SOP) Medford et al. (2013); Shim and Tahan (2016); Russ et al. (2016); Rančić and Burkard (2017); Zhang et al. (2018); Yang et al. (2019), which is far away from the charging transition regions, namely (1, 1) to (0, 2) or (2, 0), where the notation (nL,nR)(n_{L},n_{R}) denotes the number of electron in the left and right dot. By operating the qubit near the sweet spot, the derivative of the exchange energy with respect to detuning is minimized. Therefore, the sensitivity of the qubits to the leading order of charge noise is reduced. In the SOP region, the dipole moment is dominated by the longitudinal components, which can be used to construct the CPHASE two-qubit gate via the longitudinal coupling Harvey et al. (2018). On the other hand, the transverse dipole moment is very small and thus the transverse coupling is infeasible. A latest work Abadillo-Uriel et al. (2019) has suggested that there exists a family of the so-called transverse sweet spots (TSS). It is shown that even though TSS are far from the conventional SOP, they are still having the merit of the sweet spot as the same with the SOP. Near the TSS region, the transverse dipole moment is large while the longitudinal one is suppressed, which is contrary to the case for the SOP. Therefore, they have enabled strong transverse couplings. This is promising to design a two-qubit iSWAP gate by transversely coupling two qubits mediated by a cavity resonator, as shown in the recent works Blais et al. (2007); Srinivasa et al. (2016); Harvey et al. (2018); Abadillo-Uriel et al. (2019); Zhang et al. (2020a, b).

In this work, we have investigated the ST qubits under the realistic time-dependent noise environment in DQD. By introducing an AC-driven detuning near the TSS, one is able to achieve arbitrary single-qubit gates. In addition, we have studied the performance of the gates in two different regions, which corresponds to the TSS value of εSS>0\varepsilon_{\rm{SS}}>0 and εSS<0\varepsilon_{\rm{SS}}<0, respectively. Here, εSS\varepsilon_{\rm{SS}} is the detuning at TSS. When operating the gate in the region εSS>0\varepsilon_{\rm{SS}}>0, the leakage is minimized and the detuning noise remains only in the zz direction. We compare the naive (primitive) gate with several typical pulse sequences, namely, the non-cyclic geometric gate Liu et al. (2020), the cyclic geometric gate Zhao et al. (2017); Zhang et al. (2020a) and the CORPSE gate Cummins et al. (2003); Bando et al. (2013), which are well-known to suppress the zz-component noise. We find that the gate performance under the time-dependent noise for each type of gate has strong connection with the noise spectrum. Whether the naive gates can be improved sensitively depends on the chosen rotations. On the other hand, we have also found that using the strong dipole coupling between the DQDs and the cavity resonator, a two-qubit entangling gate can be achieved. In our work, the fidelity for both single- and two-qubit gates are surpassing 99%.

II Model

Refer to caption
Figure 1: (a) The evolution path of state |ψ+\left|{{\psi_{+}}}\right\rangle, along the non-cyclic path A-B-C to construct the non-cyclic geometric gate. (b) The evolution path of state |ψ+\left|{{\psi_{+}}}\right\rangle, along the cyclic path D-E-F-D so as to obtain the geometric gate. (c) Dipole coupling of two DQDs to the cavity resonator. For both single- and two-qubit gates, the detuning is operated at the TSS denoted as εSS\varepsilon_{\rm{SS}}.
Refer to caption
Figure 2: (a) and (b): Energy spectrum of the DQD Hamiltonian HSTH_{\rm{ST}} in Eq. 1 as a function of the detuning value ε\varepsilon considering the cases ΔB<τ\Delta B<\tau (a) and ΔB>τ\Delta B>\tau (b). (c) and (d): The corresponding qubit energy ωq=EeEg\omega_{q}=E_{e}-E_{g} is drawn from (a) and (b), respectively. When ΔB<τ\Delta B<\tau, the TSS point appears at ε<0\varepsilon<0. For comparison, it appears at ε>0\varepsilon>0 when ΔB>τ\Delta B>\tau. Parameters for (a) and (c): ΔB/2π=1.5GHz\Delta B/2\pi=1.5\ \rm{GHz}, τ/2π=1.75GHz\tau/2\pi=1.75\ \rm{GHz} while for (b) and (d): ΔB/2π=2.5GHz\Delta B/2\pi=2.5\ \rm{GHz}, τ/2π=1.5GHz\tau/2\pi=1.5\ \rm{GHz}.

The Hamiltonian for the ST qubits in the basis states spanned by {|T0(1,1),|S(1,1),|S(0,2)}\left\{\left|T_{0}(1,1)\right\rangle,|S(1,1)\rangle,|S(0,2)\rangle\right\} is given by Abadillo-Uriel et al. (2019)

HST=(0ΔB0ΔB02τ02τε).H_{\rm{ST}}=\left(\begin{array}[]{ccc}0&\Delta B&0\\ \Delta B&0&\sqrt{2}\tau\\ 0&\sqrt{2}\tau&-\varepsilon\end{array}\right). (1)

Here, the spin state is defined as |,=cLcR|𝒱\left|\uparrow,\downarrow\right\rangle=c_{L\uparrow}^{\dagger}c_{R\downarrow}^{\dagger}|\mathcal{V}\rangle where cjμc_{j\mu}^{\dagger} creates an electron with spin μ\mu at the jjth quantum dot and 𝒱\mathcal{V} denotes vacuum Wang et al. (2014). |S(1,1)=(|,|,)/2|S(1,1)\rangle=(\left|\uparrow,\downarrow\right\rangle-\left|\downarrow,\uparrow\right\rangle)/\sqrt{2} and |T0(1,1)=(|,+|,)/2|T_{0}(1,1)\rangle=(\left|\uparrow,\downarrow\right\rangle+\left|\downarrow,\uparrow\right\rangle)/\sqrt{2} are the spin singlet and triplet state, respectively. While |S(0,2)|S(0,2)\rangle is another singlet state when both of the two electrons occupy the right dot. The parameter ΔB\Delta B represents the magnetic field gradient between the two quantum dots, while τ\tau and ε\varepsilon are the tunneling and the detuning values between the two dots (see Fig. 1(c)). Note that the parameters here are in energy units by taking =1\hbar=1. Also, we assume the global Zeeman field is large enough, therefore, the other two triplet states |,\left|\uparrow,\uparrow\right\rangle and |,\left|\downarrow,\downarrow\right\rangle are split off and thus can be ignored.

We first show how to seek for the proper TSS detuning point εSS\varepsilon_{\rm{SS}} to define the qubit subspace. As shown in Fig. 2, the energy structure of the system is plotted by diagonalizing HSTH_{\rm{ST}}, where the eigenbasis is denoted as {|g,|e,|f}\{|g\rangle,|e\rangle,|f\rangle\}. The qubit states are defined as the two lowest eigenbasis states, namely {|0=|e,|1=|g}\left\{|0\rangle=|e\rangle,|1\rangle=|g\rangle\right\}. Introducing an AC drive ε(t)=εACcos(ωt+ϕ(t))\varepsilon^{\prime}(t)=\varepsilon_{\rm{AC}}\cos(\omega t+\phi(t)) to the detuning parameter oscillating near the given εSS\varepsilon_{\rm{SS}}, namely, ε=εSS+ε\varepsilon=\varepsilon_{\rm{SS}}+\varepsilon^{\prime}, the total Hamiltonian of the qubit system is therefore

Hq=n=13En(εSS)σnn+Hint,H_{q}=\sum_{n=1}^{3}E_{n}(\varepsilon_{\rm{SS}})\sigma_{nn}+H_{\rm{int}}, (2)

with

Hint=ε|S(0,2)S(0,2)|=εm,ndmnσmn..\begin{aligned} H_{\rm{int}}=-\varepsilon^{\prime}|S(0,2)\rangle\langle S(0,2)|=\varepsilon^{\prime}\sum_{m,n}d_{mn}\sigma_{mn}.\end{aligned}. (3)

Here, En(εSS)E_{n}(\varepsilon_{\rm{SS}}) (the numbers 1, 2, and 3 represent states |g|g\rangle, |e|e\rangle and |f|f\rangle) is the eigenvalue of HSTH_{\rm{ST}} which is dependent on the chosen εSS\varepsilon_{\rm{SS}}, σmn=|mn|\sigma_{mn}=|m\rangle\langle n| with m,nm,n being the eigenstates, and dmn=m|d^|nd_{mn}=\langle m|\hat{d}|n\rangle is the dipole matrix element with d^=HST/ε\hat{d}=\partial H_{\mathrm{ST}}/\partial\varepsilon. When the AC drive is in resonance with the qubit energy, i.e. ωq=EeEg=ω\omega_{q}=E_{e}-E_{g}=\omega, the system under the rotating wave approximation (RWA) in the rotating frame can be well described by an effective two-level structure (the detail can be seen in Appendix. A):

HAC=Ω02(cosϕ(t)σx+sinϕ(t)σy),H_{\rm{AC}}=\frac{\Omega_{0}}{2}\left(\cos\phi(t)\ \sigma_{x}+\sin\phi(t)\ \sigma_{y}\right), (4)

Here, the Rabi frequency is Ω0=|dge|εAC\Omega_{0}=|d_{ge}|\varepsilon_{\rm{AC}}. Then, using this piecewise continuous Hamiltonian one is able to realize arbitrary rotation around the axis on the xx-yy plane, denoted as R(𝒓,θ)R(\boldsymbol{r},\theta) where 𝒓\boldsymbol{r} is the rotation axis and θ\theta the rotation angle. The rotation suffering noise is as described in Appendix. B. Further, Any single-qubit gate out of the xx-yy plane can be decomposed into a related “xx-yy-xx” (or “yy-xx-yy”) sequence Nielsen and Chuang (2002), which we call as “naive pulse sequence”.

