Dynamics of qudit gates and effects of spectator modes on
optimal control pulses

A. Barış Özgüler [email protected] Fermi National Accelerator Laboratory, Batavia, IL, 60510 Superconducting Quantum Materials and Systems Center (SQMS), Fermilab    Joshua A. Job Lockheed Martin Advanced Technology Center, Sunnyvale, CA, 94089 Superconducting Quantum Materials and Systems Center (SQMS), Fermilab
(March 25, 2024)
Abstract

Qudit gates for high-dimensional quantum computing can be synthesized with high precision using numerical quantum optimal control techniques. Large circuits are broken down into modules and the tailored pulses for each module can be used as primitives for a qudit compiler. Application of the pulses of each module in the presence of extra modes may decrease their effectiveness due to crosstalk. In this paper, we address this problem by simulating qudit dynamics for circuit quantum electrodynamics (cQED) systems. As a test case, we take pulses for single-qudit SWAP gates optimized in isolation and then apply them in the presence of spectator modes each of which are in Fock states. We provide an experimentally relevant scaling formula that can be used as a bound on the fidelity decay. Our results show that frequency shift from spectator mode populations has to be 0.1%less-than-or-similar-toabsentpercent0.1\lesssim 0.1\%≲ 0.1 % of the qudit’s nonlinearity in order for high-fidelity single-qudit gates to be useful in the presence of occupied spectator modes.

I Introduction

With the demonstrations of qudit control in quantum devices, such as trapped ions [1], photonic processors [2], and circuit quantum electrodynamics (cQED) systems [3, 4, 5, 6, 7], many computational levels can be successfully manipulated in order to design and execute quantum algorithms [8]. Compared to its qubit counterparts, high-dimensional quantum computing has many advantages, some of which are lower-depth circuits, noise improvement with hardware-efficient solutions [8, 9, 10, 11] and efficient means for large-scale quantum information experiments to be performed in the lab, such as black hole dynamics modeled as a scrambling unitary [12].

Quantum devices can be controlled optimally via external fields [13, 14, 15]. Gates can be designed in modules (1- and 2-qudit gates), such as in Ref. [16] for bosonic modes. To be able to use synthesized gates in the entire space by preserving their fidelity, one needs to check if the modules function across the entire space. We leverage Juqbox.jl [17] to synthesize qudit SWAP gates with B-spline parametrization following the techniques in [18, 19]. SWAP operations provide simple, yet effective demonstrations for the effects of frequency shifts, which alter the ideal transitions between energy levels and cause fidelity decay.

We outline the rest of the paper. In Section II, we provide the effective Hamiltonian of the driven qudit when it interacts with spectator modes, each of which are in Fock states. In Section III, the infidelity scaling is given analytically and compared with the numerical result. Finally, in Section IV, we conclude the paper by discussing future work, including ways to alleviate the fidelity decay.

II Effective Hamiltonian and frequency shift in the presence of spectator modes

We focus here on a cQED system with many oscillators/modes. The system Hamiltonian in the rotating frame for each oscillator is given by [13, 14]:

H=iξi2(n^in^in^i)j>iξijn^in^j,𝐻subscript𝑖subscript𝜉𝑖2subscript^𝑛𝑖subscript^𝑛𝑖subscript^𝑛𝑖subscript𝑗𝑖subscript𝜉𝑖𝑗subscript^𝑛𝑖subscript^𝑛𝑗H=-\sum_{i}\frac{\xi_{i}}{2}(\hat{n}_{i}\hat{n}_{i}-\hat{n}_{i})-\sum_{j>i}\xi% _{ij}\hat{n}_{i}\hat{n}_{j},italic_H = - ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ( over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - ∑ start_POSTSUBSCRIPT italic_j > italic_i end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (1)

where ξisubscript𝜉𝑖\xi_{i}italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the self-Kerr for each oscillator i𝑖iitalic_i, and ξijsubscript𝜉𝑖𝑗\xi_{ij}italic_ξ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the cross-Kerr between oscillators i𝑖iitalic_i and j𝑗jitalic_j. If we take the state at time t𝑡titalic_t to be a product state of the form |ψ|jnjtensor-productket𝜓ketsubscriptproduct𝑗subscript𝑛𝑗\ket{\psi}\otimes\ket{\prod_{j}n_{j}}| start_ARG italic_ψ end_ARG ⟩ ⊗ | start_ARG ∏ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩, with the state on the target oscillator |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩ and spectator modes in Fock states {nj}subscript𝑛𝑗\{n_{j}\}{ italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT }, it is easy to see that the action of the system Hamiltonian will be

