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

Hybrid noise protection of logical qubits for universal quantum computation

Zhao-Ming Wang1111[email protected], Feng-Hua Ren2222[email protected], Mark S. Byrd3333[email protected], and Lian-Ao Wu4,5444[email protected] 1 College of Physics and Optoelectronic Engineering, Ocean University of China, Qingdao 266100, China
2 School of Information and Control Engineering, Qingdao University of Technology, Qingdao 266520, China
3 Department of Physics, Southern Illinois University, Carbondale, Illinois 62901-4401, USA
4 Department of Physics, University of the Basque Country UPV/EHU, 48080 Bilbao, Spain IKERBASQUE Basque Foundation for Science, 48013 Bilbao, Spain
5 EHU Quantum Center, University of the Basque Country UPV/EHU, Leioa, Biscay 48940, Spain
Abstract

Quantum computers now show the promise of surpassing any possible classical machine. However, errors limit this ability and current machines do not have the ability to implement error correcting codes due to the limited number of qubits and limited control. Therefore, dynamical decoupling (DD) and encodings that limit noise with fewer qubits are more promising. For these reasons, we put forth a model of universal quantum computation that has many advantages over strategies that require a large overhead such as the standard quantum error correcting codes. First, we separate collective noise from individual noises on physical qubits and use a decoherence-free subspace (DFS) that uses just two qubits for its encoding to eliminate collective noise. Second, our bath model is very general as it uses a spin-boson type bath but without any Markovian assumption. Third, we are able to either use a steady global magnetic field or to devise a set of DD pulses that remove much of the remaining noise and commute with the logical operations on the encoded qubit. This allows removal of noise while implementing gate operations. Numerical support is given for this hybrid protection strategy which provides an efficient approach to deal with the decoherence problems in quantum computation and is experimentally viable for several current quantum computing systems. This is emphasized by a recent experiment on superconducting qubits which shows promise for increasing the number of gates that can be implemented reliably with some realistic parameter assumptions.

I Introduction

Quantum algorithms Shor97 have been developed that are much more efficient than the best known classical counterpart. However, it is difficult to perform reliable calculations on a quantum computer due to the well-known problem of decoherence. A variety of schemes for combating the deleterious environment effects have been proposed, including quantum-error correction, DD and DFS/noiseless encodings. Quantum-error correction was shown by Shor et al. Shor95 and subsequent work extended this and demonstrated that techniques exist that can be used to significantly reduce the quantum error rate Knill ; Gottesman ; Lidar ; Joschka . However, fully implemented quantum error correction technology cannot be implemented reliably in the noisy intermediate-scale quantum era NISQ ; Sankar and thus some decoherence is unavoidable sc_review_1 . As a result, DD or DFS encodings have been sought to reduce decoherence and extend the functionality of these machines. DD is an active protection strategy and it requires an efficient control of qubits in the presence of noise Wu2003 ; Uhrig ; Sekiguchi . On the other hand, DFSs Duan ; Zanardi ; Brown are a type of passive protection. Both of these two schemes have their advantages and limitations, e.g., DFSs encodings may not be easy to design and have been primarily applicable to collective noise. DD schemes can combat both collective and individual noises but the controls may be difficult to implement experimentally Jing2013 ; PyshkinSR . In this paper, we design a hybrid scheme for reliable quantum gate operations. Specifically, we use DD or a steady global magnetic field and DFSs to deal with the X,YX,Y noises and the collective ZZ noises while enabling reliable computation.

Computation in subspaces remains unaffected by the interaction with the environment when the interaction Hamiltonian has a certain symmetry Duan . Coherence control of two logical qubits encoded in a DFS has been demonstrated, and the DFS encoding has proven to have high fidelity Henry . The exchange-only gating scheme of DiVincenzo et al. encodes three physical qubits into a logical qubit DiVincenzo ; Kempe . The number of physical qubits and gate operations will be ameliorated for the XXZ type couplings Wu ; Lidar2002prl ; WuA ; Nori . Numerous experimental application of DFSs have been demonstrated, such as in trapped ions KIELPINSKIV , in an optical Deutsch-Jozsa algorithm Mohseni , in a linear-optical experiment Kwiat , and in NMR Fortunato .