Refer to caption
Figure 3: The derivative ωq/ε\partial{\omega_{q}}/\partial{\varepsilon} around the TSS. Parameters for (a): ΔB/2π=1.5GHz\Delta B/2\pi=1.5\ \rm{GHz}, τ/2π=1.75GHz\tau/2\pi=1.75\ \rm{GHz} while for (b): ΔB/2π=2.5GHz\Delta B/2\pi=2.5\ \rm{GHz}, τ/2π=1.5GHz\tau/2\pi=1.5\ \rm{GHz}.

Normally, HintH_{\rm{int}} can induce the transverse and longitudinal coupling of the qubit. At the TSS, |dge||d_{ge}| is closed to the maximum value (about 0.5) and the longitudinal component is small which can be further suppressed because it appears as the fast-oscillating terms. Therefore, it is wise to operate the qubit near the TSS. On the other hand, the TSS can be found in two different regions, i.e. ΔB<τ\Delta B<\tau and ΔB>τ\Delta B>\tau. When ΔB<τ\Delta B<\tau, the corresponding detuning value of the TSS is negative, i.e. εSS<0\varepsilon_{\rm{SS}}<0 (see Fig. 2(a) and (c)). In this case, we have ωef=EfEeωq\omega_{ef}=E_{f}-E_{e}\simeq\omega_{q} and it is easy to introduce leakage to the excited state |f|f\rangle. The mechanism of the leakage effect is described in Appendix. A. For comparison, when ΔB>τ\Delta B>\tau, we have εSS>0\varepsilon_{\rm{SS}}>0 (see Fig. 2(b) and (d)), where ωef\omega_{ef} is substantially larger than ωq\omega_{q} and therefore the leakage is suppressed. To strongly suppress the noise effect, we consider implementing geometric gate in the region ΔB>τ\Delta B>\tau and using εAC/2π=0.1GHz\varepsilon_{\rm{AC}}/2\pi=0.1\ \rm{GHz} in this work, which is reasonable in the experiment Kim et al. (2015). Hereafter, we stick to the parameters: ΔB/2π=2.5GHz\Delta B/2\pi=2.5\ \rm{GHz}, τ/2π=1.5GHz\tau/2\pi=1.5\ \rm{GHz}. The corresponding TSS is εSS/2π=1.91935GHz\varepsilon_{\rm{SS}}/2\pi=1.91935\ \rm{GHz} and the dipole element is |dge|=0.45055|d_{ge}|=0.45055. Considering εAC/2π=0.1GHz\varepsilon_{\rm{AC}}/2\pi=0.1\ \rm{GHz}, the Rabi frequency is thus Ω0/2π=0.045055GHz\Omega_{0}/2\pi=0.045055\ \rm{GHz}.

III Results

III.1 Single-qubit gates

Here, we focus on introducing how to design the non-cyclic geometric gate using the piecewise continuous Hamiltonian Liu et al. (2020) (the other types of pulse sequences is introduced in Appendix. B), and then we will discuss the noise effect caused by the charge noise in the DQD.

The quantum gate considered here is using two-pieces of Hamiltonian in Eq. 4. In each of the piece, the Hamiltonian is with the form as

Hn(t)={Ω02(cos(ϕ0+π2)σx+sin(ϕ0+π2)σy),TAt<TBΩ02(cos(ϕ0+ϕ1+π2)σx+sin(ϕ0+ϕ1+π2)σy),TB<tTCH_{n}(t)=\left\{\begin{array}[]{ll}\frac{\Omega_{0}}{2}(\cos(\phi_{0}+\frac{\pi}{2})\ \sigma_{x}+\sin(\phi_{0}+\frac{\pi}{2})\ \sigma_{y}),&T_{A}\leqslant t<T_{B}\\ \frac{\Omega_{0}}{2}(\cos(\phi_{0}+\phi_{1}+\frac{\pi}{2})\ \sigma_{x}+\sin(\phi_{0}+\phi_{1}+\frac{\pi}{2})\ \sigma_{y}),&T_{B}<t\leqslant T_{C}\end{array}\right. (5)

where TBTA=χ0/Ω0T_{B}-T_{A}=\chi_{0}/\Omega_{0} and TCTB=β0/Ω0T_{C}-T_{B}=\beta_{0}/\Omega_{0}. Then, this two-connecting evolution together at the final time forms a desired quantum gate as

Un(χ0,ϕ0,ϕ1,β0)\displaystyle{U_{n}}\left({{\chi_{0}},{\phi_{0}},{\phi_{\rm{1}}},{\beta_{0}}}\right) =UCB(TC,TB)UBA(TB,TA)\displaystyle={U_{CB}}\left({{T_{C}},{T_{B}}}\right){U_{BA}}\left({{T_{B}},T_{A}}\right) (6)
=(cosχ02cosβ02sinχ02sinβ02eiϕ1cosχ02sinβ02ei(ϕ0+ϕ1)cosβ02sinχ02eiϕ0cosχ02sinβ02ei(ϕ0+ϕ1)+cosβ02sinχ02eiϕ0cosχ02cosβ02sinχ02sinβ02eiϕ1)\displaystyle=\left(\begin{array}[]{ccc}\cos\frac{\chi_{0}}{2}\cos\frac{\beta_{0}}{2}-\sin\frac{\chi_{0}}{2}\sin\frac{\beta_{0}}{2}e^{-i\phi_{1}}&-\cos\frac{\chi_{0}}{2}\sin\frac{\beta_{0}}{2}e^{-i\left(\phi_{0}+\phi_{1}\right)}-\cos\frac{\beta_{0}}{2}\sin\frac{\chi_{0}}{2}e^{-i\phi_{0}}\\ \cos\frac{\chi_{0}}{2}\sin\frac{\beta_{0}}{2}e^{i\left(\phi_{0}+\phi_{1}\right)}+\cos\frac{\beta_{0}}{2}\sin\frac{\chi_{0}}{2}e^{i\phi_{0}}&\cos\frac{\chi_{0}}{2}\cos\frac{\beta_{0}}{2}-\sin\frac{\chi_{0}}{2}\sin\frac{\beta_{0}}{2}e^{i\phi_{1}}\end{array}\right)

Here, χ0\chi_{0}, ϕ0\phi_{0}, ϕ1\phi_{1} and β0\beta_{0} are determined by the chosen evolution operator. In our simulation, the specific parameters are present in Appendix. D. Actually, Un(χ0,ϕ0,ϕ1,β0){U_{n}}\left({{\chi_{0}},{\phi_{\rm{0}}},{\phi_{\rm{1}}},{\beta_{0}}}\right) can enable arbitrary rotation on the Bloch sphere. For example, Un(χ0,π2,π,β0){U_{n}}({\chi_{0}},\frac{\pi}{2},\pi,{\beta_{0}}) and Un(χ0,π,π,β0){U_{n}}({\chi_{0}},-\pi,\pi,{\beta_{0}}) can implement a rotation around xx axis and yy axis, respectively, the rotation angle for both of which have the same form as β0χ0\beta_{0}-\chi_{0}. Note that, for a desired rotation around the xx or yy axis, even though one has many choices to pick up χ0\chi_{0} and β0\beta_{0}, the conditions for them are clear: χ0>0\chi_{0}>0 and β0>0\beta_{0}>0. In this way, the evolution time is positive and it ensures the gate will not degenerate to the naive pulse sequence.

The evolution of the quantum gate can be visualized by using the orthogonal dressed states

|ψ+(t)=cosχ2ei2η|0+sinχ2ei2η|1\displaystyle\left|\psi_{+}(t)\right\rangle=\cos\frac{\chi}{2}e^{-\frac{i}{2}\eta}|0\rangle+\sin\frac{\chi}{2}e^{\frac{i}{2}\eta}|1\rangle (7)
|ψ(t)=sinχ2ei2η|0cosχ2ei2η|1\displaystyle\left|\psi_{-}(t)\right\rangle=\sin\frac{\chi}{2}e^{-\frac{i}{2}\eta}|0\rangle-\cos\frac{\chi}{2}e^{\frac{i}{2}\eta}|1\rangle

As shown in Fig. 1(a), the whole evolution path of the quantum gate is divided into two proper parts, which are denoted as path AB and BC. In the first part of the evolution, |ψ+\left|\psi_{+}\right\rangle starts from a given point A, and evolves along the geodesic line up to the north pole B. Then, in the second part, it goes down to a given point C along another geodesic line. Using this dressed states basis, the evolution operator Un(χ0,ϕ0,ϕ1,β0){U_{n}}\left({{\chi_{0}},{\phi_{\rm{0}}},{\phi_{\rm{1}}},{\beta_{0}}}\right) can also be written as

Un=eiγ|ψ+(TC)ψ+(TA)|+eiγ|ψ(TC)ψ(TA)|\displaystyle{U_{n}}={e^{i{\gamma}}}\left|{{\psi_{+}}({T_{C}})}\right\rangle\left\langle{{\psi_{+}}({T_{A}})}\right|+{e^{-i{\gamma}}}\left|{{\psi_{-}}({T_{C}})}\right\rangle\left\langle{{\psi_{-}}({T_{A}})}\right| (8)

where γ\gamma is the related phase that is obtained. This two-piece path in the dressed state basis is not closed. On the other hand, the evolution path in the dressed state basis can alternatively be cyclic, as shown in Fig. 1(b) which can form a geometric gate. The detail of the geometric gate is as shown in Ref. Zhao et al. (2017). For this geometric gate, its whole evolution time is fixed to be 2π/Ω02\pi/\Omega_{0} Zhang et al. (2020c). By carefully selecting proper χ0\chi_{0} and β0\beta_{0} so as to satisfy χ0+β0<2π\chi_{0}+\beta_{0}<2\pi, the evolution time for the non-cyclic geometric gate can be always shorter than the geometric ones. Note that, in Ref.Liu et al. (2020), the authors have pointed out that the non-cyclic geometric gate introduced here is non-cyclic and non-Abelian. However, for the case mnm\neq n with m,n+,m,n\in+,-, one finds 0τψm(t)|(t)|ψn(t)𝑑t0\int_{0}^{\tau}\left\langle\psi_{m}(t)|\mathcal{H}(t)|\psi_{n}(t)\right\rangle dt\neq 0, namely, the dynamical phase in this case is present. Therefore, it does not satisfy the parallel transport condition and cannot be the non-Abelian geometric gate Kult et al. (2006).

