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

Minimizing state preparation times in pulse-level variational molecular simulations

Ayush Asthana Department of Chemistry, Virginia Tech, Blacksburg, VA 24061, USA    Chenxu Liu Department of Physics, Virginia Tech, Blacksburg, VA 24061, USA    Oinam Romesh Meitei Department of Chemistry, Massachusetts Institute of Technology, Cambridge, MA 02139, USA    Sophia E. Economou Department of Physics, Virginia Tech, Blacksburg, VA 24061, USA    Edwin Barnes Department of Physics, Virginia Tech, Blacksburg, VA 24061, USA    Nicholas J. Mayhall [email protected] Department of Chemistry, Virginia Tech, Blacksburg, VA 24061, USA
Abstract

Quantum simulation on NISQ devices is severely limited by short coherence times. A variational pulse-shaping algorithm known as ctrl-VQE was recently proposed to address this issue by eliminating the need for parameterized quantum circuits, which lead to long state preparation times. Here, we find the shortest possible pulses for ctrl-VQE to prepare target molecular wavefunctions for a given device Hamiltonian describing coupled transmon qubits. We find that the time-optimal pulses that do this have a bang-bang form consistent with Pontryagin’s maximum principle. We further investigate how the minimal state preparation time is impacted by truncating the transmons to two versus more levels. We find that leakage outside the computational subspace (something that is usually considered problematic) speeds up the state preparation, further reducing device coherence-time demands. This speedup is due to an enlarged solution space of target wavefunctions and to the appearance of additional channels connecting initial and target states.

I Introduction

Noisy intermediate-scale quantum (NISQ) devices have qubits with low coherence times and frequent gate errors. This constrains today’s quantum algorithms to short circuit depths and has inspired many algorithms aimed at minimizing the quantum resources needed to obtain accurate results [1, 2, 3, 4, 5, 6, 7]. Despite rapid advances on both the algorithmic and hardware fronts, the path to useful quantum advantage remains elusive [8, 9, 10, 11, 12, 13, 14, 15]. Thus, generating unitary transformations in a short time on a quantum device is of fundamental and practical importance in quantum computing.

Because most qubit platforms contain at least two levels, one must typically designate two of the states (i.e., |0\ket{0} and |1\ket{1}) as the “computational basis” (the states to which the computational problem is mapped). Any uncontrolled population of states outside of the computational space (referred to as “leakage”) is generally harmful from a computational perspective. For example, superconducting transmon qubits are one of the leading approaches for realizing scalable [16] quantum devices with high fidelity and low noise in gates [17]. However, due to the typically low anharmonicity of the transmon qubits, higher energy levels are easily populated. While these higher energy states can sometimes be leveraged as an additional quantum resource  [18, 19, 20, 21, 22, 23, 24, 25, 26], if the quantum logic gates can be extended to explicitly act on the increased computational space, leakage is generally harmful, leading to difficulties in error mitigation and measurement  [27, 28, 29]. To minimize leakage, one typical approach is to constrain the power of the drive or to use optimal control techniques [30, 31] to implement fast quantum gates with high fidelity on transmon qubits [32, 33].

On the algorithms side, variational quantum algorithms (VQAs) [34] are among the most promising approaches for solving a range of both optimization and simulation problems on NISQ era devices due to their resilience to noise [35]. Although VQAs have been developed for applications including quantum chemistry [36, 12, 14], optimization problems [37, 38], quantum neural networks [39, 40, 41, 42] and many more, our focus is on the electronic structure problem, which deals with finding the low-lying energy states of a molecule  [36, 12, 12]. The electronic structure problem is the cornerstone of quantum chemical calculations, which are believed to be one of the leading fields in which quantum computers can demonstrate an advantage over classical computers. Molecular simulations on a quantum computer have already reproduced some known results for small molecules using variational quantum eigensolver (VQE) algorithms [1]. VQEs, a subset of VQAs, are classical-quantum algorithms that approximate the ground state of a many-body Hamiltonian by optimizing a parameterized quantum circuit that prepares a trial state wavefunction (|Ψ(θ)\ket{\Psi(\theta)}) to minimize the molecular energy. The optimal parameter values are given by

θ=arg minθ(Ψ(θ)|H^mol|Ψ(θ)),\displaystyle\theta^{*}=\text{arg min}_{\theta}(\langle\Psi(\theta)|\hat{H}_{\text{mol}}|\Psi(\theta)\rangle), (1)

where H^mol\hat{H}_{\text{mol}} is the molecular Hamiltonian.

It has recently been recognized that VQAs can be both better understood and significantly improved by establishing connections with quantum optimal control theory (QOC) [43, 44, 45, 46, 47]. VQAs involve the optimization of parameters in discrete quantum circuit logic elements, whereas QOC functions more broadly to optimize a time-dependent drive Hamiltonian, typically by determining optimal pulse shapes of applied fields. Every parameterized quantum circuit corresponds to a pulse, but not all pulses have an associated parameterized quantum circuit. Although the two approaches share the same objective, to reach the final state that minimizes a cost function (e.g., molecular energy), QOC-based VQAs are inherently more powerful than circuit-based VQAs.

This connection is directly exploited by the recently developed ctrl-VQE algorithm [43], which defines a QOC-based ansatz at the device level, taking pulse parameters as optimization parameters and performing VQA optimization steps to minimize the molecular energy. Ctrl-VQE has been shown to reduce the coherence-time requirements for molecular simulations by several orders of magnitude, therefore showcasing the potential to simulate strongly correlated systems on NISQ-era quantum computers.

In addition to the reduction in coherence-time demands demonstrated in Ref. [43], connections between QOC and VQAs also provide critical conceptual insights regarding the complexity of numerical optimization. For controllable quantum systems with unconstrained quantum resources, the control landscape has been shown to be free from local minima [48]. The absence of local minima is important for the optimization process to reach the target state without getting stuck in traps. On the other hand, such traps often occur in parameterized gate-based VQAs unless more than a critical number of parameters are included (referred to as overparameterization) [49]. A conceptual understanding of numerical optimization is important for the development of scalable quantum computing algorithms.