DFSs can be quite useful for collective baths and recent experiment shows that collective baths do exist. Charge noises in a superconducting multiqubit circuit chip have been found highly correlated on a length scale over 600 micrometres Wilen . Recently, quantum computation in DFS has been proved to be possible and the computation is robust against collective decoherence in quantum systems Pyshkin . However, collective baths are special and individual baths Unruh ; JingjunSR are more common. More generally, individual and collective baths both coexist in many systems. In this paper, we consider a mixed bath, which includes both collective Duan98 ; Yavuz and individual noises Unruh . We use a simple encoding scheme with one logical qubit encoded by two physical qubits as in Ref. Wu . Also, the mitigation of noises was addressed by algebraic means complemented with numerical optimal control in the Markovian regime of Lindblad or Bloch-Redfield type JPB . Considering the non-Markovian environmental noises, for the gate operations, we calculate the fidelity dynamics. On the other hand, both the individual and the collective X,YX,Y noises can be eliminated by a steady global magnetic field on our entire quantum computation, so that we don’t require any additional physical operations. Similarly, we can also use the DD technique and a global leakage elimination operator (LEO), formulated as iσiz\sum_{i}\sigma^{z}_{i} to eliminate both collective and individual X,YX,Y noises. Since this global LEO commutes with all logical operations in the entire quantum computation process, it can be implemented independent of the gating operations. As for the leftover individual ZZ noise, we study the effects of the environment parameters on the obtainable rotation angle for given fidelity and find the region where gates remain accurate even if the individual ZZ noise is relatively strong. We show some threshold within which the universal set of gates works perfectly.

II Model

The total Hamiltonian has the form

Htot=Hs+Hb+Hint,H_{tot}=H_{s}+H_{b}+H_{int}, (1)

where HsH_{s} is the system Hamiltonian, Hb=j=0NHbjH_{b}=\sum_{j=0}^{N}H_{b}^{j} is the bath Hamiltonian and HintH_{int} is the interaction Hamiltonian. Suppose j=0j=0 corresponds to a collective bath and j=1,2,..Nj=1,2,..N correspond to NN-independent individual bath operators, then Hbj=kωkjbkjbkjH_{b}^{j}=\sum_{k}\omega_{k}^{j}b_{k}^{j{\dagger}}b_{k}^{j}. ωkj\omega_{k}^{j} is the boson’s frequency of the kkth mode and bkj,bkjb_{k}^{j{\dagger}},b_{k}^{j} are the bosonic creation and annihilation operators. The interaction reads

Hint=j,k(gkjLjbkj+gkjLjbkj),H_{int}=\sum\nolimits_{j,k}(g_{k}^{j\ast}L_{j}^{{\dagger}}b_{k}^{j}+g_{k}^{j}L_{j}b_{k}^{j{\dagger}}), (2)

where LjL_{j} are system operators, the subscript indicates the coupling to the jjth bath, and gkjg^{j}_{k} is the coupling constant between the system and kkth mode of the jjth bath. Clearly, L0L_{0} and gk0g^{0}_{k} describe the coupling between the system and the collective bath.

Assume initially that the bath is in a thermal equilibrium state at temperature TjT_{j} with the density operator ρj(0)=eβHbj/Zj,\rho_{j}(0)=e^{-\beta H_{b}^{j}}/Z_{j}, where Zj=Z_{j}=Tr[eβHbj][e^{-\beta H_{b}^{j}}] is the partition function, and β=1/(KBTj)\beta=1/(K_{B}T_{j}). The initial density matrix operator is assumed to be in a product state with the bath, ρ(0)=ρs(0)ρb(0)=|ψ0ψ0|j=1Nρj(0)\rho(0)=\rho_{s}(0)\otimes\rho_{b}(0)=\left|\mathbf{\psi}_{0}\right\rangle\left\langle\mathbf{\psi}_{0}\right|\bigotimes\limits_{j=1}^{N}\rho_{j}(0). Here ρs(t)\rho_{s}(t) and ρb(t)\rho_{b}(t) are the system and bath density matrix, respectively. Here we assume an uncorrelated initial state between the system and baths. However for a correlated one, the construction of the density matrix does not maintain the positivity of the density matrix Alicki . For non-Markovian baths, its asymptotic state strongly depends on the initial conditions Dariusz . A measure to quantify the influence of the initial state of an open system on its dynamics is proposed recently, and conditions under which the asymptotic state exists are derived Wenderoth . In this paper we use a newly developed theoretical tool, which is referred to as the non-Markovian quantum state diffusion (QSD) approach Diosi98 ; YTPRA ; Strunz1999 ; Wang2021 ; Ren2020 . The non-Markovian master equation is given by Wang2021

tρs\displaystyle\frac{\partial}{\partial t}\rho_{s} =\displaystyle= i[Hs,ρs]+j{[Lj,ρsO¯zj(t)][Lj,O¯zj(t)ρs]\displaystyle-i[H_{s},\rho_{s}]+\sum\nolimits_{j}\{[L_{j},\rho_{s}\overline{O}_{z}^{j{\dagger}}(t)]-[L_{j}^{{\dagger}},\overline{O}_{z}^{j}(t)\rho_{s}] (3)
+[Lj,ρsO¯wj(t)][Lj,O¯wj(t)ρs]},\displaystyle+[L_{j}^{{\dagger}},\rho_{s}\overline{O}_{w}^{j{\dagger}}(t)]-[L_{j},\overline{O}_{w}^{j}(t)\rho_{s}]\},