Then, we study the noise effect of this AC-driven system to the non-cyclic geometric gate. Considering the charge noise effect, the detuning value turns to be ε(t)=εSS+εACcos(ωt+ϕ(t))+δε(t)\varepsilon(t)=\varepsilon_{\rm{SS}}+\varepsilon_{\rm{AC}}\cos(\omega t+\phi(t))+\delta\varepsilon(t), where δε(t)\delta\varepsilon(t) is the time-dependent charge noise. δε(t)\delta\varepsilon(t) results in the error for the chosen εSS\varepsilon_{\rm{SS}}, i.e. εSSεSS+δε(t)\varepsilon_{\rm{SS}}\rightarrow\varepsilon_{\rm{SS}}+\delta\varepsilon(t) and further induce the drift for the energy-level structure, namely, EnEn+δEnE_{n}\rightarrow E_{n}+\delta E_{n}. Assuming that εAC\varepsilon_{\rm{AC}}, δε(t)εSS\delta\varepsilon(t)\ll\varepsilon_{\rm{SS}}, we can expand the qubit energy as

ωqωq(εSS)+δωq,\displaystyle\omega_{q}\approx\omega_{q}\left(\varepsilon_{\rm{SS}}\right)+\delta\omega_{q}, (9)

where we have δωq(ωq/ε)δε\delta\omega_{q}\simeq(\partial{\omega_{q}}/\partial{\varepsilon})\ \delta\varepsilon when operating near the TSS point. This in turn leads to error for HACH_{\rm{AC}} in the rotating frame (see Appendix. A):

HAC=Ω02(cosϕ(t)σx+sinϕ(t)σy)+δωq/2σz.H_{\rm{AC}}^{\prime}=\frac{\Omega_{0}}{2}\left(\cos\phi(t)\ \sigma_{x}+\sin\phi(t)\ \sigma_{y}\right)+\delta\omega_{q}/2\ \sigma_{z}. (10)

Note that here we have assumed δε\delta\varepsilon is a constant value for simplicity and its time-dependent effect will be considered later. At the TSS point the qubit is first-order insensitive to the charge noise because of ωq/ε=0\partial{\omega_{q}}/\partial{\varepsilon}=0. However, when it is away from the TSS point, this first-order charge noise effect cannot be ignored anymore. In Fig. 3, we show the derivative ωq/ε\partial{\omega_{q}}/\partial{\varepsilon} as a function of the detuning value ε\varepsilon around the TSS point. We find that the derivative in Fig. 3(a), which corresponds to ΔB<τ\Delta B<\tau, is much smaller than that one when ΔB>τ\Delta B>\tau, as shown in Fig. 3(b). This means the qubit is more insensitive to the charge noise when operating in this region. However, as stated above, the leakage effect in this region is substantially large regardless of this superiority to the charge noise. Meanwhile, when ΔB>τ\Delta B>\tau, ωq/ε\partial{\omega_{q}}/\partial{\varepsilon} grows almost linearly with εεSS\varepsilon-\varepsilon_{\rm{SS}}. This implies that small AC drive is more appropriate in order to avoid large charge noise. As stated above, we have taken εAC/2π=0.1GHz\varepsilon_{\rm{AC}}/2\pi=0.1\ \rm{GHz}. Considering all the operating region of the detuning value ε\varepsilon we have ωq/εωq/ε|ε=εAC0.02\partial{\omega_{q}}/\partial{\varepsilon}\leq\partial{\omega_{q}}/\partial{\varepsilon}|_{\varepsilon=\varepsilon_{\rm{AC}}}\simeq 0.02. According to Ref. Mielke et al. (2020), the standard deviation related to the detuning charge noise is with the order of σε=1μeV\sigma_{\varepsilon}=1\ \mu\rm{eV}. Thus, the deviation with respect to the qubit energy can be 0<σωq(ωq/ε)σε0.02μeV0<\sigma_{\omega q}\leq(\partial{\omega_{q}}/\partial{\varepsilon})\ \sigma_{\varepsilon}\simeq 0.02\ \mu\rm{eV}. In our simulation we have used σωq=0.02μeV\sigma_{\omega q}=0.02\ \mu\rm{eV}.

Note that, for the ST qubits in semiconductor quantum dot, there are several extra noise channels that should be carefully treated. For example, the charge noise can also bring fluctuation for the tunneling value τ\tau, leading to tunneling noise. However, the magnitude of tunneling noise is much weaker than that of the detuning noise Koski et al. (2020). Here, the error in the AC-driven field εAC\varepsilon_{\rm{AC}} is neglected, which is also much weaker than the detuning noise Abadillo-Uriel et al. (2019). In addition, the fluctuation in the Overhauser field (nuclear noise) in GaAs can lead to unwanted error in ΔB\Delta B and also introduce relaxation Reilly et al. (2008). To deal with the nuclear noise, the silicon-based semiconductor quantum dot with isotopic purification is appreciated, where the nuclear noise is strongly suppressed Huang et al. (2019). Meanwhile, the valley in Si/SiGe platforms may introduce extra valley-spin coupling leading to relaxation Zhang et al. (2019). Fortunately, recent experiment indicated that by using the silicon metal-oxide-semiconductor (Si-MOS) platform the valley splitting can be as high as 0.8 meV Kawakami et al. (2014); Rančić and Burkard (2016). In this way, the valley effect can also be safely neglected. According to the latest experiment, the relaxation time based on silicon platform has reached T1=9sT_{1}=9\ \rm{s} Ciriano-Tejel et al. (2021). In this work, we focus on the suppression on δωq\delta\omega_{q} due to the detuning noise, and we simulate the single-qubit gate fidelity by using the two-level structure as shown in Eq. 10.

Refer to caption
Figure 4: Filter transfer function used to calculate individual gate fidelity. The gate in each panel is (a)R(x^,π/2)R(\hat{x},\pi/2), (b)R(z^,π/2)R(\hat{z},\pi/2), (c)R(x^+y^z^,4π/3)R(\hat{x}+\hat{y}-\hat{z},4\pi/3) and (d)R(x^+z^,π)R(\hat{x}+\hat{z},\pi). The design of these gates can be seen in Table. 1. The fidelity for the naive, the CORPSE, the geometric and the quantum gates are: (a) nai=99.78%\mathcal{F}_{\rm{nai}}=99.78\%, cor=99.61%\mathcal{F}_{\rm{cor}}=99.61\%, geo=99.59%\mathcal{F}_{\rm{geo}}=99.59\%, non=99.60%\mathcal{F}_{\rm{non}}=99.60\%. (b) nai=98.86%\mathcal{F}_{\rm{nai}}=98.86\%, cor=98.71%\mathcal{F}_{\rm{cor}}=98.71\%, geo=99.47%\mathcal{F}_{\rm{geo}}=99.47\%, non=99.47%\mathcal{F}_{\rm{non}}=99.47\%. (c) nai=99.39%\mathcal{F}_{\rm{nai}}=99.39\%, cor=99.16%\mathcal{F}_{\rm{cor}}=99.16\%, geo=98.89%\mathcal{F}_{\rm{geo}}=98.89\%, non=99.47%\mathcal{F}_{\rm{non}}=99.47\%. (d) nai=99.29%\mathcal{F}_{\rm{nai}}=99.29\%, cor=99.01%\mathcal{F}_{\rm{cor}}=99.01\%, geo=99.18%\mathcal{F}_{\rm{geo}}=99.18\%, non=99.29%\mathcal{F}_{\rm{non}}=99.29\%. Other parameters: Ω0/2π=0.045055GHz\Omega_{0}/2\pi=0.045055\ \rm{GHz}, S(ω)=A/(ωt0)αS(\omega)=A/(\omega t_{0})^{\alpha} with α=1\alpha=1 and At0=103At_{0}=10^{-3}. The cutoffs are ωir/2π=100kHz\omega_{\mathrm{ir}}/2\pi=100\ \mathrm{kHz} and ωuv/2π=20GHz\omega_{\mathrm{uv}}/2\pi=20\ \mathrm{GHz}.