The minimum evolution time (MET) required to rotate one quantum state to another has been studied using both constrained and unconstrained controls [50, 51, 52]. The field required to drive a MET transformation forms the “time-optimal” control, and the parameters for this field are the time-optimal control parameters. Coming from classical control theory, Pontryagin’s maximum principle has been applied to study optimal control for VQAs with a continuous-time evolution, using an amplitude-constrained drive [45, 53, 54]. Yang et al. [45] found that the time-optimal controls exhibit a bang-bang form (control parameters saturate the bounds) due to the linear dependence on control waveforms in the drive Hamiltonian. Brady et al. [53] showed that more general optimal waveform shapes may exist (depending on the problem), as a key assumption in earlier work is not valid universally.

In this work, we show through numerical simulations that ctrl-VQE yields time-optimal control fields that prepare target molecular ground states in the shortest time possible for a given transmon device. These optimal pulses have a bang-bang form that is fully consistent with the Pontryagin maximum principle, as we confirm with analytical calculations. Using these time-optimal solutions, we investigate the effect of additional states outside the computational subspace, as naturally occur in transmon systems. We find that, surprisingly and contrary to typical scenarios, leakage outside of the computational space can significantly improve the performance of ctrl-VQE by further reducing the state preparation time. We show that this speedup can be attributed to two factors: (i) an expansion of the solution space of target wavefunctions, and (ii) the appearance of shorter quantum trajectories that are made accessible via the leakage states. These results constitute an important step toward overcoming coherence-time limitations of quantum simulation on NISQ devices.

This paper is organized in the following way. We first discuss the ctrl-VQE algorithm (Sec. II.1) and time-optimal controls (Sec. II.2). We then provide the computational details for the simulations in this paper in Sec. III. Results related to time-optimal controls in ctrl-VQE (IV.1) and faster state evolution using leakage states (IV.2) are discussed in Sec. IV. Finally, we present the summary and outlook in Sec. V.

II Algorithm and time-optimal controls

II.1 ctrl-VQE

The ctrl-VQE approach uses a QOC-based wavefunction ansatz at the device level, instead of parameterized quantum circuits, for VQE simulations [43]. Because ctrl-VQE creates a device-specific ansatz built from parameterized control fields, the Hamiltonian of the device needs to be defined upfront. In our simulations, we use superconducting transmon qubits with constant, always-on, interqubit couplings. The device Hamiltonian is given by

H^D=qωqaqaqqδq2aqaqaqaq+pqgpqapaq,\displaystyle\hat{H}_{D}=\sum_{q}\omega_{q}a^{\dagger}_{q}a_{q}-\sum_{q}\frac{\delta_{q}}{2}a^{\dagger}_{q}a^{\dagger}_{q}a_{q}a_{q}+\sum_{\expectationvalue{pq}}g_{pq}a^{\dagger}_{p}a_{q}, (2)

where ωq\omega_{q} is the transition frequency of the |0|1\ket{0}\rightarrow\ket{1} transition (resonance frequency) for qubit qq, δq\delta_{q} is the qubit anharmonicity, gpqg_{pq} is the constant coupling rate for a pair of transmons, and the brackets in pq\expectationvalue{pq} indicate a restricted sum over directly connected transmons. The control Hamiltonian that drives the system under the rotating wave approximation is given by

H^C(t)=qΩq(t)(eiνqtaq+eiνqtaq),\displaystyle\hat{H}_{C}(t)=\sum_{q}\Omega_{q}(t)(e^{i\nu_{q}t}a_{q}+e^{-i\nu_{q}t}a^{\dagger}_{q}), (3)

where Ωq(t)\Omega_{q}(t) and νq\nu_{q} are the time-dependent amplitude and (time-independent) frequency of the drive on qubit qq. The state evolves according to the time-dependent Schrödinger equation in the interaction picture,

t|ΨI(t)=iH^I,C(t)|ΨI(t),\displaystyle\frac{\partial}{\partial t}\ket{\Psi_{I}(t)}=-i\hat{H}_{I,C}(t)\ket{\Psi_{I}(t)}, (4)

where the |ΨI(t)\ket{\Psi_{I}(t)} is the wavefunction, and the interaction picture control Hamiltonian is given by

H^I,C(t)=eiH^DtH^C(t)eiH^Dt.\displaystyle\hat{H}_{I,C}(t)=e^{i\hat{H}_{D}t}\hat{H}_{C}(t)e^{-i\hat{H}_{D}t}. (5)

To obtain the final state at the time, TT, the time-dependent Schrödinger equation is solved, while the optimization is carried out on the drive amplitudes and frequency (Ω(t)n\Omega(t)_{n}, νn\nu_{n} for each qubit), which are varied to minimize the molecular energy given by

Emol=Ψ~I(T)|H^mol|Ψ~I(T).\displaystyle E_{\text{mol}}=\langle\tilde{\Psi}_{I}(T)|\hat{H}_{\text{mol}}|\tilde{\Psi}_{I}(T)\rangle. (6)

Note that H^mol\hat{H}_{\text{mol}} is the molecular Hamiltonian mapped to the qubit system, and |Ψ~I(T)\ket{\tilde{\Psi}_{I}(T)} is the wavefunction at the final evolution time (|ΨI(T)\ket{\Psi_{I}(T)}) projected onto the computational subspace, which is the subspace spanned by the computational basis states (|0\ket{0} and |1\ket{1}) of each qubit. The wavefunction in the computational basis is then normalized, although the unnormalized wavefunction can be used in the computations to eliminate any leakage. TT is the total time of evolution and is kept fixed during a simulation. The ctrl-VQE ansatz can be summarized as

|ΨItrial(Ωn(t),νn)=𝕋^et=0T𝑑tH^I,C(t,Ω(t)n,νn)|ψI(0),\displaystyle|\Psi_{I}^{\text{trial}}(\Omega_{n}(t),\nu_{n})\rangle=\hat{\mathbb{T}}e^{-\int_{t=0}^{T}dt\hat{H}_{I,C}(t,\Omega(t)_{n},\nu_{n})}|\psi_{I}(0)\rangle, (7)

where 𝕋^\hat{\mathbb{T}} is the time-ordering operator, and |ψI(0)\ket{\psi_{I}(0)} is the initial state, usually taken as the Hartree-Fock (HF) wavefunction. This trial wavefunction is evaluated at each iteration, and the optimization parameters are varied to minimize the molecular energy. In what follows, we suppress the subscript II on the states for notational simplicity.