where O¯z,(w)j(t)=0t𝑑sαz,(w)j(ts)Oz,(w)j(t,s)\overline{O}_{z,(w)}^{j}(t)=\int_{0}^{t}ds\alpha_{z,(w)}^{j}(t-s)O_{z,(w)}^{j}(t,s) and αz,(w)j(ts)\alpha_{z,(w)}^{j}(t-s) is the correlation function. The operator OO is an ansatz and is assumed to be noise-independent here. The operator OO is an ansatz and is assumed to be noise-independent here. Generally the O¯\overline{O} operators contain noises except for some special cases, such as the case that the system Hamiltonian commutes with the Lindblad operators YTPRA . Also, when the bath couples weakly to the system, the noise-dependent O¯z,(w)j(t,z,w)\overline{O}_{z,(w)}^{j}(t,z^{*},w^{*}) operator is approximated well by a time-independent operator O¯z,(w)j(t)\overline{O}_{z,(w)}^{j}(t) Diosi98 ; YTPRA ; Struntz . Now we use the Lorentz-Drude spectrum as an example to obtain the correlation function, where the spectral density is Jj(ω)=Γjπωj1+(ωj/γj)2J_{j}(\omega)=\frac{\Gamma_{j}}{\pi}\frac{\omega_{j}}{1+(\omega_{j}/\gamma_{j})^{2}} Ritschel ; Meier . Here Γj\Gamma_{j} represents the strength of the jjth pair system-bath coupling. γj\gamma_{j} is the characteristic frequency of the jjth bath. In the high temperature or low frequency limit, references Wang2021 ; Ren2020 therefore derives closed equations for O¯z,(w)j\overline{O}_{z,(w)}^{j} to numerically solve the non-Markovian master equation (3)

O¯zj/t\displaystyle\partial\overline{O}_{z}^{j}/\partial t =\displaystyle= (ΓjTjγjiΓjγj2)Lj/2γjO¯zj\displaystyle\left(\Gamma_{j}T_{j}\gamma_{j}-i\Gamma_{j}\gamma_{j}^{2}\right)L_{j}/2-\gamma_{j}\overline{O}_{z}^{j} (4)
[iHs+j(LjO¯zj+LjO¯wj),O¯zj],\displaystyle-[iH_{s}+\sum\nolimits_{j}(L_{j}^{{\dagger}}\overline{O}_{z}^{j}+L_{j}\overline{O}_{w}^{j}),\overline{O}_{z}^{j}],
O¯wj/t\displaystyle\partial\overline{O}_{w}^{j}/\partial t =\displaystyle= ΓjTjγjLj/2γjO¯wj\displaystyle\Gamma_{j}T_{j}\gamma_{j}L_{j}^{{\dagger}}/2-\gamma_{j}\overline{O}_{w}^{j} (5)
[iHs+j(LjO¯zj+LjO¯wj),O¯wj].\displaystyle-[iH_{s}+\sum\nolimits_{j}(L_{j}^{{\dagger}}\overline{O}_{z}^{j}+L_{j}\overline{O}_{w}^{j}),\overline{O}_{w}^{j}].

Simple encodings of one logical qubit into two physical qubits have been suggested to avoid difficult-to-implement single-qubit control terms Wu . In this case, the DFS, encoded in one pair of spins, is |0L\left|0\right\rangle_{L}, |1L\left|1\right\rangle_{L}. The subscript LL denotes the logical qubit. |0L\left|0\right\rangle_{L}=|01\left|01\right\rangle, |1L\left|1\right\rangle_{L}=|10\left|10\right\rangle. The single qubit gates in this subspace can be written in terms of the generators of SU(2) as follows Wu : Tx=(σ1xσ2x+σ1yσ2y)/2T_{x}=(\sigma^{x}_{1}\sigma^{x}_{2}+\sigma^{y}_{1}\sigma^{y}_{2})/2, Ty=(σ1yσ2xσ1xσ2y)/2T_{y}=(\sigma^{y}_{1}\sigma^{x}_{2}-\sigma^{x}_{1}\sigma^{y}_{2})/2, Tz=(σ1zσ2z)/2T_{z}=(\sigma^{z}_{1}-\sigma^{z}_{2})/2.

Now we consider the quantum gate operations of one logical qubit in the presence of noise which is separated into individual and collective parts. The quantum gate fidelity is defined as F(t)=𝑑ψ(0)ψ(0)|Uρs(t)U|ψ(0)F(t)=\int{d\psi(0)\left\langle\psi(0)\right|U^{\dagger}\rho_{s}(t)U\left|\psi(0)\right\rangle}, where U=eiHst|ψ(0)U=e^{-iH_{s}t}\left|\psi(0)\right\rangle and |ψ(0)\left|\psi(0)\right\rangle is the arbitrary initial state of the system. ρs(t)\rho_{s}(t) is the system’s reduced density matrix in Eq. (3).