In the real evolution process, the noise is time-dependent and is always described by the so-called 1/fα1/f^{\alpha} noise model. The power spectral density of the 1/fα1/f^{\alpha}-type noise has the form as S(ω)=A/(ωt0)αS(\omega)=A/(\omega t_{0})^{\alpha} Yang and Wang (2016). Here, AA denotes the noise amplitude, t0t_{0} the time unit, and the exponent α\alpha the noise correlation. For the semiconductor quantum-dot environment, the charge noise is typically about α1\alpha\simeq 1. The noise amplitude can be determined by Zhang et al. (2017)

ωirωuvAε(ωt0)α𝑑ω=πσωq2.\displaystyle\int_{\omega_{\mathrm{ir}}}^{\omega_{\mathrm{uv}}}\frac{A_{\varepsilon}}{\left(\omega t_{0}\right)^{\alpha}}d\omega=\pi\sigma_{\omega q}^{2}. (11)

Here, we take the cutoffs as ωir/2π=100kHz\omega_{\mathrm{ir}}/2\pi=100\ \mathrm{kHz} and ωuv/2π=20GHz\omega_{\mathrm{uv}}/2\pi=20\ \mathrm{GHz} Abadillo-Uriel et al. (2019). In our simulation, we take the time unit as t0=1/Ω0t_{0}=1/\Omega_{0}. Therefore, by substituting α=1\alpha=1, we estimate Aεt0103A_{\varepsilon}t_{0}\simeq 10^{-3}. Here, the 1/f1/f noise in the simulation is generated by using the method as described in Yang and Wang (2016). On the other hand, for the piecewise continuous Hamiltonian (see Eq. 4), the time-dependent noise spectrum and the infidelity can be well characterized via the filter transfer function Green et al. (2012, 2013); Paz-Silva and Viola (2014). The detail on how to analyze the fidelity using this method can be seen in Ref. Green et al. (2013). Here, we briefly introduce this method and focus on the zz-component noise. For the desired gate rotation U(t)U(t) which satisfies the Schrodinger equation iU˙(t)=H(t)U(t)i\dot{U}(t)=H(t)U(t), one can define a specific control matrix Green et al. (2013)

Rij(t)=[𝑹(t)]ij=Tr[U(t)σiU(t)σj]/2,\displaystyle R_{ij}(t)=[\boldsymbol{R}(t)]_{ij}=\operatorname{Tr}[U^{\dagger}(t)\sigma_{i}U(t)\sigma_{j}]/2, (12)

with its Fourier transform as

Rij(ω)=iω0T𝑑tRij(t)eiωt,\displaystyle R_{ij}(\omega)=-i\omega\int_{0}^{T}dtR_{ij}(t)e^{i\omega t}, (13)

where i,j{x,y,z}i,j\in\{x,y,z\}. And then, the average gate fidelity in the noise environment that is determined by the specific cross-spectral density matrix Sij(ω)S_{ij}(\omega) can be Green et al. (2013)

av\displaystyle\mathcal{F}_{\mathrm{av}}\simeq 112πi,j,k=x,y,zdωω2Sij(ω)Rjk(ω)Rik(ω).\displaystyle 1-\frac{1}{2\pi}\sum_{i,j,k=x,y,z}\int_{-\infty}^{\infty}\frac{\mathrm{d}\omega}{\omega^{2}}S_{ij}(\omega)R_{jk}(\omega)R_{ik}^{*}(\omega). (14)

Since we only consider the zz-component noise, the cross-spectral density is with the form of Sz(ω)S_{z}(\omega). In this way, Eq. 14 reduces to

av=\displaystyle\mathcal{F}_{\mathrm{av}}= 112π𝑑ωSz(ω)Fz(ω)ω2\displaystyle 1-\frac{1}{2\pi}\int_{-\infty}^{\infty}d\omega S_{z}(\omega)\frac{F_{z}(\omega)}{\omega^{2}} (15)

where

Fz(ω)=k=x,y,zRzk(ω)Rzk(ω)\displaystyle F_{z}(\omega)=\sum_{k=x,y,z}R_{zk}(\omega)R_{zk}^{*}(\omega) (16)

is the filter transfer function for the zz-component detuning noise. In Appendix. C, we have given a step-by-step example to calculate the filter transfer function. One can then use it to calculate the fidelity.

Considering the 1/f1/f noise spectrum, one can further derive the fidelity as

av=112πωirωuvAϵωt0Fz(ω)ω2\displaystyle\mathcal{F}_{\mathrm{av}}=1-\frac{1}{2\pi}\int_{\omega_{\mathrm{ir}}}^{\omega_{\mathrm{uv}}}\frac{A_{\epsilon}}{\omega t_{0}}\frac{F_{z}(\omega)}{\omega^{2}} (17)

Note that, to simplify the calculation one can use the unit of Ω0\Omega_{0} and the corresponding cutoffs turn to be ωir=ωir/Ω02.220×103\omega_{\mathrm{ir}}^{\prime}=\omega_{\mathrm{ir}}/\Omega_{0}\simeq 2.220\times 10^{-3} and ωuv=ωuv/Ω04.439×102\omega_{\mathrm{uv}}^{\prime}=\omega_{\mathrm{uv}}/\Omega_{0}\simeq 4.439\times 10^{2}.

From Eq. 15 it is clear that for a given noise spectrum the related filter transfer function i.e. Fz(ω)/ω2F_{z}(\omega)/\omega^{2} has positive relationship with the infidelity. Therefore, it can be used to characterize the noise effect. Fig. 4 shows the results of the filter transfer function for the naive, the CORPSE, the geometric and the non-cyclic geometric quantum gates. In the plot, we show four different kinds of gates (R(x^,π/2)R(\hat{x},\pi/2), R(z^,π/2)R(\hat{z},\pi/2), R(x^+y^z^,4π/3)R(\hat{x}+\hat{y}-\hat{z},4\pi/3) and R(x^+z^,π)R(\hat{x}+\hat{z},\pi)), which are the representatives for the rotation around different axis in the single-qubit Clifford group. The design of these gates can be seen in Table. 1. The performance of other gates in the Clifford group is similar and is thus not shown. We find that in the low-frequency region i.e. ω/Ω0<101\omega/\Omega_{0}<10^{-1}, the filter transfer function for the CORPSE gate is normally lower than others. This means that the CORPSE gate is more robust to the low-frequency noise. However, it stands out in the high-frequency region 101<ω/Ω0<10010^{-1}<\omega/\Omega_{0}<10^{0}. When the frequency is high enough, the performance for all the gates are not distinguishable. Using these filter transfer function results, we further calculate the gate fidelity as shown in the plot. It is shown that whether the non-cyclic geometric gate can offer improvement over the naive gates depends on the specific noise spectrum up to the gates. For example, for the gate R(x^,π/2)R(\hat{x},\pi/2) in (a), it performs worse than the naive gate. Meanwhile, for the gate R(x^+z^,π)R(\hat{x}+\hat{z},\pi) in (d), it performs the same as the naive gate. On the other hand, both R(z^,π/2)R(\hat{z},\pi/2) in (b) and R(x^+y^z^,4π/3)R(\hat{x}+\hat{y}-\hat{z},4\pi/3) in (c) can outperform the naive gate. On the other hand no improvement is offered by the CORPSE gate.

III.2 Two-qubit gate

The two-qubit entangling gate can be implemented by coupling two ST qubits via a cavity resonator as shown in Fig. 1(c). In the DQD eigenbasis, the total Hamiltonian of the hybrid system composed of the qubits and the resonator reads Abadillo-Uriel et al. (2019)

Htot=\displaystyle H_{\rm{tot}}= ωraa+k=12m=13Em(k)σmm(k)\displaystyle\omega_{r}a^{\dagger}a+\sum_{k=1}^{2}\sum_{m=1}^{3}E_{m}^{(k)}\sigma_{mm}^{(k)} (18)
+k=12m,n=13g(k)dmn(k)(a+a)σmn(k),\displaystyle+\sum_{k=1}^{2}\sum_{m,n=1}^{3}g^{(k)}d_{mn}^{(k)}\left(a+a^{\dagger}\right)\sigma_{mn}^{(k)},

where aa^{\dagger} (aa) is the photon creation (annihilation) operator for the resonator and ωr\omega_{r} is the intrinsic frequency of the resonator, while g(k)g^{(k)} is the coupling strength between the kkth DQD and the resonator. Em(k)E_{m}^{(k)} and σmn(k)\sigma_{mn}^{(k)} are the eigenvalue and the Pauli matrix for the kkth DQD. By moving to the rotating frame defined by U=eiH0tU^{\prime}=e^{-iH_{0}t} where

H0=ωraa+k=12m=13Em(k)σmm(k),\displaystyle H_{0}=\omega_{r}a^{\dagger}a+\sum_{k=1}^{2}\sum_{m=1}^{3}E_{m}^{(k)}\sigma_{mm}^{(k)}, (19)

one finds

Hint=\displaystyle H_{\rm{int}}^{\prime}= k=12m=13g(k)dmm(k)eiωrtσmm(k)a+k=12m<ng(k)dmn(k)(σmnaei(Em(k)En(k)+ωr)t+σmnaei(Em(k)En(k)ωr)t)+h.c.\displaystyle\sum_{k=1}^{2}\sum_{m=1}^{3}g^{(k)}d_{mm}^{(k)}e^{i\omega_{r}t}\sigma_{mm}^{(k)}a^{\dagger}+\sum_{k=1}^{2}\sum_{m<n}g^{(k)}d_{mn}^{(k)}(\sigma_{mn}a^{\dagger}e^{i(E_{m}^{(k)}-E_{n}^{(k)}+\omega_{r})t}+\sigma_{mn}ae^{i(E_{m}^{(k)}-E_{n}^{(k)}-\omega_{r})t})+h.c. (20)

In the experiments, the resonator frequency can be as high as several GHz, while the coupling strength is typically less than 100 MHz Abadillo-Uriel et al. (2019), indicating |g(k)dmn(k)|/2π50|g^{(k)}d_{mn}^{(k)}|/2\pi\leq 50 MHz and therefore we have |g(k)dmm(k)|ωr|g^{(k)}d_{mm}^{(k)}|\ll\omega_{r}. Also, we can set the proper parameters in the Hamiltonian Eq. 1 to ensure |g(k)dmn(k)||Em(k)En(k)ωr||g^{(k)}d_{mn}^{(k)}|\ll|E_{m}^{(k)}-E_{n}^{(k)}-\omega_{r}|. In this way, the leakage to the excited states |f|f\rangle (related to the terms dgfd_{gf} and defd_{ef}) and the longitudinal coupling between the qubits and the resonator (related to the terms dggd_{gg}, deed_{ee} and dffd_{ff}) are strongly suppressed. This is owing to the fact that they can be regarded as the fast-oscillating terms. Further, if we set ωq(k)=ωr\omega_{q}^{(k)}=\omega_{r}, namely, both of the qubits are in resonance with the resonator, Eq. 20 reduces to