Refer to caption
Figure 1: Qualitative depiction of state evolution in Hilbert space of qudits with two or more levels. The time-optimal path in the computational subspace (yellow surface) is shown as the path with evolution time T=TT=T^{*}. Many paths are possible to reach the solution |Ψ(T)|\Psi(T)\rangle when T>TT>T^{*} (gray curvy line). Access to leakage states provides new paths (shown by darker red paths) that reach the solution faster than TT^{*} when the final state is projected onto the computational subspace.

II.2 Minimum evolution time and time-optimal controls

Control problems aim to find a target state that minimizes the cost function, which in the case of molecular simulation is the molecular energy [Eq. (6)]. In the qualitative depiction in Fig. 1, this state is represented by |Ψ(T)\ket{\Psi(T)}. Starting from the initial state, represented by |Ψ(0)\ket{\Psi(0)} (see Fig. 1), the state evolves in time through the action of driving pulses until it reaches the target state. There exists a minimum pulse duration (the MET) that is required to carry out this state evolution. Control pulses with durations that exceed the MET (T>TT>T^{*} in Fig. 1 where TT^{*} is the MET for qubits) can drive the system along one of the infinitely many possible trajectories in Hilbert space to reach the target state, while time-optimal pulses drive the initial state to the target state along special trajectories in a time equal to the MET. The MET generally depends on the constraints imposed on the driving pulses. Ctrl-VQE generates controllable evolutions in the control theory sense, as the entire Hilbert space can be spanned to arbitrary precision by varying unconstrained parameters of the driving pulses. In ctrl-VQE, as the total pulse duration (TT) increases, the Hilbert space accessible to the qubit system increases as well.

In the present study, we numerically find TT^{*} and the associated time-optimal controls by starting at a large enough pulse duration and reducing the pulse duration until we find the shortest pulse at which the system can reach the target state to the desired accuracy. The pulse is divided into many step-wise constant segments with practical bounds on amplitudes and frequencies. Dividing the pulse into many adjustable segments provides a large amount of freedom in shaping the pulse, essentially allowing the ctrl-VQE algorithm to choose any possible optimal pulse shape in the optimization process. Because the numerical optimization of time-optimal controls is quite challenging, it is difficult to ensure that one has converged to TT^{*}. As a guarantee, in what follows we compare our numerical solutions to analytical results obtained from Pontryagin’s theory to confirm that our numerically obtained METs are correct.

III Computational details

The calculations are carried out using a local version of the ctrlq code [43, 55]. As a test system for our simulations, we have chosen the problem of finding the ground state energy of the H2 molecule at a bond distance of 1.5Å1.5~{}\text{\AA}, which has relatively strong correlations. Molecular integrals are generated using PySCF [56] and STO-3G basis set is used in this work. We have used the parity mapping with Z2 symmetry reduction to map molecular Hamiltonian on the qubit basis. The initial state is taken as the unentangled HF state for this problem, and the target state is the exact ground state (i.e., full configuration interaction (FCI)). The numerical optimization of pulse amplitudes and frequencies is carried out using the l-BFGS-b optimization algorithm, and the pulses are taken to be piece-wise constant with 100 segments (unless otherwise stated). The time-discretized amplitude, Ω(t)\Omega(t), is given by

Ωk(t)={c1 0tt1c2t1tt2cntn1tT\displaystyle\Omega_{k}(t)=\begin{cases}c_{1}&\>0\leq t\leq t_{1}\\ c_{2}&\>t_{1}\leq t\leq t_{2}\\ \dots\\ c_{n}&\>t_{n-1}\leq t\leq T\end{cases} (8)

for qubit kk, where c1,c2cnc_{1},c_{2}\dots c_{n} represent amplitude parameters for each pulse segment, and TT is the final time. The pulse amplitude is constrained between ±20\pm 20 MHz unless otherwise mentioned, which is based on common values for control signals in transmon qubits [57]. The carrier signal frequency νk\nu_{k} given in Eq. (3) is constrained such that its difference from the qubit frequency, νkωk\nu_{k}-\omega_{k}, is in the range ±1\pm 1 GHz, where ωk\omega_{k} is the qubit frequency. A convergence error threshold of 10810^{-8} Hartree in molecular energy is used in the optimization process, which corresponds to close to unit fidelity with the target FCI wavefunction. Since we encounter many local minima when the pulse duration is close to its minimum value, we consider MET to be the minimum pulse duration at which a solution could be reached in 1000 random initializations. The results are minimally affected by increasing the number of pulse segments or Trotter steps (see SI for details). The parameters in the device Hamiltonian, ωq\omega_{q}, ηq\eta_{q} and gpqg_{pq}, used in the simulations are given in Table 1.

Transmon
1 2
ω/(2π)\omega/(2\pi) 4.8080 4.8333
δ/(2π)\delta/(2\pi) 0.3102 0.2916
121\rightarrow 2
g/(2π)g/(2\pi) 0.01831
Table 1: Device parameters used in Eq. (3) for a two-transmon system. All values are in GHz.

IV Results and discussion

In the following subsections, we discuss our main results related to time-optimal controls in the ctrl-VQE algorithm applied to transmon qubits. We first discuss the time-optimal pulse shape for ctrl-VQE through the application of Pontryagin’s maximum principle and numerical simulations. Next, we qualitatively analyze the near-MET evolution of the two-qudit state, with a particular focus on the role of leakage to higher transmon levels outside the computational subspace.

IV.1 Time-optimal control in ctrl-VQE

Refer to caption
(a) qubit system
Refer to caption
(b) qutrit system
Figure 2: (a) How the optimized pulse shape changes for a two-qubit system as the pulse duration is reduced from 20.0020.00 ns to 15.0015.00 ns. (b) How the optimized pulse shape changes for a two-qutrit system as the pulse duration is reduced from 14.0014.00 ns to 8.948.94 ns. The H2 ground state energy is calculated in these simulations at a bond distance of 1.5Å1.5~{}\text{\AA}. The pulse shape evolves into bang-bang form as the pulse duration is restricted towards the minimum evolution time (MET) needed to produce the ground state.