H|ψj|nj=[ξ12(n^1n^1n^1)j>1ξ1jnjn^1+C]|ψj|jnjsubscripttensor-product𝑗𝐻ket𝜓ketsubscript𝑛𝑗subscripttensor-product𝑗delimited-[]subscript𝜉12subscript^𝑛1subscript^𝑛1subscript^𝑛1subscript𝑗1subscript𝜉1𝑗subscript𝑛𝑗subscript^𝑛1𝐶ket𝜓ketsubscriptproduct𝑗subscript𝑛𝑗\begin{split}H\ket{\psi}\otimes_{j}\ket{n_{j}}&=\Bigl{[}-\frac{\xi_{1}}{2}(% \hat{n}_{1}\hat{n}_{1}-\hat{n}_{1})\\ &\quad-\sum_{j>1}\xi_{1j}n_{j}\hat{n}_{1}+C\Bigr{]}\ket{\psi}\otimes_{j}\ket{% \prod_{j}n_{j}}\end{split}start_ROW start_CELL italic_H | start_ARG italic_ψ end_ARG ⟩ ⊗ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | start_ARG italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ end_CELL start_CELL = [ - divide start_ARG italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ( over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL - ∑ start_POSTSUBSCRIPT italic_j > 1 end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT 1 italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_C ] | start_ARG italic_ψ end_ARG ⟩ ⊗ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | start_ARG ∏ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ end_CELL end_ROW (2)

where C𝐶Citalic_C is a constant formed by the action of H𝐻Hitalic_H on the spectator modes which we ignore from here on as it only generates a global phase.

The above Hamiltonian does not generate any evolution on the spectator modes (because their initial state is a Fock state), and so focusing only on the target mode and suppressing the subscript 1111 for ease of notation we get the effective Hamiltonian on the target mode:

Heff=ξ2(n^n^n^)jξjnjn^,subscript𝐻𝑒𝑓𝑓𝜉2^𝑛^𝑛^𝑛subscript𝑗subscript𝜉𝑗subscript𝑛𝑗^𝑛H_{eff}=-\frac{\xi}{2}(\hat{n}\hat{n}-\hat{n})-\sum_{j}\xi_{j}n_{j}\hat{n},italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = - divide start_ARG italic_ξ end_ARG start_ARG 2 end_ARG ( over^ start_ARG italic_n end_ARG over^ start_ARG italic_n end_ARG - over^ start_ARG italic_n end_ARG ) - ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG , (3)

where the first term H0ξ2(n^n^n^)subscript𝐻0𝜉2^𝑛^𝑛^𝑛H_{0}\equiv-\frac{\xi}{2}(\hat{n}\hat{n}-\hat{n})italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≡ - divide start_ARG italic_ξ end_ARG start_ARG 2 end_ARG ( over^ start_ARG italic_n end_ARG over^ start_ARG italic_n end_ARG - over^ start_ARG italic_n end_ARG ) is the time-independent part of the driven qudit Hamiltonian, εjξjnj𝜀subscript𝑗subscript𝜉𝑗subscript𝑛𝑗\varepsilon\equiv\sum_{j}\xi_{j}n_{j}italic_ε ≡ ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the perturbation parameter and Vn^𝑉^𝑛V\equiv-\hat{n}italic_V ≡ - over^ start_ARG italic_n end_ARG is the shift operator appearing due to spectators. In essence, the cross-Kerr between the target mode and the spectator modes in Fock states produces a frequency shift on the target mode. We can write in short:

Heff=H0+εV.subscript𝐻𝑒𝑓𝑓subscript𝐻0𝜀𝑉H_{eff}=H_{0}+\varepsilon V.italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_ε italic_V . (4)
Refer to caption
Figure 1: (Top) Infidelity for the labeled SWAP operation arising from a frequency shift ε=jξjnj𝜀subscript𝑗subscript𝜉𝑗subscript𝑛𝑗\varepsilon=\sum_{j}\xi_{j}n_{j}italic_ε = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT from the presence of njsubscript𝑛𝑗n_{j}italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT photons in the j𝑗jitalic_jth spectator mode with cross-Kerr strength to the target mode ξjsubscript𝜉𝑗\xi_{j}italic_ξ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT relative to the ideal/target gate. We exclude zero occupation as the x-axis value would be 00, but the infidelity for that case is that of the optimal control pulse without spectator modes, namely (104)ordersuperscript104\order{10^{-4}}( start_ARG 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT end_ARG ) to (103)ordersuperscript103\order{10^{-3}}( start_ARG 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) for each gate, as seen at the smallest ε𝜀\varepsilonitalic_ε. The slope for small ε/ξ=(104)𝜀𝜉ordersuperscript104\varepsilon/\xi=\order{10^{-4}}italic_ε / italic_ξ = ( start_ARG 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT end_ARG ) is 2absent2\approx 2≈ 2, meaning that infidelity scales quadratically with ε𝜀\varepsilonitalic_ε. The flat region at very small ε𝜀\varepsilonitalic_ε is the region when the perturbation is negligible, while for larger ε𝜀\varepsilonitalic_ε higher order terms (in part due to saturation near infidelity 1absent1\approx 1≈ 1) take effect. (Bottom) Rescaled fidelity curves such that the value at the ε=104𝜀superscript104\varepsilon=10^{-4}italic_ε = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT for each curve is equal, so as to highlight the similarity of the slope in that region.