Hint\displaystyle H_{\rm{int}}^{\prime}\approx k=12Ω(k)σge(k)a+h.c.\displaystyle\sum_{k=1}^{2}\Omega^{(k)}\sigma_{ge}^{(k)}a^{\dagger}+h.c. (21)

where Ω(k)=g(k)dge(k)\Omega^{(k)}=g^{(k)}d_{ge}^{(k)}. In the single-excitation subspace spanned by {|eg0,|ge0,|gg1}\{|eg0\rangle,|ge0\rangle,|gg1\rangle\}, Eq. 21 can establish an effective resonant three-level Λ\Lambda system. Here, |abc|abc\rangle represents the first and the second qubit and the resonator. When the evolution time satisfies 0tΩ(t)dt=π\int_{0}^{t}\Omega(t)\mathrm{d}t^{\prime}=\pi, a two-qubit entangling gate is obtained Zhou et al. (2017); Egger et al. (2019); Li et al. (2020); Zhang et al. (2020b). In the computational subspace, which corresponds to the zero-photon subspace {|gg0,|ge0,|eg0,|ee0}\{|gg0\rangle,|ge0\rangle,|eg0\rangle,|ee0\rangle\} one finds

Uent(ξ)=(10000cosξsinξ00sinξcosξ00001),\displaystyle U_{\rm{ent}}\left(\xi\right)=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&\cos\xi&\sin\xi&0\\ 0&\sin\xi&-\cos\xi&0\\ 0&0&0&-1\end{array}\right), (22)

where we have assumed Ω=(Ω(1))2+(Ω(2))2\Omega=\sqrt{\left(\Omega^{(1)}\right)^{2}+\left(\Omega^{(2)}\right)^{2}} and tanξ/2=Ω(1)/Ω(2)\tan\xi/2=-\Omega^{(1)}/\Omega^{(2)}.

The performance of the two-qubit gate can be evaluated via numerically solving the master equation Blais et al. (2007)

ρ˙=i[Hint,ρ]+Γa𝒟[a]+Γ2𝒟[σz]+Γ1𝒟[σ],\displaystyle\dot{\rho}=-i\left[H_{\rm{int}}^{\prime},\rho\right]+\Gamma_{a}\mathcal{D}[a]+\Gamma_{2}\mathcal{D}[\sigma_{z}]+\Gamma_{1}\mathcal{D}[\sigma], (23)

where

𝒟[L]=(2LρLLLρρLL)/2.\displaystyle\mathcal{D}[L]=\left(2L\rho L^{\dagger}-L^{\dagger}L\rho-\rho L^{\dagger}L\right)/2. (24)

Here, Γa\Gamma_{a} denotes the internal decay effect of the resonator, while Γ1\Gamma_{1} and Γ2\Gamma_{2} denote the relaxation and pure dephasing rate, respectively. Meanwhile, we assume the pure dephasing is mainly owing to the charge noise and the fast-oscillating terms, both of which have been considered in the Hamiltonian HintH_{\rm{int}}^{\prime}, we therefore set Γ2=0\Gamma_{2}=0. Note that, the relaxation of the ST qubits is complicated. However, as stated above, the relaxation rate in the silicon-based platform can be substantially low, since the relaxation time in the experiment has reached T1=9sT_{1}=9\ \rm{s} Ciriano-Tejel et al. (2021). Therefore, the relaxation of the coupled system can be mainly owing to the leakage to the excited state |f|f\rangle and the resonator decay.

Under the action of the entangling gate Uent(ξ=π/2)U_{\rm{ent}}\left(\xi=-\pi/2\right) which corresponds to g(1)=g(2)=gg^{(1)}=g^{(2)}=g^{\prime}, the initial state |ge0|ge0\rangle will transform into state |eg0|eg0\rangle at the final evolution time. In Fig. 5(a) and (b) we show the population of the states |ge0|ge0\rangle and |eg0|eg0\rangle, which corresponds to the cases ΔB<τ\Delta B<\tau and ΔB>τ\Delta B>\tau, respectively. Here, we set g=2π×100MHzg^{\prime}=2\pi\times 100\ \rm{MHz}. The resonator decay is taken as Γa/2π=0.028MHz\Gamma_{a}/2\pi=0.028\ \rm{MHz} according to the recent experiment, which corresponds to the resonator quality factor of about Q=105Q=10^{5} Samkharadze et al. (2016). The population of |eg0|eg0\rangle in Fig. 5(a) is obviously smaller than the case in Fig. 5(b) due to the strong leakage effect. To deeply study the robustness of the two-qubit gate, we further plot the fidelity as a function of σωq/g\sigma_{\omega q}/g^{\prime}, as shown in Fig. 5 (c) and (d). To ensure convergence, we have averaged the fidelities over 1000 implementations for each σωq/g\sigma_{\omega q}/g^{\prime}. When the noise is small, i.e. σωq/g0\sigma_{\omega q}/g^{\prime}\simeq 0, the fidelities for both cases are closed to 1. As σωq/g\sigma_{\omega q}/g^{\prime} increases, the fidelity in Fig. 5(c) drops quickly, and it is down to 92% for the large noise of σωq/g0.2\sigma_{\omega q}/g^{\prime}\simeq 0.2. For comparison, the fidelity in Fig. 5(d) falls more slowly. When σωq/g0.2\sigma_{\omega q}/g^{\prime}\simeq 0.2, the fidelity is still surpassing 97%. Thus, to achieve high-fidelity entangling gate, one is wise to operate it in the region ΔB>τ\Delta B>\tau, which is consistent with the single-qubit case. Assuming that the charge noise here is with the same level as that of the single-qubit case, namely σωq/g0.05\sigma_{\omega q}/g^{\prime}\simeq 0.05, the fidelity in Fig. 5(d) can be as high as 99.63%.

Refer to caption
Figure 5: (a) and (b): the population of the states |ge0|ge0\rangle and |eg0|eg0\rangle. (c) and (d): the fidelity of the entangling gate Uent(ξ=π/2)U_{\rm{ent}}\left(\xi=-\pi/2\right) as a function of σωq/g\sigma_{\omega q}/g^{\prime}. Parameters: for (a) and (c): ΔB/2π=1.5GHz\Delta B/2\pi=1.5\ \rm{GHz}, τ/2π=1.75GHz\tau/2\pi=1.75\ \rm{GHz} while for (b) and (d): ΔB/2π=2.5GHz\Delta B/2\pi=2.5\ \rm{GHz}, τ/2π=1.5GHz\tau/2\pi=1.5\ \rm{GHz}. The coupling strength is taken as g(1)=g(2)=g=2π×100MHzg^{(1)}=g^{(2)}=g^{\prime}=2\pi\times 100\ \rm{MHz}. The resonator decay is taken as Γa/2π=0.028MHz\Gamma_{a}/2\pi=0.028\ \rm{MHz}, which corresponds to the resonator quality factor of about Q=105Q=10^{5} Samkharadze et al. (2016).

IV Conclusions

We theoretically propose the universal gates for the ST qubits. We have shown that the single-qubit gates can be achieved by introducing an AC drive on the detuning near the TSS. The large dipole moments of the DQDs at the TSS has enabled strong coupling between the qubits and the cavity resonator, which leads to a two-qubit entangling gates. We have discussed the implementation of the gates in two different regions for the TSS. When operating in the region εSS>0\varepsilon_{\rm{SS}}>0, both the single- and two-qubit gates are having fidelity higher than 99%. Our method offers an alternative tool to realize high-fidelity ST qubits.

ACKNOWLEDGMENTS

This work was supported by Key-Area Research and Development Program of GuangDong Province (Grant No. 2018B030326001), the National Natural Science Foundation of China (Grant No. 11905065, 11874156), the Project funded by China Postdoctoral Science Foundation (Grant No. 2019M652928), the Science and Technology Program of Guangzhou (Grant No. 2019050001).