Time-optimal controls for quantum algorithms have been previously studied using quantum optimal control theory with continuous-time controls [45, 53, 54, 58]. Pontryagin’s principle is applied to derive conditions under which an associated cost function is minimized at the MET. We will follow the same procedure and use Pontryagin’s principle to find conditions for time-optimal controls for the ctrl-VQE algorithm. The dynamics of the multi-qudit state in ctrl-VQE is governed by the time-dependent Schrödinger equation given in Eqs. (4)-(5). The cost function in the case of the electronic structure problem takes the form of Eq. (6). This can be formulated into a time-optimal control problem, with the modified cost function (not to be confused with the cost function associated with the ctrl-VQE algorithm) to be minimized as

J\displaystyle J =Ψ(T)|H^mol|Ψ(T)\displaystyle=\langle\Psi(T)|\hat{H}_{\text{mol}}|\Psi(T)\rangle
+0Tλ(t)|(iH^I,C(t)|Ψ(t)|Ψ˙(t))𝑑t+h.c.\displaystyle+\int_{0}^{T}\langle\lambda(t)|\left(-i\hat{H}_{I,C}(t)|\Psi(t)\rangle-|\dot{\Psi}(t)\rangle\right)dt+h.c. (9)

This corresponds to minimizing the molecular energy at the final time TT, under the constraint that the wavefunction is obtained as a proper solution to the device Schrödinger equation. Here, λ(t)|\bra{\lambda(t)} and |λ(t)\ket{\lambda(t)} are treated as Lagrange multipliers. The control function 111conventionally known as control Hamiltonian but we use the edited name to avoid confusion with drive Hamiltonian (H¯C\bar{H}_{C}) may be defined following Pontryagin’s maximum principle [60, 54, 53] and Eq. (4), taking the form

\displaystyle\mathbb{H} =λ(t)|iH^I,C(t)|Ψ(t)Ψ(t)|iH^I,C(t)|λ(t),\displaystyle=\langle\lambda(t)|i\hat{H}_{I,C}(t)|\Psi(t)\rangle-\langle\Psi(t)|i\hat{H}_{I,C}(t)|\lambda(t)\rangle, (10)

with the time evolution of state |Ψ(t)|\Psi(t)\rangle and co-state |λ(t)|\lambda(t)\rangle, and end point of co-state |λ(T)|\lambda(T)\rangle given by

|Ψ˙(t)=iHI,C|Ψ(t),\displaystyle|\dot{\Psi}(t)\rangle=-iH_{I,C}|\Psi(t)\rangle, (11)
|λ˙(t)=iHI,Cλ(t),\displaystyle|\dot{\lambda}(t)\rangle=-iH_{I,C}\lambda(t), (12)
|λ(T)=H^mol|Ψ(T).\displaystyle|\lambda(T)\rangle=\hat{H}_{\text{mol}}|\Psi(T)\rangle. (13)

The derivation of the control function \mathbb{H} and details of the application of Pontryagin’s principle are given in the SI. At the optimal control (Ω(t)\Omega^{*}(t),Ψ(t)\Psi^{*}(t),λ(t)\lambda^{*}(t) where denotes the optimal path), Pontryagin’s maximum principle states that

(t,Ωn(t),Ψ(t),λ(t))(t,Ωn(t),Ψ(t),λ(t)),\displaystyle\mathbb{H}(t,\Omega_{n}^{*}(t),\Psi^{*}(t),\lambda^{*}(t))\geq\mathbb{H}(t,\Omega_{n}(t),\Psi^{*}(t),\lambda^{*}(t)), (14)

at each point in time. Notice that the \mathbb{H} formed in ctrl-VQE is an explicit function of time. \mathbb{H} may be written as a sum of \mathbb{H} for each qubit as

=qq,\displaystyle\mathbb{H}=\sum_{q}\mathbb{H}_{q}, (15)

where qq is the qubit index. For ctrl-VQE, the control function for each qubit may be written as

q=2Reλ(t)|iΩq(t)eiH^Dt(eiνqtaq+eiνqtaq)eiH^Dt|Ψ(t),=Ωq(t)ϕq(t),\displaystyle\begin{split}\mathbb{H}_{q}=&2\text{Re}\langle\lambda(t)|-i\Omega_{q}(t)e^{i\hat{H}_{D}t}(e^{i\nu_{q}t}a_{q}\\ &+e^{-i\nu_{q}t}a^{\dagger}_{q})e^{-i\hat{H}_{D}t}|\Psi(t)\rangle,\\ =&\Omega_{q}(t)\phi_{q}(t),\end{split} (16)

where the switching function ϕq(t)\phi_{q}(t) is given by

ϕq(t)=2Reλ(t)|ieiH^Dt(eiνqtaq+eiνqtaq)eiH^Dt|Ψ(t).\displaystyle\begin{split}\phi_{q}(t)=&2\text{Re}\langle\lambda(t)|-ie^{i\hat{H}_{D}t}(e^{i\nu_{q}t}a_{q}+e^{-i\nu_{q}t}a^{\dagger}_{q})\\ &e^{-i\hat{H}_{D}t}|\Psi(t)\rangle.\end{split} (17)

Through Eq. (14) to (17), it can be shown that the optimal controls Ωq(t)\Omega^{*}_{q}(t) obey the following condition,

Ωq(t)={max boundif ϕq(t)>0min boundif ϕq(t)<0undefinedif ϕq(t)=0.\Omega^{*}_{q}(t)=\begin{cases}\text{max bound}&\text{if }\phi_{q}(t)>0\\ \text{min bound}&\text{if }\phi_{q}(t)<0\\ \text{undefined}&\text{if }\phi_{q}(t)=0.\end{cases} (18)

The result of this is that optimal controls follow a “bang-bang” structure when the switching function has a non-zero value; specifically the controls are maximized at time tt in the case of positive ϕq(t)\phi_{q}(t) and vice versa. The optimal solution is undefined in the case of a singular switching function ϕq(t)=0\phi_{q}(t)=0 as the control function becomes independent of control parameters in that case. This condition, ϕq(t)=0\phi_{q}(t)=0, is well studied and is known as the “singularity” condition [60]. For a single-qubit system, Eq. 17 is zero for an extended period of time only when λ(t)=Ψ(t)\lambda(t)=\Psi(t) for an extended period. This is not possible in the electronic structure problem due to Eq. (13); therefore, the optimal protocol will follow a bang-bang form for a single qubit. To study the case of two qubits, as in the case of the H2 molecular calculation, we investigate the optimal pulse shape by carrying out ctrl-VQE transmon simulations and limiting the pulse duration to the requisite MET.