III Scaling of the infidelity

For the quantum control problem of gate synthesis, the target action U𝑈Uitalic_U is known, and we wish to find a time-dependent drive Hamiltonian Hd(t)subscript𝐻𝑑𝑡H_{d}(t)italic_H start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_t ) such that the evolution from H(t)=H0+Hd(t)𝐻𝑡subscript𝐻0subscript𝐻𝑑𝑡H(t)=H_{0}+H_{d}(t)italic_H ( italic_t ) = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_t ) produces the target unitary. The propagator for the driven qudit without spectator modes is given by:

U0(t)=𝒯exp[i0t𝑑t(H0+Hd(t))],subscript𝑈0𝑡𝒯𝑖Planck-constant-over-2-pisuperscriptsubscript0𝑡differential-dsuperscript𝑡subscript𝐻0subscript𝐻𝑑superscript𝑡\displaystyle U_{0}(t)=\mathcal{T}\exp\left[-\frac{i}{\hbar}\int_{0}^{t}dt^{% \prime}\left(H_{0}+H_{d}(t^{\prime})\right)\right],italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) = caligraphic_T roman_exp [ - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ) ] , (5)

where Hdsubscript𝐻𝑑H_{d}italic_H start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is the drive term synthesizing the target gate, i.e., U0(t=T)subscript𝑈0𝑡𝑇U_{0}(t=T)italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t = italic_T ) is the target gate U𝑈Uitalic_U, T𝑇Titalic_T is the gate time and 𝒯𝒯\mathcal{T}caligraphic_T is the time-ordering operator. To design a gate using optimal control techniques, we optimize coefficients of pulse control terms a^+a^^𝑎superscript^𝑎\hat{a}+\hat{a}^{\dagger}over^ start_ARG italic_a end_ARG + over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT and a^a^^𝑎superscript^𝑎\hat{a}-\hat{a}^{\dagger}over^ start_ARG italic_a end_ARG - over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT acting on the target oscillator. Thus, they commute with the spectator modes. Using the same drive terms, the effective propagator of the qudit due to spectator shifts is given by:

Ueff(t)=𝒯exp[i0t𝑑t(Heff+Hd(t))].subscript𝑈𝑒𝑓𝑓𝑡𝒯𝑖Planck-constant-over-2-pisuperscriptsubscript0𝑡differential-dsuperscript𝑡subscript𝐻𝑒𝑓𝑓subscript𝐻𝑑superscript𝑡\displaystyle U_{eff}(t)=\mathcal{T}\exp\left[-\frac{i}{\hbar}\int_{0}^{t}dt^{% \prime}\left(H_{eff}+H_{d}(t^{\prime})\right)\right].italic_U start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ( italic_t ) = caligraphic_T roman_exp [ - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_H start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ) ] . (6)

Here, we assume that the drive field is not coupled to the spectator modes, which is a safe assumption for sufficiently detuned frequencies.

Fidelity between the ideal gate and shifted gate is defined as:

F|Tr(Ulog(T))d|2,𝐹superscript𝑇𝑟subscript𝑈𝑙𝑜𝑔𝑇𝑑2F\equiv\Bigg{|}\frac{Tr(U_{log}(T))}{d}\Bigg{|}^{2},italic_F ≡ | divide start_ARG italic_T italic_r ( italic_U start_POSTSUBSCRIPT italic_l italic_o italic_g end_POSTSUBSCRIPT ( italic_T ) ) end_ARG start_ARG italic_d end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (7)

where Ulog(t)U0(t)Ueff(t)subscript𝑈𝑙𝑜𝑔𝑡superscriptsubscript𝑈0𝑡subscript𝑈𝑒𝑓𝑓𝑡U_{log}(t)\equiv U_{0}^{\dagger}(t)\,U_{eff}(t)italic_U start_POSTSUBSCRIPT italic_l italic_o italic_g end_POSTSUBSCRIPT ( italic_t ) ≡ italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_t ) italic_U start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT ( italic_t ) is the propagator in the logical frame of the driven qudit 111For an in-depth discussion on the concepts of the logical frame and the echo operator, please refer to Eqs. (294-295) in Ref. [21], which corresponds to Eqs. (8-9) in this manuscript., d𝑑ditalic_d is the norm of U0subscript𝑈0U_{0}italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

Ulogsubscript𝑈𝑙𝑜𝑔U_{log}italic_U start_POSTSUBSCRIPT italic_l italic_o italic_g end_POSTSUBSCRIPT is defined in terms of the perturbation as:

Ulog(t)=𝒯exp[i0t𝑑tV~(t)],subscript𝑈𝑙𝑜𝑔𝑡𝒯𝑖Planck-constant-over-2-pisuperscriptsubscript0𝑡differential-dsuperscript𝑡~𝑉superscript𝑡\displaystyle U_{log}(t)=\mathcal{T}\exp\left[-\frac{i}{\hbar}\int_{0}^{t}dt^{% \prime}\tilde{V}(t^{\prime})\right],italic_U start_POSTSUBSCRIPT italic_l italic_o italic_g end_POSTSUBSCRIPT ( italic_t ) = caligraphic_T roman_exp [ - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT over~ start_ARG italic_V end_ARG ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] , (8)

where V~(t)U0(t)VU0(t)~𝑉𝑡superscriptsubscript𝑈0𝑡𝑉subscript𝑈0𝑡\tilde{V}(t)\equiv U_{0}^{\dagger}(t)VU_{0}(t)over~ start_ARG italic_V end_ARG ( italic_t ) ≡ italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_t ) italic_V italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ). For small perturbation ε𝜀\varepsilonitalic_ε, Ulogsubscript𝑈𝑙𝑜𝑔U_{log}italic_U start_POSTSUBSCRIPT italic_l italic_o italic_g end_POSTSUBSCRIPT is expanded via Baker–Campbell–Hausdorff formula as in Ref. [21]:

Ulog(t)exp[i(εV¯t+12ε2Γ(t)+𝒪(ε3))],similar-to-or-equalssubscript𝑈𝑙𝑜𝑔𝑡𝑖Planck-constant-over-2-pi𝜀¯𝑉𝑡12superscript𝜀2Γ𝑡𝒪superscript𝜀3\displaystyle U_{log}(t)\simeq\ \exp\left[-\frac{i}{\hbar}\Bigl{(}\varepsilon% \bar{V}t+\frac{1}{2}\varepsilon^{2}\Gamma(t)+\mathcal{O}(\varepsilon^{3})\Bigr% {)}\right],italic_U start_POSTSUBSCRIPT italic_l italic_o italic_g end_POSTSUBSCRIPT ( italic_t ) ≃ roman_exp [ - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG ( italic_ε over¯ start_ARG italic_V end_ARG italic_t + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Γ ( italic_t ) + caligraphic_O ( italic_ε start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) ) ] , (9)

where V¯¯𝑉\bar{V}over¯ start_ARG italic_V end_ARG is the time average of V~(t)~𝑉𝑡\tilde{V}(t)over~ start_ARG italic_V end_ARG ( italic_t ):

V¯(t)=1T0tV~(t)𝑑t,¯𝑉𝑡1𝑇superscriptsubscript0𝑡~𝑉superscript𝑡differential-dsuperscript𝑡\bar{V}(t)=\frac{1}{T}\int_{0}^{t}\tilde{V}(t^{\prime})dt^{\prime},over¯ start_ARG italic_V end_ARG ( italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_T end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT over~ start_ARG italic_V end_ARG ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , (10)

and Γ(t)Γ𝑡\Gamma(t)roman_Γ ( italic_t ) is the integral of the time correlation function:

Γ(t)=i0t𝑑ttt𝑑t′′[V~(t),V~(t′′)].Γ𝑡𝑖Planck-constant-over-2-pisuperscriptsubscript0𝑡differential-dsuperscript𝑡superscriptsubscriptsuperscript𝑡𝑡differential-dsuperscript𝑡′′~𝑉superscript𝑡~𝑉superscript𝑡′′\Gamma(t)=\frac{i}{\hbar}\int_{0}^{t}dt^{\prime}\int_{t^{\prime}}^{t}dt^{% \prime\prime}[\tilde{V}(t^{\prime}),\tilde{V}(t^{\prime\prime})].roman_Γ ( italic_t ) = divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_d italic_t start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT [ over~ start_ARG italic_V end_ARG ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , over~ start_ARG italic_V end_ARG ( italic_t start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ) ] . (11)

Ulog(t)subscript𝑈𝑙𝑜𝑔𝑡U_{log}(t)italic_U start_POSTSUBSCRIPT italic_l italic_o italic_g end_POSTSUBSCRIPT ( italic_t ) is:

Ulog(t)I+X+X22+,similar-to-or-equalssubscript𝑈𝑙𝑜𝑔𝑡𝐼𝑋superscript𝑋22U_{log}(t)\simeq\ I+X+\frac{X^{2}}{2}+...,italic_U start_POSTSUBSCRIPT italic_l italic_o italic_g end_POSTSUBSCRIPT ( italic_t ) ≃ italic_I + italic_X + divide start_ARG italic_X start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG + … , (12)

where I𝐼Iitalic_I is the identity matrix and Xi(εV¯t+12ε2Γ(t))𝑋𝑖Planck-constant-over-2-pi𝜀¯𝑉𝑡12superscript𝜀2Γ𝑡X\equiv-\frac{i}{\hbar}\Bigl{(}\varepsilon\bar{V}t+\frac{1}{2}\varepsilon^{2}% \Gamma(t)\Bigr{)}italic_X ≡ - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG ( italic_ε over¯ start_ARG italic_V end_ARG italic_t + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Γ ( italic_t ) ). In simple terms, Equations 8 and 9, describe the concept of dynamical decoupling. They could be used to manipulate the reference frame for a quantum system to isolate and control specific types of interactions, with the goal of preserving the system’s quantum state against disturbances, which is crucial for the functioning of quantum computers and error correction in quantum information processing.