Appendix A Effective two-level Hamiltonian and the RWA effect

Here, we seek how to obtain the effective two-level Hamiltonian for the single-qubit quantum gate and estimate the RWA effect. The total Hamiltonian written in the eigenstates considering the detuning noise reads

H=n=13(En(εSS)+δEn)σnn+εm,n=13dmnσmn,H=\sum_{n=1}^{3}(E_{n}(\varepsilon_{\rm{SS}})+\delta E_{n})\sigma_{nn}+\varepsilon^{\prime}\sum_{m,n=1}^{3}d_{mn}\sigma_{mn}, (25)

where EnE_{n} is the nnth energy level (the number 1, 2, and 3 here represent states |g|g\rangle, |e|e\rangle and |f|f\rangle), δEn\delta E_{n} is the drift of the energy level induced by the detuning noise, and ε(t)=εACcos(ωt+ϕ(t))\varepsilon^{\prime}(t)=\varepsilon_{\rm{AC}}\cos(\omega t+\phi(t)). In the rotating frame defined by Ui=ei(n=13En|nn|)tU_{i}=e^{-i(\sum_{n=1}^{3}E_{n}|n\rangle\langle n|)t}, we further have

Hrot\displaystyle H_{\rm{rot}} =UiHUiiUiUit\displaystyle=U_{i}^{\dagger}HU_{i}-iU_{i}^{\dagger}\frac{\partial U_{i}}{\partial t} (26)
=\displaystyle= Hleak+Hcompu,\displaystyle H_{\rm{leak}}+H_{\rm{compu}},

where

Hleak=εAC2(dff(ei(ϕ+ωt)+ei(ϕ+ωt))+δEfdef(ei(ϕ+ωtωeft)+ei(ϕ+ωt+ωeft))dgf(ei(ϕ+ωtωgft)+ei(ϕ+ωt+ωgft))def(ei(ϕ+ωtωeft)+ei(ϕ+ωt+ωeft))00dgf(ei(ϕ+ωtωgft)+ei(ϕ+ωt+ωgft))00),\displaystyle\small H_{\rm{leak}}=\frac{\varepsilon_{\rm{AC}}}{2}\left(\begin{array}[]{ccc}d_{ff}(e^{-i(\phi+\omega t)}+e^{i(\phi+\omega t)})+\delta E_{f}&d_{ef}(e^{-i(\phi+\omega t-\omega_{ef}t)}+e^{i(\phi+\omega t+\omega_{ef}t)})&d_{gf}(e^{-i(\phi+\omega t-\omega_{gf}t)}+e^{i(\phi+\omega t+\omega_{gf}t)})\\ d_{ef}(e^{i(\phi+\omega t-\omega_{ef}t)}+e^{-i(\phi+\omega t+\omega_{ef}t)})&0&0\\ d_{gf}(e^{i(\phi+\omega t-\omega_{gf}t)}+e^{-i(\phi+\omega t+\omega_{gf}t)})&0&0\end{array}\right), (27)

and

Hcompu=(0000εACdee2(ei(ϕ+ωt)+ei(ϕ+ωt))+δEeεACdge2(ei(ϕ+ωtωqt)+ei(ϕ+ωt+ωqt))0εACdge2(ei(ϕ+ωtωqt)+ei(ϕ+ωt+ωqt))εACdgg2(ei(ϕ+ωt)+ei(ϕ+ωt))+δEg),\displaystyle\small H_{\rm{compu}}=\left(\begin{array}[]{ccc}0&0&0\\ 0&\frac{\varepsilon_{\rm{AC}}d_{ee}}{2}(e^{-i(\phi+\omega t)}+e^{i(\phi+\omega t)})+\delta E_{e}&\frac{\varepsilon_{\rm{AC}}d_{ge}}{2}(e^{-i(\phi+\omega t-\omega_{q}t)}+e^{i(\phi+\omega t+\omega_{q}t)})\\ 0&\frac{\varepsilon_{\rm{AC}}d_{ge}}{2}(e^{i(\phi+\omega t-\omega_{q}t)}+e^{-i(\phi+\omega t+\omega_{q}t)})&\frac{\varepsilon_{\rm{AC}}d_{gg}}{2}(e^{-i(\phi+\omega t)}+e^{i(\phi+\omega t)})+\delta E_{g}\end{array}\right), (28)

Here we can see that the counter-rotating terms appear in both HleakH_{\rm{leak}} and HcompuH_{\rm{compu}}, the former of which leads to leakage to the excited state |f|f\rangle, while the latter would result in gate error. Consider the RWA condition is met, namely, |εACdmn2|ω,|ω±ωgf|,|ω±ωef||\frac{\varepsilon_{\rm{AC}}d_{mn}}{2}|\ll\omega,|\omega\pm\omega_{gf}|,|\omega\pm\omega_{ef}|, and the resonance condition is also satisfied, i.e. ωq=ω\omega_{q}=\omega, then, all the counter-rotating terms vanish. The Hamiltonian is reduced to the effective two-level structure as

Hcompu=(δωq2εACdge2eiϕεACdge2eiϕδωq2)\displaystyle H_{\rm{compu}}^{{}^{\prime}}=\left(\begin{array}[]{cc}\frac{\delta\omega_{q}}{2}&\frac{\varepsilon_{\rm{AC}}d_{ge}}{2}e^{-i\phi}\\ \frac{\varepsilon_{\rm{AC}}d_{ge}}{2}e^{i\phi}&-\frac{\delta\omega_{q}}{2}\end{array}\right) (29)
=Ω02(cosϕσx+sinϕσy)+δ2σz\displaystyle=\frac{\Omega_{0}}{2}\left(\cos\phi\ \sigma_{x}+\sin\phi\ \sigma_{y}\right)+\frac{\delta}{2}\ \sigma_{z}

where Ω0=|dge|εAC\Omega_{0}=|d_{ge}|\varepsilon_{\rm{AC}} and δ=δωq=δEeδEg\delta=\delta\omega_{q}=\delta E_{e}-\delta E_{g}. To suppress the counter-rotating terms, one should carefully choose the parameters for the TSS point and the value of εAC\varepsilon_{\rm{AC}} should not be too large. For our chosen parameters as shown in the main text, Eq. 29 is always valid. This can be estimated by plotting the evolution of all the three eigenstates when operating a π\pi-pulse rotation (i.e. a NOT gate). The evolution of the states can be numerically calculated via solving the von Neumann equation

iρt=[Hrot,ρ]\displaystyle i\hbar\frac{\partial\rho}{\partial t}=[H_{\rm{rot}},\rho] (30)

As shown in Fig. 6, for a given initial state |0|0\rangle (denoted as black line), the population of the state |f|f\rangle (denoted as blue line) is always with the order of 10410^{-4}. This means that the leakage to the excited state |f|f\rangle is strongly suppressed and thus can be safely ignored. On the other hand, the population of state |0|0\rangle at the final evolution time is with the order of 10610^{-6}. This indicates that the gate error due to the counter-rotating terms in the computational subspace is substantially weak. Therefore, the RWA is highly effective, and we would focus on the two-level Hamiltonian as described in Eq. 29

Refer to caption
Figure 6: The population of the computational states |0|0\rangle and |1|1\rangle, and the leakage state |f|f\rangle when operating a π\pi-pulse rotation. Parameters: ΔB/2π=2.5GHz\Delta B/2\pi=2.5\ \rm{GHz}, τ/2π=1.5GHz\tau/2\pi=1.5\ \rm{GHz} and εAC/2π=0.1GHz\varepsilon_{\rm{AC}}/2\pi=0.1\ \rm{GHz}. For a given initial state |0|0\rangle, the population of the state |f|f\rangle (denoted as black line) is always with the order of 10410^{-4}. On the other hand, the population of state |0|0\rangle at the final evolution time is with the order of 10610^{-6}.

Appendix B Expansion of the fidelity

Here, we analytically compare the fidelity of the naive, the CORPSE, the geometric and the quantum gates. We first write down the evolution operators for each type of the gates and then derive the fidelity expressions.

Refer to caption
Figure 7: Infidelity for several gates. The gate in each panel is (a)R(x^,π/2)R(\hat{x},\pi/2), (b)R(z^,π/2)R(\hat{z},\pi/2), (c)R(x^+y^z^,4π/3)R(\hat{x}+\hat{y}-\hat{z},4\pi/3) and (d)R(x^+z^,π)R(\hat{x}+\hat{z},\pi). The design of these gates can be seen in Table. 1.

In Sec. A, we have shown that the system can be treated well as a two-level system, where the dominant detuning noise appears in the diagonal term. Therefore, we would calculate the fidelity by using the expression in Eq. 29. For this piecewise-continuous Hamiltonian, the corresponding one-piece evolution operator suffering noise reads

Rδ(ϕ,θ)exp[i(Ω02(cosϕσx+sinϕσy)+δ2σz)θΩ0]\displaystyle R_{\delta}(\phi,\theta)\equiv\exp\left[-i(\frac{\Omega_{0}}{2}\left(\cos\phi\ \sigma_{x}+\sin\phi\ \sigma_{y}\right)+\frac{\delta}{2}\ \sigma_{z})\frac{\theta}{\Omega_{0}}\right] (31)

Here, Rδ(ϕ,θ)R_{\delta}(\phi,\theta) denotes arbitrary rotation around the axis in the xx-yy plane by an angle θ\theta, where the axis on the Bloch sphere is determined by 𝒓=(cosϕ,sinϕ,0)\boldsymbol{r}=(\cos\phi,\sin\phi,0). In our case, the single-piece evolution operator Rδ(ϕ,θ)R_{\delta}(\phi,\theta) is regarded as the so-called naive gate.

Typically, Rδ(ϕ,θ)R_{\delta}(\phi,\theta) can be dynamically corrected by using the CORPSE gates which includes 3 elementary pulse Rδ(ϕ,θ)R_{\delta}(\phi,\theta) Cummins et al. (2003); Bando et al. (2013):

Rcorpse(ϕ,θ)=Rδ(ϕc3,θc3).Rδ(ϕc2,θc2).Rδ(ϕc1,θc1).\displaystyle R_{\rm{corpse}}(\phi,\theta)=R_{\delta}(\phi_{c3},\theta_{c3}).R_{\delta}(\phi_{c2},\theta_{c2}).R_{\delta}(\phi_{c1},\theta_{c1}). (32)

where

θc1=2n1π+θ/2arcsin[sin(θ/2)/2],θc2=2n2π2arcsin[sin(θ/2)/2],θc3=2n3π+θ/2arcsin[sin(θ/2)/2],ϕc1=ϕc2π=ϕc3=ϕ.\displaystyle\begin{array}[]{l}\theta_{c1}=2n_{1}\pi+\theta/2-\arcsin[\sin(\theta/2)/2],\\ \theta_{c2}=2n_{2}\pi-2\arcsin[\sin(\theta/2)/2],\\ \theta_{c3}=2n_{3}\pi+\theta/2-\arcsin[\sin(\theta/2)/2],\\ \phi_{c1}=\phi_{c2}-\pi=\phi_{c3}=\phi.\end{array} (33)

where nin_{i} (i=1,2,3i=1,2,3) is being integer. Here, we consider using the short CORPSE pulse, which corresponds to n1=n3=0n_{1}=n_{3}=0, n2=1n_{2}=1.