III Results and discussions

First consider a single-qubit gate with the system Hamiltonian Hs=JTx,(z)H_{s}=-JT_{x,(z)}. Here JJ is the coupling constant. Suppose both physical spins encounter a collective bath Hb0H_{b}^{0}, and each of them couples to an individual bath with Hamiltonian HbjH_{b}^{j} (j=1,2j=1,2). We use a parameter α\alpha to represent the degree of mixing of the two types of baths. The interaction can then be rewritten as

Hint=cos2ασ0B0+sin2α(σ1B1+σ2B2),\displaystyle H_{int}=\cos^{2}\alpha\vec{\sigma_{0}}\cdot\vec{B_{0}}+\sin^{2}\alpha(\vec{\sigma_{1}}\cdot\vec{B_{1}}+\vec{\sigma_{2}}\cdot\vec{B_{2}}), (6)

where the subscript 0 denotes the collective bath, 1,21,2 denote the individual baths of spin 11 and 22, respectively. The Pauli vector σi=(σix,σiy,σiz)\vec{\sigma_{i}}=(\sigma_{i}^{x},\sigma_{i}^{y},\sigma_{i}^{z}) and the bath operators Bi=(Bix,Biy,Biz)\vec{B_{i}}=(B_{i}^{x},B_{i}^{y},B_{i}^{z}) with Bix(y,z)B_{i}^{x(y,z)} (i=0,1,2i=0,1,2) representing the X(Y,Z)X(Y,Z) noises. Comparing with the interaction in Eq. (2), Li=σiL_{i}=\sigma_{i} and Bi=k(gkibki+gkibki)B_{i}=\sum\nolimits_{k}(g_{k}^{i\ast}b_{k}^{i}+g_{k}^{i}b_{k}^{i{\dagger}}). Note that the Lindblad operator L=L=σxL=L^{\dagger}=\sigma_{x} or σz\sigma_{z} for spin boson or dephasing. Then the rotating-wave approximation IntravaiaA ; IntravaiaB is not used here. For generality, we can also use αx,(y,z)\alpha_{x,(y,z)} to denote the mixture degree of the X(Y,Z)X(Y,Z) noises. For example, αz=0\alpha_{z}=0 (π/2\pi/2) correspond to a collective (two individual) ZZ noise. The bath can be written as Hb=cos2αHb0+sin2α(Hb1+Hb2)H_{b}=\cos^{2}\alpha H_{b}^{0}+\sin^{2}\alpha(H_{b}^{1}+H_{b}^{2}).

At first, assume there are only X,YX,Y noises present, so

Hint\displaystyle H_{int} =\displaystyle= cos2αxσ0xB0x+sin2αxi=12σixBix\displaystyle\cos^{2}\alpha_{x}\sigma_{0}^{x}B_{0}^{x}+\sin^{2}\alpha_{x}\sum\nolimits_{i=1}^{2}\sigma_{i}^{x}B_{i}^{x} (7)
+cos2αyσ0yB0y+sin2αyi=12σiyBiy.\displaystyle+\cos^{2}\alpha_{y}\sigma_{0}^{y}B_{0}^{y}+\sin^{2}\alpha_{y}\sum\nolimits_{i=1}^{2}\sigma_{i}^{y}B_{i}^{y}.

We observe that if [σ1z+σ2z,Tx,(y,z)]=0[\sigma^{z}_{1}+\sigma^{z}_{2},T_{x,(y,z)}]=0, these interactions can be eliminated by adding an LEO Hamiltonian Wu2002 ; Wang2020

HLEO=c(t)(σ1z+σ2z),\displaystyle H_{LEO}=c(t)(\sigma^{z}_{1}+\sigma^{z}_{2}), (8)

where c(t)c(t) is the control function. We emphasize that the significant advantage of the LEO Hamiltonian is that, due to [σ1z+σ2z,Tx,(y,z)]=0[\sigma^{z}_{1}+\sigma^{z}_{2},T_{x,(y,z)}]=0, adding of such an LEO does not interfere with the gate operation.

Refer to caption
Refer to caption
Figure 1: (Color on line) (a) The fidelity FF versus θ/π\theta/\pi for different parameter hh for logical qubits. The X,Y,ZX,Y,Z noises are all individual; (b) The fidelity FF versus θ/π\theta/\pi under DD control. θ=Jt\theta=Jt can be viewed as a rotation angle of the gates. The pulses are present at τ\tau and absent at the next τ\tau. Hs=JTxH_{s}=-JT_{x}, Γ=0.005\Gamma=0.005, γ=1\gamma=1, T=50T=50.
Refer to caption
Refer to caption
Refer to caption
Figure 2: (Color online) The number of gates NN (F=0.95F=0.95) versus α\alpha with different γ\gamma (a) or temperature TT (b). T=10T=10 in (a) and γ=2\gamma=2 in (b). Γ=0.005\Gamma=0.005. Hs=JTxH_{s}=-JT_{x}. (c) Comparison of the rotation angle θ\theta (F=0.95F=0.95) versus α\alpha between physical qubit and logical qubit with different Tx,(z)T_{x,(z)}. Γ=0.005\Gamma=0.005 (0.010.01) for the logical (physical) qubit. γ=10\gamma=10, T=50T=50, L=σzL=\sigma^{z}.