In agreement with the analytic results, our simulations reveal the emergence of bang-bang pulse controls for time-optimal simulations of the H2 ground state at a bond length of 1.5Å1.5~{}\text{\AA}. Fig. 2 presents the pulse shapes obtained after optimization at different pulse durations for both 2-level qubits (Fig. 2(a)) and 3-level qutrits (Fig. 2(b)). Similar to the cartoon depiction in Fig. 1, there are many ways of reaching the final state when the evolution time exceeds the MET (i.e., at T>TT>T^{*}). In the top two panels, we present one of the many solutions found, noting that both qubits and qutrits exhibit a highly unstructured pulse shape. In the middle two panels (1616 ns and 1010 ns, respectively), both qubits and qutrits begin to develop a bang-bang-like structure. As the pulse duration is further reduced, the pulse shape becomes increasingly more uniform until it reaches a well organized bang-bang shape near a MET of 1515 ns and 8.948.94 ns for qubits and qutrits, respectively.

Refer to caption
(a) qubit 1
Refer to caption
(b) qubit 2
Figure 3: (a) The switching function ϕ1(t)\phi_{1}(t) (yellow dashed) and pulse shape Ω1(t)\Omega_{1}^{*}(t) (green) for the optimal drive on qubit 1 in two-level transmon qubits. (b) The switching function ϕ2(t)\phi_{2}(t) (yellow dashed) and pulse shape Ω2(t)\Omega_{2}^{*}(t) (red) for the optimal drive on qubit 2 in two-level transmon qubits. The plots are made using 100 pulse segments, 1000 Trotter steps in the simulation, and a pulse duration of T=14.93T=14.93 ns, reaching an accuracy of 10610^{-6} Hartree relative to the FCI energy.

As the evolution time is decreased (such that we reach our VQE trial state faster), the computational task of optimizing a pulse becomes increasingly difficult due to the quickly rising number of local parameter traps. This makes it difficult to confirm that one has actually converged to TT^{*}. In order to verify that we are near TT^{*}, in Fig. 3, we compare the switching function from Eq. (17), ϕq(t)\phi_{q}(t), with the optimized pulses to check the agreement with results obtained using Pontryagin’s maximum principle. By plotting the function ϕq(t)\phi_{q}(t) 222Here we use 100 pulse segments and 1000 Trotter steps. We tested smaller and larger Trotter steps and found no appreciable change in results. atop the optimized pulses in Fig. 3, we can readily see that the pulse switching times coincide with the roots of the switching function, and that at each time, the pulse is maximized in the direction determined by the sign of ϕq(t)\phi_{q}(t), in agreement with Pontryagin’s maximum principle, i.e., when ϕq(t)\phi_{q}(t) is positive (negative), Ωq(t)\Omega_{q}(t) should assume the largest positive (negative) value allowed by the constraints.

For the calculation of ϕq(t)\phi_{q}(t), we first carry out a ctrl-VQE simulation using time-optimal controls to evaluate |Ψ(t)\ket{\Psi(t)} at each time step. We then form |λ(T)\ket{\lambda(T)}, and back-propagate it using Eqs. (12) and (13). Finally, we evaluate ϕq(t)\phi_{q}(t) through Eq. (17). It can be observed that there is no region in time where the function ϕq(t)\phi_{q}(t) is 0 for a continuous period of time. Further, the times at which ϕq(t)\phi_{q}(t) vanishes coincide with the times at which the pulse changes sign, consistent with the analytical prediction of Eq. (18). This can also be understood from the fact that ϕq(t)\phi_{q}(t) is proportional to the analytical derivative derived in Ref. [43].

IV.2 Minimum evolution time for a system of transmon qudits

IV.2.1 Leakage speeds up evolution

The ctrl-VQE optimized pulses for H2 shown in Fig. 2 demonstrate that the ground state is obtained significantly faster when higher transmon levels are available. The MET found for a system of qubits for this problem was 15.00 ns while the MET for a system of qutrits was 8.94 ns. This amounts to a 40\sim 40% reduction in evolution time when a qutrit system is used. Considering that short coherence times are one of the most restrictive bottlenecks on NISQ devices, this reduction in state preparation time constitutes a significant practical advantage of this approach.

Of course, this decreased evolution time is not entirely free of cost. By populating higher energy states (as depicted in Fig. 1), the final molecular energy can only be computed by first projecting onto the computational space and renormalizing the expectation value. However, assuming one can resolve the difference between computational states and leakage states in a measurement, one can simply post-select, discarding measurements of leakage states, but keeping track of how often such measurement outcomes are obtained. The percentage of leakage state measurements then directly informs one how to normalize the objective function. With increasing weight in the leakage space, the number of discarded measurements increases, which requires one to perform an increasing number of shots. However, this increased shot-count overhead is directly determined by the extent of the leakage, which is also partially controlled by the amplitude constraints on the pulse.

Refer to caption
Figure 4: Success probability of the simulation to reach the target state vs pulse duration. The probability is computed using 100 random initializations and with an energy error of less than 10810^{-8} Hartree compared to the FCI solution. The plot shows that transmons truncated to two levels require more evolution time as compared to transmons with more than two levels.

In addition to the decreased time of evolution, access to leakage states also improves the ability to optimize the pulses themselves. Fig. 4 shows the probability of finding the target solution plotted against pulse duration for transmons with 2, 3, 4, and 5 levels retained. The probability is determined by using 100 random initializations and counting how many of them succeed in preparing the ground state of the H2H_{2} molecule (to within 10810^{-8} Hartree). The results show that the probability to reach the target state drops abruptly at 9\sim 9 ns for qutrits and at \sim19 ns for qubits. It is evident that the computational difficulty to reach the target state in the case of qubits is significantly higher, and it takes more optimization steps for the simulation to converge in the case of qubits as compared to qutrits (increase in optimisation steps near MET was also observed in Ref. [52]). This indicates that a computational phase transition occurs at the MET. The speedup obtained using qutrits is not further enhanced by using more than three levels in our two-transmon calculations. This is likely due to the fact that the higher levels are far detuned from the microwave drive, and are thus not easily populated by the microwave drive. Therefore, the contribution of these states to preparing the final state is negligible compared to that of the third transmon level.