Fidelity is then expressed as (suppressing the time parameter of V¯(t=T)¯𝑉𝑡𝑇\bar{V}(t=T)over¯ start_ARG italic_V end_ARG ( italic_t = italic_T ) for notational simplicity):

F 1[Tr(V¯2)Tr2(V¯)]T22d2ε2.similar-to-or-equals𝐹1delimited-[]𝑇𝑟superscript¯𝑉2𝑇superscript𝑟2¯𝑉superscript𝑇2superscriptPlanck-constant-over-2-pi2superscript𝑑2superscript𝜀2F\simeq\ 1-\frac{\Bigl{[}Tr(\bar{V}^{2})-Tr^{2}(\bar{V})\Bigr{]}T^{2}}{\hbar^{% 2}\,d^{2}}\varepsilon^{2}.italic_F ≃ 1 - divide start_ARG [ italic_T italic_r ( over¯ start_ARG italic_V end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) - italic_T italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( over¯ start_ARG italic_V end_ARG ) ] italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (13)

The normalized trace term in the coefficient of ε2superscript𝜀2\varepsilon^{2}italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the variance of V¯¯𝑉\bar{V}over¯ start_ARG italic_V end_ARG for the maximally mixed state and the coefficient of ε2superscript𝜀2\varepsilon^{2}italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is known as fidelity susceptibility for time-independent systems [22, 23]. We leave the detailed examination of this trace term with time-averaged operators for future work. We compare this analytical scaling (ε2similar-toabsentsuperscript𝜀2\sim\varepsilon^{2}∼ italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) with numerical results below (cf.𝑐𝑓cf.italic_c italic_f . Fig. 1).

Transitions between the Fock states |iket𝑖\ket{i}| start_ARG italic_i end_ARG ⟩ and |jket𝑗\ket{j}| start_ARG italic_j end_ARG ⟩ in the oscillator are generated by control pulses at the transition frequency between the states, in our case that is ξ2(i2j2+ij)𝜉2superscript𝑖2superscript𝑗2𝑖𝑗\frac{\xi}{2}(i^{2}-j^{2}+i-j)divide start_ARG italic_ξ end_ARG start_ARG 2 end_ARG ( italic_i start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_j start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_i - italic_j ). The spectator modes, however, shift these frequencies by kξknk(ij)subscript𝑘subscript𝜉𝑘subscript𝑛𝑘𝑖𝑗\sum_{k}\xi_{k}n_{k}(i-j)∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_i - italic_j ). To demonstrate the effect of this frequency shift, we can optimize a set of control pulses to produce a swap gate between |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ and |jket𝑗\ket{j}| start_ARG italic_j end_ARG ⟩ on a spectator mode. For concreteness, we use ω/2π=4.8𝜔2𝜋4.8\omega/2\pi=4.8italic_ω / 2 italic_π = 4.8 GHz and ξ/2π=0.22𝜉2𝜋0.22\xi/2\pi=0.22italic_ξ / 2 italic_π = 0.22 GHz, with the self-Kerr of the spectator modes being modulated as some fraction of ξ𝜉\xiitalic_ξ and cross-Kerr parameters equal to βjξsubscript𝛽𝑗𝜉\beta_{j}\,\xiitalic_β start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ξ with parameter βjsubscript𝛽𝑗\beta_{j}italic_β start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT varying for each mode j𝑗jitalic_j. We use these system parameters so that our gates are directly comparable to those in Section 7 of Ref. [18]. Other parameters, such as which SWAPs will be generated (SWAPs from |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ to |3,|4,|5,|6ket3ket4ket5ket6\ket{3},\ket{4},\ket{5},\ket{6}| start_ARG 3 end_ARG ⟩ , | start_ARG 4 end_ARG ⟩ , | start_ARG 5 end_ARG ⟩ , | start_ARG 6 end_ARG ⟩) and the time for each gate (140,215,265,140215265140,215,265,140 , 215 , 265 , and 425425425425 ns respectively) are also taken from that section, along with the use of a single guard level (which implies that a SWAP to state |kket𝑘\ket{k}| start_ARG italic_k end_ARG ⟩ has k+1𝑘1k+1italic_k + 1 levels actively participating in the gate and k+2𝑘2k+2italic_k + 2 states simulated in the optimization and frequency-shifted calculations). Our only difference is that we restricted our optimization of the control parameters for the ideal (without spectator modes) case to 200 iterations. Note that our simulations were performed for closed systems but decoherence is not a bottleneck for this work since pulse durations are much shorter than typical coherence times for cQED systems, such as superconducting qubits and cavities [24, 3, 25, 7, 26, 6].

A SWAP gate between level |iket𝑖\ket{i}| start_ARG italic_i end_ARG ⟩ and |jket𝑗\ket{j}| start_ARG italic_j end_ARG ⟩ is defined as:

SWAP|i|j=I+|ij|+|ji||ii||jj|.subscriptSWAPket𝑖ket𝑗𝐼𝑖𝑗𝑗𝑖𝑖𝑖𝑗𝑗\text{SWAP}_{\ket{i}\longleftrightarrow\ket{j}}=I+\outerproduct{i}{j}+% \outerproduct{j}{i}-\outerproduct{i}{i}-\outerproduct{j}{j}.SWAP start_POSTSUBSCRIPT | start_ARG italic_i end_ARG ⟩ ⟷ | start_ARG italic_j end_ARG ⟩ end_POSTSUBSCRIPT = italic_I + | start_ARG italic_i end_ARG ⟩ ⟨ start_ARG italic_j end_ARG | + | start_ARG italic_j end_ARG ⟩ ⟨ start_ARG italic_i end_ARG | - | start_ARG italic_i end_ARG ⟩ ⟨ start_ARG italic_i end_ARG | - | start_ARG italic_j end_ARG ⟩ ⟨ start_ARG italic_j end_ARG | . (14)

SWAP gates are vital for shifting matrix elements around and moving quantum states around lattices of qubits, while partial SWAP operations can generate entanglement and more complicated superpositions. Here, our choice of simple SWAPs between the ground state and various excited states of the single oscillator is meant as only an example to illustrate the effect of spectator mode shift of the target oscillator’s transition frequencies.