Suppose the control function is a constant c(t)=hc(t)=h, here hh, for example, could be the magnetic field. In Fig. 1 (a), we plot the fidelity FF as a function of the normalized time θ/π\theta/\pi for different hh. Here θ=Jt\theta=Jt can be viewed as a rotation angle and is taken to be θ=2π\theta=2\pi in the total evolution. We assume the collective bath has the same magnitude as the individual baths, i.e., Γ0=Γ1=Γ2=Γ\Gamma_{0}=\Gamma_{1}=\Gamma_{2}=\Gamma, γ0=γ1=γ2=γ\gamma_{0}=\gamma_{1}=\gamma_{2}=\gamma, and T0=T1=T2=TT_{0}=T_{1}=T_{2}=T. The environmental parameters are taken to be Γ=0.005\Gamma=0.005, γ=1\gamma=1, T=50T=50. If there is no control, the fidelity will decrease quickly with increasing θ\theta for Hs=JTxH_{s}=-JT_{x}. As expected, the X,YX,Y or X,Y,ZX,Y,Z noises will significantly reduce the gate fidelity.

We next compare this with control added. For simplicity, we only add a constant pulse. Noting h>>Jh>>J and that this control does not affect the gate operation. (They commute.) Fig. 1(a) shows that with the increasing hh, FF has a drastic increase. When h=20h=20, F=1F=1 for X,YX,Y noises, indicating the interaction has been removed. When it also has ZZ noise, the control is not as effective. We also check the case Hs=JTzH_{s}=-JT_{z} and find similar behavior to that in Fig. 1(a). The constant pulse in the above discussion is often difficult to approximate well in experiments. Bang-Bang are idealistic since they assume a relatively strong, fast pulse sequences c(t)=πiδ(tτi)/2c(t)=\pi\sum_{i}\delta(t-\tau_{i})/2 PyshkinSR . (They must be strong relative to the natural, or drift, Hamiltonian.) Thus these are also often difficult to be implement experimentally Jing2013 . Nonperturbative DD which uses a finite pulse intensity and finite pulse intervals is much more practical for effective control Jing2013 ; Wang2012 . Next we show how nonperturbative DD can be used to eliminate X,YX,Y noises.

In the numerical simulation, we use rectangular pulses and the above LEO Hamiltonian to simulate a δ\delta-function pulse. We use c(t)=50c(t)=50 (even nn) and c(t)=0c(t)=0 for nτ<t<(n+1)τn\tau<t<(n+1)\tau (odd nn), τ=0.01π\tau=0.01\pi. For this choice, the integral satisfies 0τc(s)𝑑s=π/2\int_{0}^{\tau}c(s)ds=\pi/2 which is required by theory in one control period τ\tau Wang2020 . However, there is no need to stick to π/2\pi/2, nonperturbative DD only requires a large constant Jing2013 ; Wang2012 , the control function can even be noisy Jing2014 ; Long . In Fig. 1(b), we plot the fidelity FF versus the parameter θ/π\theta/\pi under DD control. The results again show that the simulated DD pulses are effective to remove the X,YX,Y noises (αx=αy=π/2\alpha_{x}=\alpha_{y}=\pi/2 or αx=αy=π/4\alpha_{x}=\alpha_{y}=\pi/4 in Fig. 1(b)), both for individual and collective types. As expected, it fails for the case that the X,Y,ZX,Y,Z noises (αx=αy=αz=π/2\alpha_{x}=\alpha_{y}=\alpha_{z}=\pi/2) in Fig. 1(b)). We also plot the case where we have both the individual X,Y,ZX,Y,Z noises plus collective ZZ noise (αx=αy=π/2,αz=π/4\alpha_{x}=\alpha_{y}=\pi/2,\alpha_{z}=\pi/4). We find that in this case the protection of the DFS encodings is still effective and the fidelity evolution for these two cases are the same. To summarize, the advantage of our hybrid strategy is that if there are X,YX,Y noises and only collective ZZ noise, reliable quantum gate operation can be realized by adding a control that effectively removes the interaction and the effects of the noises can be completely eliminated.