We find that the leakage at pulse durations close to the MET for qutrits is 50%\sim 50\%, indicating a necessary 2-fold increase in the number of shots required. We have further seen that this leakage can be optimized to a desired fraction of the total population by introducing a penalty term in our cost function to penalize leakage over the desired amount. For instance, using a 0.01 Hartree penalty per 1% leakage for any amount of leakage over 10%10\% of the total population (such that a leakage of 11% will result in an energy penalty of 0.01 Hartree, while any amount of leakage below 10% will have no energy penalty) reduces the MET from 15 ns to about 12.5 ns (see SI for details). So a penalty can be introduced in the cost function to keep the leakage below a target threshold, while still taking advantage of faster state evolution. However, we find that a penalty term that reduces the leakage to zero does not result in any speedup in our simulations. The no-leakage case is the same as performing the simulation without projecting onto the computational space (the space spanned by the |0|0\rangle and |1|1\rangle levels of the qudits), which has also not resulted in a faster state evolution in our simulations. This indicates that some population in the higher-lying qudit levels is necessary at the final time to achieve this faster state evolution. Another important observation is that most of the leakage lies in specific states, namely |02|02\rangle and |20|20\rangle in the present simulations.

It is also worth noting that for pulse durations TT significantly greater than the minimum time TT^{*}, we do not observe any traps (getting stuck in local minima) for either qubits or qutrits. This can be understood through results in quantum optimal control which state that a controllable system is free from traps unless the drive is severely constrained [52]. Once the drive has a duration significantly greater than the minimum duration needed for the state evolution, the system reaches the target state without getting hindered by traps. Local traps are often encountered in gate-based VQEs unless a minimum number of parameters are included in the circuit (over-parameterization). A major benefit of the reduced MET is that it allows this trap-free (analogous to over-parameterization in gate-based VQE) region to be reached with shorter pulses in ctrl-VQE.

IV.2.2 Mechanism of evolution speedup in qudits

Refer to caption
(a) Qubits
Refer to caption
(b) Qutrits
Figure 5: Population in various basis states vs evolution time for (a) qubits and (b) qutrits. The solid lines are states in the computational space, while the dashed lines are states that involve the extra transmon level. The initial state is the HF state, while the final state is the FCI wavefunction of H2 at a bond distance of 1.5 Å.
Refer to caption
(a) Two qubits
Refer to caption
(b) Two qutrits
Figure 6: Transition probability from state |01|01\rangle to |10|10\rangle through various intermediate states (mm in Eq. 23) for (a) two qubits and (b) two qutrits. The dotted yellow line shows the sum of contributions from all channels to the state |10|10\rangle.
Refer to caption
(a) Two qubits
Refer to caption
(b) Two qutrits
Figure 7: Major channels involved at second order in evolution from state |01|01\rangle to |10|10\rangle in the case of (a) two qubits and (b) two qutrits. The orange channels are the new channels that arise in the qutrit case.

We investigate the mechanism of the observed speedup using a numerical analysis based on time-dependent perturbation theory to determine which processes contribute the most toward reaching the target wavefunction in the case of qubits and qutrits. In the H2 molecular problem, the initial state is the HF state |01\ket{01}. The FCI target state is a linear combination of |01|01\rangle and |10|10\rangle, where the population in the former is 6.926.92 times that of the latter. Thus, some population is transferred from |01|01\rangle to |10|10\rangle during the evolution. The population of each state as a function of time is shown in Fig. 5, both for transmon qubits (a) and qutrits (b). It is evident that the evolution is very different between the two cases. In the case of qubits, the states |00|00\rangle and |11|11\rangle get populated during the evolution and depopulate completely by the end of the evolution. These two states assist with the transfer of population from |01|01\rangle to |10|10\rangle. In the case of qutrits, the states |00|00\rangle and |11|11\rangle receive very little population throughout the evolution, while states |02|02\rangle and |20|20\rangle exhibit a significant population buildup that remains at the end of the evolution. This striking disparity indicates that higher transmon levels provide alternate channels for reaching the target state. To determine the exact transition channels at work, we analyze the evolution using time-dependent perturbation theory. The unitary evolution operator U(0,T)U(0,T) that carries out the state evolution, |Ψ(T)=U(0,T)|Ψ(0)|\Psi(T)\rangle=U(0,T)|\Psi(0)\rangle, can be expanded in a Dyson series:

U(0,tf)\displaystyle U(0,t_{f}) =I+(i)0tf𝑑tHI,C(t)\displaystyle=I+(-i)\int_{0}^{t_{f}}dtH_{I,C}(t)
+(i)20tf𝑑t10t1𝑑t2HI,C(t1)HI,C(t2)+\displaystyle+(-i)^{2}\int_{0}^{t_{f}}dt_{1}\int_{0}^{t_{1}}dt_{2}H_{I,C}(t_{1})H_{I,C}(t_{2})+\dots (19)

We consider the transition from the initial state |01\ket{01} to the final state |10\ket{10}. The zeroth-order transition amplitude is zero, 01|10=0\langle 01|10\rangle=0, while the first- and second-order transition amplitudes from state |01|01\rangle to |10|10\rangle are

A0110(1)=\displaystyle A_{01\rightarrow 10}^{(1)}= 0tf10|(i)HI,C(t)|01𝑑t,\displaystyle\int_{0}^{t_{f}}\langle 10|(-i)H_{I,C}(t)|01\rangle dt, (20)
A0110(2)=\displaystyle A_{01\rightarrow 10}^{(2)}= 0tf𝑑t10t1𝑑t210|(i)2HI,C(t1)HI,C(t2)|01.\displaystyle\int_{0}^{t_{f}}dt_{1}\int_{0}^{t_{1}}dt_{2}\langle 10|(-i)^{2}H_{I,C}(t_{1})H_{I,C}(t_{2})|01\rangle. (21)

The second-order transition amplitude can be rewritten as a sum over transitions through intermediate states:

A0110(2)=\displaystyle A_{01\rightarrow 10}^{(2)}= m0tf𝑑t10t1𝑑t210|(i)HI,C(t1)|m\displaystyle\sum_{m}\int_{0}^{t_{f}}dt_{1}\int_{0}^{t_{1}}dt_{2}\langle 10|(-i)H_{I,C}(t_{1})|m\rangle
×m|(i)HI,C(t2)|10,\displaystyle\quad\times\langle m|(-i)H_{I,C}(t_{2})|10\rangle, (22)
=\displaystyle= mA01m01,\displaystyle\sum_{m}A_{01\rightarrow m\rightarrow 01}, (23)

where the summation is over all the eigenstates of the transmons. We have checked that evolving the wavefunction using the second-order evolution operator U(2)(0,T)U^{(2)}(0,T) yields the target state up to an infidelity of 0.01\sim 0.01, and hence second order is sufficient for a qualitative study of the state evolution. The first-order transition probability, A0110(1)=0A_{01\rightarrow 10}^{(1)}=0, vanishes for both qubits and qutrits, as the static coupling between transmon qubits is treated as a part of the device Hamiltonian, which is diagonal in the computational basis. Thus the transition from initial to final state is entirely through higher-order processes.