For our quantum gate as shown in Eq. 6 in the main text, the corresponding evolution operator is

Rnon(χ0,ϕ0,ϕ1,β0)=Rδ(ϕn2,θn2).Rδ(ϕn1,θn1)\displaystyle R_{\rm{non}}(\chi_{0},\phi_{0},\phi_{\rm{1}},\beta_{0})=R_{\delta}(\phi_{n2},\theta_{n2}).R_{\delta}(\phi_{n1},\theta_{n1}) (34)

where

θq1=χ0,θq2=β0,ϕq1=ϕ0+π/2,ϕq2=ϕ0+ϕ1+π/2.\displaystyle\begin{array}[]{l}\theta_{q1}=\chi_{0},\\ \theta_{q2}=\beta_{0},\\ \phi_{q1}=\phi_{0}+\pi/2,\\ \phi_{q2}=\phi_{0}+\phi_{1}+\pi/2.\end{array} (35)

Using Rnon(χ0,ϕ0,ϕ1,β0)R_{\rm{non}}(\chi_{0},\phi_{0},\phi_{\rm{1}},\beta_{0}) one is able to design arbitrary rotation. For example, the rotation around xx axis can be designed as Rnon(χ0,ϕ0=π2,ϕ1=π,β0)R_{\rm{non}}(\chi_{0},\phi_{0}=\frac{\pi}{2},\phi_{\rm{1}}=\pi,\beta_{0}), and the rotation angle is β0χ0\beta_{0}-\chi_{0}.

On the other hand, the geometric gates considered in this paper are also based on the 3 elementary pulse Rδ(ϕ,θ)R_{\delta}(\phi,\theta) Zhao et al. (2017); Zhang et al. (2020a):

Rgeo(θ,ϕ,γ)=eiγ2nσ\displaystyle R_{\rm{geo}}(\theta^{\prime},\phi^{\prime},\gamma^{\prime})=e^{-i\frac{\gamma^{\prime}}{2}\vec{n}\cdot\vec{\sigma}} (36)
=\displaystyle= Rδ(ϕgeo3,θgeo3).Rδ(ϕgeo2,θgeo2).Rδ(ϕgeo1,θgeo1).\displaystyle R_{\delta}(\phi_{\rm{geo}3},\theta_{\rm{geo}3}).R_{\delta}(\phi_{\rm{geo}2},\theta_{\rm{geo}2}).R_{\delta}(\phi_{\rm{geo}1},\theta_{\rm{geo}1}).

where

θgeo1=θ,θgeo2=π,θgeo3=πθ,ϕgeo1=ϕπ/2,ϕgeo2=ϕγ/2π/2,ϕgeo3=ϕπ/2.\displaystyle\begin{array}[]{l}\theta_{\rm{geo}1}=\theta^{\prime},\\ \theta_{\rm{geo}2}=\pi,\\ \theta_{\rm{geo}3}=\pi-\theta^{\prime},\\ \phi_{\rm{geo}1}=\phi-\pi/2,\\ \phi_{\rm{geo}2}=\phi-\gamma^{\prime}/2-\pi/2,\\ \phi_{\rm{geo}3}=\phi-\pi/2.\end{array} (37)

and n=(sinθcosϕ,sinθsinϕ,cosθ)\vec{n}=(\sin\theta^{\prime}\cos\phi^{\prime},\sin\theta^{\prime}\sin\phi^{\prime},\cos\theta^{\prime}).

Here, we treat δ\delta as a quasi-static perturbation with |δ|1|\delta|\ll 1, then, we can expand the fidelity for each type of gates by using the evolution operator above. It is well known that any single-qubit gate can be further decomposed into a “xx-yy-xx” rotation Nielsen and Chuang (2002):

Rx(ϕa=0,θa)Ry(ϕb=π/2,θb)Rx(ϕc=0,θc),R_{x}\left(\phi_{a}=0,\theta_{a}\right)R_{y}\left(\phi_{b}=\pi/2,\theta_{b}\right)R_{x}\left(\phi_{c}=0,\theta_{c}\right), (38)

Therefore, we can focus on the performance of the rotation around the xx (yy) axis by an arbitrary rotation angle γ\gamma. The corresponding fidelity expressions around the xx axis for each gate are

δ(x^,γ)14(4δ2+δ2cosγ),non(x^,γ)1+14δ2(3+2cosβ0cosγ+2cosχ0),geo(x^,γ)13δ24+δ2cosγ214δ2cosγ,\displaystyle\begin{array}[]{l}\mathcal{F}_{\delta}(\hat{x},\gamma)\approx\frac{1}{4}\left(4-\delta^{2}+\delta^{2}\cos\gamma\right),\\ \mathcal{F}_{\rm{non}}(\hat{x},\gamma)\approx 1+\frac{1}{4}\delta^{2}(-3+2\cos\beta_{0}-\cos\gamma+2\cos\chi_{0}),\\ \mathcal{F}_{\rm{geo}}(\hat{x},\gamma)\approx 1-\frac{3\delta^{2}}{4}+\delta^{2}\cos\frac{\gamma}{2}-\frac{1}{4}\delta^{2}\cos\gamma,\end{array} (39)

and

corpse(x^,γ)132(32+(7+(2π+γ)2)δ4)\displaystyle\mathcal{F}_{\rm{corpse}}(\hat{x},\gamma)\approx\frac{1}{32}\left(-32+\left(7+(-2\pi+\gamma)^{2}\right)\delta^{4}\right) (40)
\displaystyle- 132δ4(6cosγ+cos2γ+2sinγ2((4π2γ)cosγ2+27+cosγ(2π+γ+sinγ))).\displaystyle\frac{1}{32}\delta^{4}\left(6\cos\gamma+\cos 2\gamma+2\sin\frac{\gamma}{2}\left((4\pi-2\gamma)\cos\frac{\gamma}{2}+\sqrt{2}\sqrt{7+\cos\gamma}(-2\pi+\gamma+\sin\gamma)\right)\right).

Here, we see that the first order effect of the detuning noise vanishes for all the gates. Meanwhile, CORPSE gate can eliminate the noise effect up to 4th order, which outperforms other gates. We also find that the gate fidelity depends on the rotation angle γ\gamma. The non-cyclic geometric gate can outperform the naive gate when π<γ<π-\pi<\gamma<\pi. On the other hand, its performance still depends on the choice of χ0\chi_{0} and β0\beta_{0}. For a simple choice of χ0=0.01γ\chi_{0}=0.01\gamma and β0=1.01γ\beta_{0}=1.01\gamma (γ=β0χ0\gamma=\beta_{0}-\chi_{0}), the non-cyclic geometric gate can correct the naive gate in the region between 2π<γ<π-2\pi<\gamma<-\pi and π<γ<2π\pi<\gamma<2\pi. The case for the rotation around the yy axis is similar and we would not discuss it in detail. In Fig. 7, we show the infidelity for several considered gates when suffering noise.

Appendix C An example to calculate filter transfer function

Here, we introduce how to step-by-step calculate the piecewise continuous filter function. We take a special case of the short CORPSE gate for example, i.e. R(x^,π/2)R(\hat{x},\pi/2), which includes three piecewise pulse. The case for other types of gate is just a simplified version. For R(x^,π/2)R(\hat{x},\pi/2) with ϕ=0\phi=0 and θ=π/2\theta=\pi/2, we can derive the useful parameters from Eq.33 as: θc1=θc3=π/4arcsin(122),θc2=2π2arcsin(122)\theta_{c1}=\theta_{c3}=\pi/4-\arcsin\left(\frac{1}{2\sqrt{2}}\right),\theta_{c2}=2\pi-2\arcsin\left(\frac{1}{2\sqrt{2}}\right), ϕ1c=ϕ3c=0\phi_{1c}=\phi_{3c}=0 and ϕ2c=π\phi_{2c}=\pi. The corresponding piecewise Hamiltonian are thus H1=Ω02σx,H2=Ω02σx,H3=Ω02σxH_{1}=\frac{\Omega_{0}}{2}\sigma_{x},H_{2}=-\frac{\Omega_{0}}{2}\sigma_{x},H_{3}=\frac{\Omega_{0}}{2}\sigma_{x}. The time interval is Tk=θck/Ω0T_{k}=\theta_{ck}/\Omega_{0} for k=1,2,3k=1,2,3.