The action on the system Hamiltonian from the spectator modes is conveyed entirely through the term εV^=n^jξjnj𝜀^𝑉^𝑛subscript𝑗subscript𝜉𝑗subscript𝑛𝑗\varepsilon\hat{V}=-\hat{n}\sum_{j}\xi_{j}n_{j}italic_ε over^ start_ARG italic_V end_ARG = - over^ start_ARG italic_n end_ARG ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. Instead of plotting the infidelity as a function of the populations in adjacent modes, we instead plot it against the parameter ε𝜀\varepsilonitalic_ε (really, ε/ξ𝜀𝜉\varepsilon/\xiitalic_ε / italic_ξ as this is the primary dynamically relevant parameter) in Fig. 1 for each of the four SWAP gates tested. Here we exclude the zero spectator mode photon occupation case (as ε=0𝜀0\varepsilon=0italic_ε = 0 in that case and thus cannot be placed on a log-log plot). We will simply note that infidelity is (104)ordersuperscript104\order{10^{-4}}( start_ARG 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT end_ARG ) to (103)ordersuperscript103\order{10^{-3}}( start_ARG 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) in the zero-noise case for each sample. We see that each SWAP gate tested shows the same scaling with ε/ξ𝜀𝜉\varepsilon/\xiitalic_ε / italic_ξ for ε/ξ0.001much-less-than𝜀𝜉0.001\varepsilon/\xi\ll 0.001italic_ε / italic_ξ ≪ 0.001, scaling with a slope of 2absent2\approx 2≈ 2 on a log-log plot, denoting quadratic scaling in ε𝜀\varepsilonitalic_ε, just as predicted in Section III. We note that these plots compare the implemented gate with spectator mode state-dependent frequency shifts to the ideal/target gate, and thus the infidelity is lower-bounded by the infidelity of the noiseless gate (resulting in a saturation behavior at small shifts).

To make this even clearer, we also plot a rescaling of these infidelities in Fig. 1, with each curve’s y-values rescaled such that at a data point for an intermediate value of ε𝜀\varepsilonitalic_ε (ε=104𝜀superscript104\varepsilon=10^{-4}italic_ε = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT) each curve has the same y-axis value. All the data lines up nearly perfectly for more than an order of magnitude from just over 105superscript10510^{-5}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT to around 103.5superscript103.510^{-3.5}10 start_POSTSUPERSCRIPT - 3.5 end_POSTSUPERSCRIPT and only diverges as ε/ξ𝜀𝜉\varepsilon/\xiitalic_ε / italic_ξ approaches 0.0010.0010.0010.001. We also provide a table, Table 1, showing the slope of the infidelity curves on the log-log plot in the region of 104superscript10410^{-4}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, estimated by taking the slope in the region between 104.05superscript104.0510^{-4.05}10 start_POSTSUPERSCRIPT - 4.05 end_POSTSUPERSCRIPT and 103.95superscript103.9510^{-3.95}10 start_POSTSUPERSCRIPT - 3.95 end_POSTSUPERSCRIPT, and all of the slopes are near 2222, ie are approximately quadratic.

SWAP gate infidelity slope
|0|3ket0ket3\ket{0}\longleftrightarrow\ket{3}| start_ARG 0 end_ARG ⟩ ⟷ | start_ARG 3 end_ARG ⟩ 1.95
|0|4ket0ket4\ket{0}\longleftrightarrow\ket{4}| start_ARG 0 end_ARG ⟩ ⟷ | start_ARG 4 end_ARG ⟩ 1.98
|0|5ket0ket5\ket{0}\longleftrightarrow\ket{5}| start_ARG 0 end_ARG ⟩ ⟷ | start_ARG 5 end_ARG ⟩ 1.94
|0|6ket0ket6\ket{0}\longleftrightarrow\ket{6}| start_ARG 0 end_ARG ⟩ ⟷ | start_ARG 6 end_ARG ⟩ 1.84
Table 1: Slope in the region of ε/ξ=104𝜀𝜉superscript104\varepsilon/\xi=10^{-4}italic_ε / italic_ξ = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT of the infidelity curves to three significant figures for each SWAP gate tested, found by taking the slope of said curve in the region 104.05<ε/ξ<103.95superscript104.05𝜀𝜉superscript103.9510^{-4.05}<\varepsilon/\xi<10^{-3.95}10 start_POSTSUPERSCRIPT - 4.05 end_POSTSUPERSCRIPT < italic_ε / italic_ξ < 10 start_POSTSUPERSCRIPT - 3.95 end_POSTSUPERSCRIPT.