Refer to caption
(a) Two qubits
Refer to caption
(b) Two qutrits
Figure 8: Role of constructive interference between the main transition channels in the case of (a) two qubits (b) two qutrits. The constructive interference (difference between orange and green lines) is more significant in the qubit case.

We show the second-order transition probability from state |01\ket{01} to |10\ket{10} for two transmon qubits and two transmon qutrits in Figs. 6a and b, respectively. The second-order transition probability is defined as

P0110|A0110(2)|2=|mA01m10(2)|2.P_{01\rightarrow 10}\equiv\left|A_{01\rightarrow 10}^{(2)}\right|^{2}=\left|\sum_{m}A_{01\rightarrow m\rightarrow 10}^{(2)}\right|^{2}. (24)

To investigate which transition path is most important in the time-optimal evolution, we compare the contributions to the transition probability from two dominant paths. In the qubit case, the paths are the ones that go through states |00\ket{00} and |11\ket{11} (see Fig. 6a orange line) while for the qutrit case, the dominant paths are the ones through |02\ket{02} and |20\ket{20} (see Fig. 6b orange line). Note that in both cases, the truncated transition probability is close to the second-order one, which shows that the transition is predominantly happening through those channels. Thus, the higher-lying leakage states are participating and providing alternate channels for state evolution. This can be visualized with the help of the schematic level diagram shown in Fig. 7. Basis states |02|02\rangle and |20|20\rangle do not exist in qubits, while this channel starts to become available in qutrits. Further, the population of basis states |00|00\rangle and |11|11\rangle is zero in the target state. This puts a constraint on the dynamics of qubits as it has to depopulate these basis states at the end of the evolution, while these basis states do not have any population in the case of qutrits. Furthermore, in Fig. 8a and b, we compare the truncated transition probability with the squares of the individual transition amplitudes. If the two transition paths constructively interfere with each other, we expect

|A01m110+A01m210|2>|A01m110|2+|A01m210|2.\displaystyle\begin{split}|A_{01\rightarrow m_{1}\rightarrow 10}&+A_{01\rightarrow m_{2}\rightarrow 10}|^{2}\\ &>|A_{01\rightarrow m_{1}\rightarrow 10}|^{2}+|A_{01\rightarrow m_{2}\rightarrow 10}|^{2}.\end{split} (25)

This is observed in the qubit case, while it is not true for the qutrit case (see Fig. 8). Details for analysis using perturbation theory is presented in the SI.

Refer to caption
Figure 9: Population buildup in state |10|10\rangle vs evolution time in the case of two qubits (2-level) and two qutrits (3-level). The basis state |10|10\rangle reaches the target population faster in the case of qutrits when the final state is projected onto the computational subspace and normalized (3-level normalized).

Finally, an additional contribution to the speedup evident in the qutrit case comes from the classical step of projecting the final qutrit state onto the computational subspace and normalizing it. This can be seen from Fig. 9, where the population in state |10|10\rangle is plotted as a function of the evolution time in the case of qubits and qutrits (unnormalized and normalized cases). It can be seen that the population in qutrits increases faster than qubits and reaches the target population faster when the final state on qutrits is projected onto the computational basis and normalized. This is because the population required by the target state in the computational subspace is smaller when we finally normalize the state. The presence of additional channels for state evolution, along with reduced population demand in the desired target state in the computational subspace together leads to a faster state evolution in qutrits compared to qubits.

Although we have focused on the electronic structure problem in the case of the H2 molecule ground state in our current work, the key results are expected to generalize. This is because any problem can be mapped to the task of finding trajectories that connect an initial and target state in Hilbert space. Changing the problem would correspond to changing these initial and target states. Including higher levels will generally provide new channels for state evolution that provide shorter paths to the target state.

V Summary

In this paper, we presented time-optimal control fields that prepare target molecular ground states on transmon quantum processors. These optimal controls were obtained using the pulse-level VQE algorithm known as ctrl-VQE. We found that the optimal controls converge to a bang-bang form, i.e., they saturate pulse amplitude constraints, as the total evolution time is reduced. We showed that this behavior is consistent with analytical results derived from Pontryagin’s maximum principle. These results suggest that ctrl-VQE prepares target states on quantum computers in the shortest possible time allowed by quantum speed limits.

In addition, we investigated how the minimal state preparation time is impacted by truncating the transmons to two or more levels. We find that the inclusion of higher transmon levels can lead to a substantial reduction in the time needed to reach the target state. We showed that this speedup is the result of two effects caused by the additional levels: (i) an enlargement of the space of wavefunctions that have perfect overlap with the target state in the computational subspace, and (ii) the appearance of new transition paths that connect the initial and target states. Together, these effects cut down the evolution time by 40\sim 40% for the simulations performed in this work.

Given the limited coherence times in NISQ devices, it is crucial that we minimize the time it takes to prepare target wavefunctions. Our results indicate that ctrl-VQE does this. Our work also shows that leakage to higher transmon levels is not always problematic. It can potentially be used to reduce the coherence time requirements needed for VQEs. Another important benefit of faster state evolution is that the overparameterized region, where the VQE algorithms become free from local minima (trap-free), is reached faster in qudits as compared with qubits. Local traps have been shown to prevent gate-based VQE algorithms from reaching the solution to the desired accuracy unless a significantly large gate depth is achieved [49].

VI Acknowledgements

This research is supported by the Department of Energy (DOE). E.B. and N.J.M acknowledge Award No. DE-SC0019199, and S.E.E. acknowledges the DOE Office of Science, National Quantum Information Science Research Centers, Co-design Center for Quantum Advantage (C2QA), Contract No. DE-SC0012704. A.A. would like to thank Dr. Luke Bertels and Dr. Sheril Avikkal Kunhippurayil for helpful discussions. The authors thank the Advanced Research Computing (ARC) facility at Virginia Tech for the computational infrastructure.