From the above analysis, the individual ZZ noise is not easily eliminated. It is therefore important to check how the mixture of the individual and collective ZZ noises affect the number of the gates NN that can be implemented for certain threshold fidelity, e.g., F(θ)=0.95F(\theta)=0.95. Recently, a quantum error correction threshold of 4.7%4.7\% using a clustering decoder has been found for a depolarizing noise model Alexis . In the following part we will only discuss ZZ noise. Now for collective bath σ0=σ1z+σ2z\sigma_{0}=\sigma^{z}_{1}+\sigma^{z}_{2}, and for individual baths σ1=σ1z,σ2=σ2z\sigma_{1}=\sigma^{z}_{1},\sigma_{2}=\sigma^{z}_{2}. In Figs. 2(a) and (b), we plot the number of the gates NN versus α\alpha for different γ\gamma and TT, respectively. We take Hs=JTxH_{s}=-JT_{x} as an example, and for Hs=JTzH_{s}=-JT_{z} we obtain similar results. NN decreases with increasing parameter α\alpha, which shows that NN can be dramatically enhanced by increasing the collective bath ratio α\alpha. We emphasize that an increase in this ratio has been realized in recent work Wilen . Our results support that thousands of gates can be implemented for a small α\alpha. From Figs. 2(a), non-Markovinity of the baths play an important role in boosting the achievable number of gates NN. Fig. 2(b) shows that the NN decreases with increasing temperature TT as expected.

DFS encodings provide passive protection for collective noises. Then to what degree does the noise differ from that for the physical qubit? For a single physical qubit gate, the corresponding Hamiltonian is H=Jσx,(z)+σzB+HBH=J\sigma^{x,(z)}+\sigma^{z}B+H_{B}. In Fig. 2(c), we compare the achievable θ\theta for the physical qubit and the logical qubit. θ\theta versus α\alpha for TxT_{x} or TzT_{z} is plotted. γ=10\gamma=10, T=50T=50 and J=1J=1 are used for both cases, while we take Γ=0.005\Gamma=0.005 (1.0) for the logical (physical) qubit. For Tx,zT_{x,z}, there is only an individual bath for one single physical qubit. Then θ\theta is a constant. However, for the logical qubit, it has both collective and individual baths. For TxT_{x}, α=π/8\alpha=\pi/8, and θ/π=4\theta/\pi=4 and for TzT_{z}, θ/π=2.6\theta/\pi=2.6. Increasing α\alpha, θ\theta begins to decrease. When α=π/2\alpha=\pi/2, i.e., there are only individual baths for the logical qubit, we find that the dynamics are the same for the two cases: the interaction strength parameter 2Γ2\Gamma (physical bit) as Γ\Gamma (logical qubit).

Refer to caption

Figure 3: (Color online) Two logical qubits which consist of two pairs of spins. The solid line in the spin pair represents the XX+YYXX+YY coupling and the dashed line between the two pairs represents the ZZZZ coupling.

Now let us consider two pairs of spins, each pair encodes one logical qubit. Assume spins 1 and 2 (3 and 4) belong to the first (second) logical qubit (See Fig. 3). Controlled operations between two logical qubit can be made by Tz1Tz2=Z2Z3T_{z1}T_{z2}=-Z_{2}Z_{3}, which implements a two-qubit entangling gate. The operator eiθTz1Tz2e^{i\theta T_{z1}T_{z2}} gives a controlled phase gate when θ=π/2\theta=\pi/2. Together with a Hadamard gate, we can get a CNOT Pyshkin . We point out that the couplings between the spins can be tuned by adding an external field Jepsen . For the two logical qubits, the system Hamiltonian is Hs=JTz1Tz2H_{s}=-JT_{z1}T_{z2}. In Figs. 4(a) and (b) we plot the angle θ/π\theta/\pi versus γ\gamma and TT for the two-qubit gate. Fig. 4 shows that the obtainable angle θ\theta decreases with increasing γ\gamma and TT. For certain parameters, θ\theta also decreases with increasing α\alpha. The parameter dependence shows a behavior similar to the single-qubit gate.

Refer to caption
Refer to caption
Figure 4: (Color online) The rotation angle θ\theta (F=0.95F=0.95) versus γ\gamma (a) and temperature TT (b) with different α\alpha values. T=10T=10, Γ=0.005\Gamma=0.005 in (a) and γ=2\gamma=2, Γ=0.005\Gamma=0.005 in (b). Hs=JTz1Tz2H_{s}=-JT_{z1}T_{z2}.