Thus, we see both from theory and simulation that the effect of spectator modes on the fidelity of a gate generated by a control pulse produced without taking into account spectator modes’ frequency shift on the target mode is approximately quadratic in the magnitude of that frequency shift, and rises rapidly to yield an almost orthogonal gate for shifts on the order of 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT times the qudit nonlinearity.

IV Conclusions

We provided a fidelity decay formula and simulated qudit gates in the presence of spectator modes in order to compare the estimated scaling and numerical results. The fidelity formula, Eq. 13, is independent of the gate, so we expect to get a similar scaling for ε0𝜀0\varepsilon\rightarrow 0italic_ε → 0 for gates other than SWAP. In our study, a “useful” gate (as mentioned in the abstract) is defined as one that maintains an operational fidelity exceeding 99.9%, a threshold crucial for fault-tolerant quantum computing. This high fidelity standard ensures the gates are sufficiently reliable for practical quantum computing applications, where precision and error minimization are essential.

We highlight that these frequency shifts yield extremely stringent bounds on interaction parameters and spectator mode occupations. For future directions, one may try to tackle alleviating the effects of fidelity decay with several useful approaches from quantum computing and error correction, such as dynamical decoupling [21, 27, 28], shortcuts to adiabaticity and steering [29], circuit optimization and machine learning [30, 31], risk-neutral approaches in robust control [19] and bosonic error correction [32].

Acknowledgment

We thank Jens Koch and Yuri Alexeev for discussions on large-scale simulations. This material is based upon work supported by the U.S. Department of Energy, Office of Science, National Quantum Information Science Research Centers, Superconducting Quantum Materials and Systems Center (SQMS) under contract number DE-AC02-07CH11359. We gratefully acknowledge the computing resources provided on Bebop, a high-performance computing cluster operated by the Laboratory Computing Resource Center at Argonne National Laboratory.