To calculate the piecewise control matrix as shown in Eq.12, define the evolution operator during each interval as Uk(t)=eitHkU_{k}(t)=e^{-itH_{k}}. Let Vij[U]=Tr(UσiUσj)/2V_{ij}[U]=\operatorname{Tr}\left(U^{\dagger}\sigma_{i}U\sigma_{j}\right)/2. The corresponding piecewise control matrix is therefore Rij(k)(ω)=iω0Tk𝑑teiωtVij[Uk(t)]R_{ij}^{(k)}(\omega)=-i\omega\int_{0}^{T_{k}}dte^{i\omega t}V_{ij}\left[U_{k}(t)\right]. We first calculate the indefinite integral for the control matrix with R(k)(t)=iω𝑑teiωtVij[Uk(t)]R^{\prime(k)}(t)=-i\omega\int dte^{i\omega t}V_{ij}\left[U_{k}(t)\right], where

R(1)(t)=R(3)(t)=(eiωt000ωeiωt(ωcosΩ0tiΩ0sinΩ0t)ω2Ω02ωeiωt(iΩ0cosΩ0t+ωsinΩ0t)ω2Ω020ωeiωt(iΩ0cosΩ0t+ωsinΩ0t)ω2Ω02ωeiωt(ωcosΩ0tiΩ0sinΩ0t)ω2Ω02),\displaystyle R^{\prime(1)}(t)=R^{\prime(3)}(t)=\left(\begin{array}[]{ccc}-e^{i\omega t}&0&0\\ 0&-\frac{\omega e^{i\omega t}(\omega\cos\Omega_{0}t-i\Omega_{0}\sin\Omega_{0}t)}{\omega^{2}-\Omega_{0}^{2}}&\frac{\omega e^{i\omega t}(i\Omega_{0}\cos\Omega_{0}t+\omega\sin\Omega_{0}t)}{\omega^{2}-\Omega_{0}^{2}}\\ 0&-\frac{\omega e^{i\omega t}(i\Omega_{0}\cos\Omega_{0}t+\omega\sin\Omega_{0}t)}{\omega^{2}-\Omega_{0}^{2}}&-\frac{\omega e^{i\omega t}(\omega\cos\Omega_{0}t-i\Omega_{0}\sin\Omega_{0}t)}{\omega^{2}-\Omega_{0}^{2}}\end{array}\right), (41)

and

R(2)(t)=(eiωt000ωeiωt(ωcosΩ0tiΩ0sinΩ0t)ω2Ω02ωeiωt(iΩ0cosΩ0t+ωsinΩ0t)ω2Ω020ωeiωt(iΩ0cosΩ0t+ωsinΩ0t)ω2Ω02ωeiωt(ωcosΩ0tiΩ0sinΩ0t)ω2Ω02),\displaystyle R^{\prime(2)}(t)=\left(\begin{array}[]{ccc}-e^{i\omega t}&0&0\\ 0&-\frac{\omega e^{i\omega t}(\omega\cos\Omega_{0}t-i\Omega_{0}\sin\Omega_{0}t)}{\omega^{2}-\Omega_{0}^{2}}&-\frac{\omega e^{i\omega t}(i\Omega_{0}\cos\Omega_{0}t+\omega\sin\Omega_{0}t)}{\omega^{2}-\Omega_{0}^{2}}\\ 0&\frac{\omega e^{i\omega t}(i\Omega_{0}\cos\Omega_{0}t+\omega\sin\Omega_{0}t)}{\omega^{2}-\Omega_{0}^{2}}&-\frac{\omega e^{i\omega t}(\omega\cos\Omega_{0}t-i\Omega_{0}\sin\Omega_{0}t)}{\omega^{2}-\Omega_{0}^{2}}\end{array}\right), (42)

Then, one can easily compute the integral results R(k)(ω)R^{(k)}(\omega). According to Eq.(29) in Ref.Green et al. (2013), one can further define a matrix 𝚲(k)\mathbf{\Lambda}^{(k)} where

𝚲(k)=V[Qk],Qk=Uk(Tk)Uk1(Tk1),,U1(T1)\mathbf{\Lambda}^{(k)}=V[Q_{k}],\ \ Q_{k}=U_{k}(T_{k})U_{k-1}(T_{k-1}),...,U_{1}(T_{1}) (43)

In this way, the control matrix is

𝑹(ω)=k=1neiωTk1𝑹k(ω)𝚲(k1)\boldsymbol{R}(\omega)=\sum_{k=1}^{n}e^{i\omega T^{\prime}_{k-1}}\boldsymbol{R}^{k}(\omega)\boldsymbol{\Lambda}^{(k-1)} (44)

where Tk1=T1+T2+,,+Tk1T^{\prime}_{k-1}=T_{1}+T_{2}+,...,+T_{k-1}.

Considering the CORPSE case for R(x^,π/2)R(\hat{x},\pi/2) with n=3n=3, we have

𝚲(1)\displaystyle\boldsymbol{\Lambda}^{(1)} =V[Q1]=V[eiθc1σx/2]\displaystyle=V\left[Q_{1}\right]=V\left[e^{-i\theta_{c1}\sigma_{x}/2}\right] (45)
𝚲(2)\displaystyle\boldsymbol{\Lambda}^{(2)} =V[Q2]=V[ei(θc2+θc1)σx/2]\displaystyle=V\left[Q_{2}\right]=V\left[e^{-i\left(-\theta_{c2}+\theta_{c1}\right)\sigma_{x}/2}\right]

and further

𝑹(ω)=𝑹(1)(ω)+eiωT1𝑹(2)(ω)𝚲(1)+eiω(T2+T1)𝑹(3)(ω)𝚲(2)\boldsymbol{R}(\omega)=\boldsymbol{R}^{(1)}(\omega)+e^{i\omega T_{1}}\boldsymbol{R}^{(2)}(\omega)\boldsymbol{\Lambda}^{(1)}+e^{i\omega\left(T_{2}+T_{1}\right)}\boldsymbol{R}^{(3)}(\omega)\boldsymbol{\Lambda}^{(2)} (46)

By inserting the zz-component of matrix 𝑹(ω)\boldsymbol{R}(\omega) into Eq.16, one can obtain the analytical filter function as

Fz(ω)ω2=\displaystyle\frac{F_{z}(\omega)}{\omega^{2}}= 2(ω2Ω02)2(ω226ωΩ0(7+i)4ωΩ0e4iωcsc1(22)Ω0Ω02(5+(7+1)cosω(π4sec1(22))4Ω0+\displaystyle\frac{2}{(\omega^{2}-\Omega_{0}^{2})^{2}}(\omega^{2}-2^{\frac{6\omega}{\Omega_{0}}}(\sqrt{7}+i)^{-\frac{4\omega}{\Omega_{0}}}e^{\frac{4i\omega\csc^{-1}(2\sqrt{2})}{\Omega_{0}}}\Omega_{0}^{2}(-5+(\sqrt{7}+1)\cos\frac{\omega(\pi-4\sec^{-1}(2\sqrt{2}))}{4\Omega_{0}}+ (47)
3cosω(π+2sec1(22))Ω0(71)cos3ω(π+4sec1(22))4Ω0)+\displaystyle 3\cos\frac{\omega(\pi+2\sec^{-1}(2\sqrt{2}))}{\Omega_{0}}-(\sqrt{7}-1)\cos\frac{3\omega(\pi+4\sec^{-1}(2\sqrt{2}))}{4\Omega_{0}})+
212ωΩ0(7+i)8ωΩ0e8iωcsc1(22)Ω0ωΩ0((71)sinω(π4sec1(22))4Ω0+\displaystyle 2^{\frac{12\omega}{\Omega_{0}}}(\sqrt{7}+i)^{-\frac{8\omega}{\Omega_{0}}}e^{\frac{8i\omega\csc^{-1}(2\sqrt{2})}{\Omega_{0}}}\omega\Omega_{0}((\sqrt{7}-1)\sin\frac{\omega(\pi-4\sec^{-1}(2\sqrt{2}))}{4\Omega_{0}}+
(7+1)sin3ω(π+4sec1(22))4Ω02sinω(π+8sec1(22))2Ω0));\displaystyle(\sqrt{7}+1)\sin\frac{3\omega(\pi+4\sec^{-1}(2\sqrt{2}))}{4\Omega_{0}}-2\sin\frac{\omega(\pi+8\sec^{-1}(2\sqrt{2}))}{2\Omega_{0}}));

Appendix D Rotations used for the filter function

As shown in Table. 1, the specific gates are used to calculate fidelity using the transfer function. The CORPSE gates are not listed since they are just the simple transform of the naive gates according to Eq. 32.

Table 1: Gates used for simulation.
Element Naive Geometric Non-cyclic geometric
R(x^,π2)R\left(\hat{x},\frac{\pi}{2}\right) R(x^,π2)R\left(\hat{x},\frac{\pi}{2}\right) Rgeo(π2,0,π4)R_{\rm{geo}}\left(\frac{\pi}{2},0,-\frac{\pi}{4}\right) Rnon(π8,π2,π,5π8)R_{\rm{non}}\left(\frac{\pi}{8},\frac{\pi}{2},\pi,\frac{5\pi}{8}\right)
R(z^,π2)R\left(\hat{z},\frac{\pi}{2}\right) R(x^,π2)R(y^,π2)R(x^,π2)R\left(\hat{x},\frac{\pi}{2}\right)R\left(\hat{y},\frac{\pi}{2}\right)R\left(-\hat{x},\frac{\pi}{2}\right) Rgeo(0,0,π4)R_{\rm{geo}}\left(0,0,-\frac{\pi}{4}\right) Rnon(π,0,π4,π)R_{\rm{non}}\left(\pi,0,\frac{\pi}{4},\pi\right)
R(x^+z^,π)R(\hat{x}+\hat{z},\pi) R(y^,π2)R(x^,π)R\left(-\hat{y},\frac{\pi}{2}\right)R(\hat{x},\pi) Rgeo(π4,0,π2)R_{\rm{geo}}\left(\frac{\pi}{4},0,-\frac{\pi}{2}\right) Rnon(π2,0,π2,π)R_{\rm{non}}\left(\frac{\pi}{2},0,-\frac{\pi}{2},\pi\right)
R(x^+y^z^,4π3)R\left(\hat{x}+\hat{y}-\hat{z},\frac{4\pi}{3}\right) R(x^,π2)R(y^,π2)R\left(-\hat{x},\frac{\pi}{2}\right)R\left(-\hat{y},\frac{\pi}{2}\right) Rgeo(πtan1(2),π4,2π3)R_{\rm{geo}}\left(\pi-\tan^{-1}(\sqrt{2}),\frac{\pi}{4},-\frac{2\pi}{3}\right) Rnon(3π2,0,π2,π2)R_{\rm{non}}\left(\frac{3\pi}{2},0,\frac{\pi}{2},\frac{\pi}{2}\right)

References