IV Experimental parameters

The dimensionless parameters used in this paper can be converted into a dimensional form for a comparison with a recent experiment using superconducting qubits Yan19 . For the Hamiltonian of a single-qubit gate, each qubit can be regarded as a spin-1/2 system, in this case, the coupling JJ is the nearest-neighbor hopping strength. The typical coupling is around 12.512.5 MHz Yan19 and /J0.08\hbar/J\simeq 0.08 μs\mu s. The tunable zz-axis coupling between qubits can be performed with an additional intermediate qubit mode and the coupling strength is governed by the flux bias applied to the coupler chen2014qubit ; huang2020superconducting . The strength JzJ_{z} can be tuned to 1010 MHz in a high-coherence superconducting circuit Kounalakis , which is near the coupling strength JJ. The zz-axis rotations on individual qubits can be performed by modulating the microwave and local magnetic fields sc_review_1 . Increasing the collective bath ratio in the experiment Wilen can greatly enhance the number of gates that can be applied, enhancing the ability to perform computations.

V Conclusions

We have designed a hybrid error reduction method using very low overhead. Motivated by recent experiments, the method uses passive protection (a DFS) to reduce collective errors while employing active correction (LEOs) for individual noise. The calculation shows that high fidelity can be obtained for low temperature, and high non-Markovianity of the baths. This strategy shows promise theoretically, and numerically. We have also provided evidence that suggests significant improvement for experiments involving superconducting qubits.

VI ACKNOWLEDGMENTS

This work was supported by the Natural Science Foundation of Shandong Province (Grant No. ZR2021LLZ004) and Fundamental Research Funds for the Central Universities (Grant No. 202364008)., the grant PID2021-126273NB-I00 funded by MCIN/AEI/10.13039/501100011033, and by “ERDF A way of making Europe” and the Basque Government through grant number IT1470-22.