References

  • Ringbauer et al. [2021] M. Ringbauer, M. Meth, L. Postler, R. Stricker, R. Blatt, P. Schindler, and T. Monz, A universal qudit quantum processor with trapped ions, arXiv preprint arXiv:2109.06903  (2021).
  • Chi et al. [2022] Y. Chi, J. Huang, Z. Zhang, J. Mao, Z. Zhou, X. Chen, C. Zhai, J. Bao, T. Dai, H. Yuan, et al., A programmable qudit-based quantum processor, Nature communications 13, 1 (2022).
  • Romanenko et al. [2020] A. Romanenko, R. Pilipenko, S. Zorzetti, D. Frolov, M. Awida, S. Belomestnykh, S. Posen, and A. Grassellino, Three-dimensional superconducting resonators at t𝑡titalic_t <<< 20 mk with photon lifetimes up to τ𝜏\tauitalic_τ= 2 s, Physical Review Applied 13, 034032 (2020).
  • Wu et al. [2020] X. Wu, S. Tomarken, N. A. Petersson, L. Martinez, Y. J. Rosen, and J. L. DuBois, High-fidelity software-defined quantum logic on a superconducting qudit, Physical Review Letters 125, 170502 (2020).
  • Alam et al. [2022] M. S. Alam, S. Belomestnykh, N. Bornman, G. Cancelo, Y.-C. Chao, M. Checchin, V. S. Dinh, A. Grassellino, E. J. Gustafson, R. Harnik, et al., Quantum computing hardware for hep algorithms and sensing, arXiv preprint arXiv:2204.08605  (2022).
  • Chakram et al. [2022] S. Chakram, K. He, A. V. Dixit, A. E. Oriani, R. K. Naik, N. Leung, H. Kwon, W.-L. Ma, L. Jiang, and D. I. Schuster, Multimode photon blockade, Nature Physics , 1 (2022).
  • Chakram et al. [2021] S. Chakram, A. E. Oriani, R. K. Naik, A. V. Dixit, K. He, A. Agrawal, H. Kwon, and D. I. Schuster, Seamless high-q microwave cavities for multimode circuit quantum electrodynamics, Physical review letters 127, 107701 (2021).
  • Wang et al. [2020] Y. Wang, Z. Hu, B. C. Sanders, and S. Kais, Qudits and high-dimensional quantum computing, Frontiers in Physics 8, 479 (2020).
  • Gustafson [2021] E. J. Gustafson, Prospects for simulating a qudit-based model of (1+ 1) d scalar qed, Physical Review D 103, 114505 (2021).
  • Gustafson [2022] E. Gustafson, Noise improvements in quantum simulations of sqed using qutrits, arXiv preprint arXiv:2201.04546  (2022).
  • Otten et al. [2021] M. Otten, K. Kapoor, A. B. Özgüler, E. T. Holland, J. B. Kowalkowski, Y. Alexeev, and A. L. Lyon, Impacts of noise and structure on quantum information encoded in a quantum memory, Physical Review A 104, 012605 (2021).
  • Blok et al. [2021] M. S. Blok, V. V. Ramasesh, T. Schuster, K. O’Brien, J.-M. Kreikebaum, D. Dahlen, A. Morvan, B. Yoshida, N. Y. Yao, and I. Siddiqi, Quantum information scrambling on a superconducting qutrit processor, Physical Review X 11, 021010 (2021).
  • Ma et al. [2021] W.-L. Ma, S. Puri, R. J. Schoelkopf, M. H. Devoret, S. Girvin, and L. Jiang, Quantum control of bosonic modes with superconducting circuits, Science Bulletin 66, 1789 (2021).
  • Blais et al. [2021] A. Blais, A. L. Grimsmo, S. Girvin, and A. Wallraff, Circuit quantum electrodynamics, Reviews of Modern Physics 93, 025005 (2021).
  • Koch et al. [2022] C. P. Koch, U. Boscain, T. Calarco, G. Dirr, S. Filipp, S. J. Glaser, R. Kosloff, S. Montangero, T. Schulte-Herbrüggen, D. Sugny, et al., Quantum optimal control in quantum technologies. strategic report on current status, visions and goals for research in europe, arXiv preprint arXiv:2205.12110  (2022).
  • Özgüler and Venturelli [2022] A. B. Özgüler and D. Venturelli, Numerical gate synthesis for quantum heuristics on bosonic quantum processors, Frontiers in Physics 10, 900612 (2022).
  • Petersson and Garcia [2021] N. A. Petersson and F. Garcia, Juqbox.jl, GitHub, https://github.com/LLNL/Juqbox.jl (2021).
  • Petersson et al. [2020] N. A. Petersson, F. M. Garcia, A. E. Copeland, Y. L. Rydin, and J. L. DuBois, Discrete adjoints for accurate numerical optimization with application to quantum control, arXiv preprint arXiv:2001.01013  (2020).
  • Anders Petersson and Garcia [2022] N. Anders Petersson and F. Garcia, Optimal control of closed quantum systems via b-splines with carrier waves, SIAM Journal on Scientific Computing 44, A3592 (2022).
  • Note [1] For an in-depth discussion on the concepts of the logical frame and the echo operator, please refer to Eqs. (294-295) in Ref. [21], which corresponds to Eqs. (8-9) in this manuscript.
  • Gorin et al. [2006] T. Gorin, T. Prosen, T. H. Seligman, and M. Žnidarič, Dynamics of loschmidt echoes and fidelity decay, Physics Reports 435, 33 (2006).
  • Gu [2010] S.-J. Gu, Fidelity approach to quantum phase transitions, International Journal of Modern Physics B 24, 4371 (2010).
  • Özgüler et al. [2020] A. B. Özgüler, C. Xu, and M. G. Vavilov, Response of a quantum disordered spin system to a local periodic drive, Physical Review B 101, 024204 (2020).
  • Nguyen et al. [2019] L. B. Nguyen, Y.-H. Lin, A. Somoroff, R. Mencia, N. Grabon, and V. E. Manucharyan, High-coherence fluxonium qubit, Physical Review X 9, 041041 (2019).
  • Place et al. [2021] A. P. Place, L. V. Rodgers, P. Mundada, B. M. Smitham, M. Fitzpatrick, Z. Leng, A. Premkumar, J. Bryon, A. Vrajitoarea, S. Sussman, et al., New material platform for superconducting transmon qubits with coherence times exceeding 0.3 milliseconds, Nature communications 12, 1 (2021).
  • Özgüler et al. [2021] A. B. Özgüler, V. E. Manucharyan, and M. G. Vavilov, Excitation dynamics in inductively coupled fluxonium circuits, arXiv preprint arXiv:2104.03300  (2021).
  • Lidar [2014] D. A. Lidar, Review of decoherence-free subspaces, noiseless subsystems, and dynamical decoupling, Quantum information and computation for chemistry , 295 (2014).
  • Pokharel and Lidar [2023] B. Pokharel and D. A. Lidar, Demonstration of algorithmic quantum speedup, Physical Review Letters 130, 210602 (2023).
  • Özgüler et al. [2018] A. B. Özgüler, R. Joynt, and M. G. Vavilov, Steering random spin systems to speed up the quantum adiabatic algorithm, Physical Review A 98, 062311 (2018).
  • Xu et al. [2022] D. Xu, A. B. Özgüler, G. D. Guglielmo, N. Tran, G. N. Perdue, L. Carloni, and F. Fahim, Neural network accelerator for quantum control, in 2022 IEEE/ACM Third International Workshop on Quantum Computing Software (QCS) (IEEE Computer Society, Los Alamitos, CA, USA, 2022) pp. 43–49.
  • Ogunkoya et al. [2024] O. Ogunkoya, J. Kim, B. Peng, A. B. Özgüler, and Y. Alexeev, Qutrit circuits and algebraic relations: A pathway to efficient spin-1 hamiltonian simulation, Physical Review A 109, 012426 (2024).
  • Cai et al. [2021] W. Cai, Y. Ma, W. Wang, C.-L. Zou, and L. Sun, Bosonic quantum error correction codes in superconducting quantum circuits, Fundamental Research 1, 50 (2021).