References

  • (1) P. W. Shor, SIAM J. Comput. 26,1484 (1997).
  • (2) P. W. Shor, Phys. Rev. A. 52, 2493 (1995).
  • (3) E. Knill, R. Laflamme, and W. H. Zurek, Science 279, 5349, 342 (1998).
  • (4) D. Gottesman, Phys. Rev. A. 57, 127 (1998).
  • (5) D. Lidar, T. Brun, Quantum error correction, Cambridge University Press (2013).
  • (6) J. Roffe, Contemp. Phys. 60, 226 (2019).
  • (7) J. Preskill, Quantum 2, 79 (2018).
  • (8) S. D. Sarma, https://www.technologyreview.com
    /2022/03/28/1048355/quantum-computing-has-a-hype-problem/
  • (9) P. Krantz, M. Kjaergaard, F. Yan, T. P. Orlando, S. Gustavsson, and W. D. Oliver, Appl. Phys. Rev. 6, 021318 (2019).
  • (10) L.-A. Wu and D. A. Lidar, Phys. Rev. Lett. 91, 097904 (2003).
  • (11) G. S. Uhrig, Phys. Rev. Lett. 102, 120502 (2009).
  • (12) Y. Sekiguchi, Y. Komura, and H. Kosaka, Phys. Rev. Applied 12, 051001 (2019).
  • (13) L.-M. Duan, and G.-C. Guo, Phys. Rev. Lett. 79, 1953 (1997).
  • (14) P. Zanardi and M. Rasetti, Phys. Rev. Lett. 79, 3306 (1997).
  • (15) K. R. Brown, J. Vala, and K. B. Whaley, Phys. Rev. A 67, 012309 (2003).
  • (16) J. Jing, L.-A. Wu, J. Q. You, and T. Yu, Phys. Rev. A 88, 022333 (2013).
  • (17) P. V. Pyshkin, D.-W. Luo, J. Jing, J. Q. You, and L.-A. Wu, Sci. Rep. 6, 37781 (2016).
  • (18) M. K. Henry, C. Ramanathan, J. S. Hodges, C. A. Ryan, M. J. Ditty, R. Laflamme, and D. G. Cory, Phys. Rev. Lett. 99, 220501 (2007).
  • (19) D. P. DiVincenzo, D. Bacon, J. Kempe, G. Burkard and K. B. Whaley, Nature (London) 408, 339 (2000).
  • (20) J. Kempe, D. Bacon, D. A. Lidar, and K. B. Whaley, Phys. Rev. A 63, 042307 (2001).
  • (21) L.-A. Wu, D. A. Lidar, Phys. Rev. A 65, 042318 (2002).
  • (22) D. A. Lidar, L.-A. Wu, Phys. Rev. Lett. 88, 017905 (2002).
  • (23) L.-A. Wu, Y.-X. Liu, F. Nori, Phys. Rev. A 80, 042315 (2009).
  • (24) L.-A. Wu, A. Miranowicz, X. Wang, Y.-x. Liu, and F. Nori,Phys. Rev. A 80, 012332 (2009).
  • (25) D. Kielpinski, V. Meyer, M. A. Rowe, C. A. Sackett, W. M. Itano, C. A. Monroe and D. J. Wineland, Science 291, 1013 (2001).
  • (26) M. Mohseni, J. S. Lundeen, K. J. Resch, and A. M. Steinberg, Phys. Rev. Lett. 91, 187903 (2003).
  • (27) P. G. Kwiat, A. J. Berglund, J. B. Altepeter, and A. G. White, Science 290, 5491, 498 (2000).
  • (28) E. M. Fortunato, L. Viola, J. Hodges, G. Teklemariam and D. G. Cory, New. J. Phys. 4, 5 (2002).
  • (29) C. D. Wilen, S. Abdullah, N. A. Kurinsky, C. Stanford, L. Cardani, G. D’Imperio, C. Tomei, L. Faoro, L. B. Ioffe, C. H. Liu, A. Opremcak, B. G. Christensen, J. L. DuBois and R. McDermott, Nature, 594, 369 (2021).
  • (30) P. V. Pyshkin, D.-W. Luo, L.-A. Wu, Phys. Rev. A 106, 012420 (2022).
  • (31) W. G. Unruh, Phys. Rev. A 51, 992 (1995).
  • (32) J. Jing, L.-A. Wu and A. del Campo, Sci. Rep. 6, 38149 (2016).
  • (33) L.-M. Duan and G.-C. Guo, Phys. Rev. A 57, 1050 (1998).
  • (34) D. D. Yavuz, J. Opt. Soc. Am. B 31, 2665 (2014).
  • (35) T. Schulte-Herbrüggen, A. Spörl, N. Khaneja and S. J. Glaser, J. Phys. B: At. Mol. Opt. Phys. 44, 154013 (2011).
  • (36) R. Alicki, K. Lendi, Quantum Dynamical Semigroups and Applications (Springer, Berlin, 1987).
  • (37) D. Chruściński, A. Kossakowski, and S. Pascazio. Phys. Rev. A 81, 032101 (2010).
  • (38) S. Wenderoth, H.-P. Breuer, and M. Thoss, Phys. Rev. A 107, 022211 (2023).
  • (39) L. Diósi, N. Gisin, and W. T. Strunz, Phys. Rev. A 58, 1699 (1998).
  • (40) W. T. Strunz, L. Diósi and N. Gisin, Phys. Rev. Lett. 82, 1801(1999).
  • (41) T. Yu, Phys. Rev. A 69, 062107 (2004).
  • (42) Z.-M. Wang, F.-H. Ren, D.-W. Luo, Z.-Y. Yan and L.-A. Wu, J. Phys. A: Math. Theor. 54, 155303 (2021).
  • (43) F.-H. Ren, Z.-M. Wang and L.-A. Wu, Phys. Rev. A. 102, 062603 (2020).
  • (44) W. T. Strunz, L. Diósi and N. Gisin, and Ting Yu Phys. Rev. Lett. 83, 4909 (1999).
  • (45) G. Ritschel and A. Eisfeld, J. Chem. Phys. 141, 094101 (2014).
  • (46) C. Meier and D. J. Tannor, J. Chem. Phys. 111, 3365 (1999).
  • (47) F. Intravaia, S. Maniscalco, and A. Messina, Phys. Rev. A 67, 042108 (2003).
  • (48) F. Intravaia, S. Maniscalco, and A. Messina, Eur. Phys. J. B. 32, 97 (2003).
  • (49) L.-A. Wu, M. S. Byrd, and D. A. Lidar, Phys. Rev. Lett. 89, 127901 (2002).
  • (50) Z.-M. Wang, F.-H. Ren, D.-W. Luo, Z.-Y. Yan, and L.-A. Wu, Phys. Rev. A. 102 , 042406 (2020).
  • (51) Z.-M. Wang, L.-A. Wu, J. Jing, B. Shao, and T. Yu, Phys. Rev. A. 86 , 032303 (2012).
  • (52) J. Jing, L.-A. Wu, T. Yu, J. Q. You, Z.-M. Wang, and L. Garcia, Phys. Rev. A. 89 , 032110 (2014).
  • (53) B.-X. Wang, T. Xin, X.-Y. Kong, S.-J. Wei, D. Ruan, and G.-L. Long, Phys. Rev. A. 97 , 042345 (2018).
  • (54) A. Schotte, G. Zhu, L. Burgelman, and F. Verstraete, Phys. Rev. X 12, 021012 (2022).
  • (55) P. N. Jepsen, J. Amato-Grill, I. Dimitrova, W. W. Ho, E. Demler and W. Ketterle, Nature 588, 403, (2020).
  • (56) G. H. Yuan and N. I. Zheludev, Science, 364, 6442, 771 (2019).
  • (57) Y. Chen, C. Neill, P. Roushan, N. Leung, M. Fang et al, Phys. Rev. Lett. 113, 220502 (2014).
  • (58) H.-L. Huang, D. Wu, D. Fan, and X. Zhu, Sci. China Inf. Sci., 63, 8 (2020).
  • (59) M. Kounalakis, C. Dickel, A. Bruno, N. K. Langford, and G. A. Steele, npj Quantum Inf. 4, 38 (2018).