References

  • Kandala et al. [2017] A. Kandala, A. Mezzacapo, K. Temme, M. Takita, M. Brink, J. M. Chow, and J. M. Gambetta, Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets, Nature 549, 242 (2017).
  • Huggins et al. [2020] W. J. Huggins, J. Lee, U. Baek, B. O’Gorman, and K. B. Whaley, A non-orthogonal variational quantum eigensolver, New Journal of Physics 22, 073009 (2020).
  • Grimsley et al. [2019] H. R. Grimsley, S. E. Economou, E. Barnes, and N. J. Mayhall, An adaptive variational algorithm for exact molecular simulations on a quantum computer, Nature communications 10, 1 (2019).
  • Tang et al. [2021] H. L. Tang, V. Shkolnikov, G. S. Barron, H. R. Grimsley, N. J. Mayhall, E. Barnes, and S. E. Economou, qubit-adapt-vqe: An adaptive algorithm for constructing hardware-efficient ansätze on a quantum processor, PRX Quantum 2, 020310 (2021).
  • Lee et al. [2018] J. Lee, W. J. Huggins, M. Head-Gordon, and K. B. Whaley, Generalized unitary coupled cluster wave functions for quantum computation, Journal of chemical theory and computation 15, 311 (2018).
  • Ryabinkin et al. [2018] I. G. Ryabinkin, T.-C. Yen, S. N. Genin, and A. F. Izmaylov, Qubit coupled cluster method: a systematic approach to quantum chemistry on a quantum computer, Journal of chemical theory and computation 14, 6317 (2018).
  • Gard et al. [2020] B. T. Gard, L. Zhu, G. S. Barron, N. J. Mayhall, S. E. Economou, and E. Barnes, Efficient symmetry-preserving state preparation circuits for the variational quantum eigensolver algorithm, npj Quantum Information 6, 1 (2020).
  • Arute et al. [2019] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, R. Biswas, S. Boixo, F. G. Brandao, D. A. Buell, et al., Quantum supremacy using a programmable superconducting processor, Nature 574, 505 (2019).
  • Zhong et al. [2020] H.-S. Zhong, H. Wang, Y.-H. Deng, M.-C. Chen, L.-C. Peng, Y.-H. Luo, J. Qin, D. Wu, X. Ding, Y. Hu, et al., Quantum computational advantage using photons, Science 370, 1460 (2020).
  • Maslov et al. [2021] D. Maslov, J.-S. Kim, S. Bravyi, T. J. Yoder, and S. Sheldon, Quantum advantage for computations with limited space, Nature Physics 17, 894 (2021).
  • Nam and Maslov [2019] Y. Nam and D. Maslov, Low-cost quantum circuits for classically intractable instances of the hamiltonian dynamics simulation problem, npj Quantum Information 5, 1 (2019).
  • Cerezo et al. [2021] M. Cerezo, A. Arrasmith, R. Babbush, S. C. Benjamin, S. Endo, K. Fujii, J. R. McClean, K. Mitarai, X. Yuan, L. Cincio, et al., Variational quantum algorithms, Nature Reviews Physics , 1 (2021).
  • Cao et al. [2019] Y. Cao, J. Romero, J. P. Olson, M. Degroote, P. D. Johnson, M. Kieferová, I. D. Kivlichan, T. Menke, B. Peropadre, N. P. Sawaya, et al., Quantum chemistry in the age of quantum computing, Chemical reviews 119, 10856 (2019).
  • Tilly et al. [2021] J. Tilly, H. Chen, S. Cao, D. Picozzi, K. Setia, Y. Li, E. Grant, L. Wossnig, I. Rungger, G. H. Booth, et al., The variational quantum eigensolver: a review of methods and best practices, arXiv preprint arXiv:2111.05176  (2021).
  • Fedorov et al. [2022] D. A. Fedorov, B. Peng, N. Govind, and Y. Alexeev, Vqe method: A short survey and recent developments, Materials Theory 6, 1 (2022).
  • Acín et al. [2018] A. Acín, I. Bloch, H. Buhrman, T. Calarco, C. Eichler, J. Eisert, D. Esteve, N. Gisin, S. J. Glaser, F. Jelezko, et al., The quantum technologies roadmap: a european community view, New Journal of Physics 20, 080201 (2018).
  • Kjaergaard et al. [2020] M. Kjaergaard, M. E. Schwartz, J. Braumüller, P. Krantz, J. I.-J. Wang, S. Gustavsson, and W. D. Oliver, Superconducting qubits: Current state of play, Annual Review of Condensed Matter Physics 11, 369 (2020).
  • Neeley et al. [2009] M. Neeley, M. Ansmann, R. C. Bialczak, M. Hofheinz, E. Lucero, A. D. O’Connell, D. Sank, H. Wang, J. Wenner, A. N. Cleland, et al., Emulation of a quantum spin with a superconducting phase qudit, Science 325, 722 (2009).
  • Lanyon et al. [2009] B. P. Lanyon, M. Barbieri, M. P. Almeida, T. Jennewein, T. C. Ralph, K. J. Resch, G. J. Pryde, J. L. O’brien, A. Gilchrist, and A. G. White, Simplifying quantum logic using higher-dimensional hilbert spaces, Nature Physics 5, 134 (2009).
  • Bechmann-Pasquinucci and Peres [2000] H. Bechmann-Pasquinucci and A. Peres, Quantum cryptography with 3-state systems, Physical Review Letters 85, 3313 (2000).
  • Cerf et al. [2002] N. J. Cerf, M. Bourennane, A. Karlsson, and N. Gisin, Security of quantum key distribution using d-level systems, Physical review letters 88, 127902 (2002).
  • Blok et al. [2021] M. S. Blok, V. Ramasesh, T. Schuster, K. O’Brien, J. 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).
  • Rosenblum et al. [2018] S. Rosenblum, P. Reinhold, M. Mirrahimi, L. Jiang, L. Frunzio, and R. J. Schoelkopf, Fault-tolerant detection of a quantum error, Science 361, 266 (2018).
  • Chow et al. [2013] J. M. Chow, J. M. Gambetta, A. W. Cross, S. T. Merkel, C. Rigetti, and M. Steffen, Microwave-activated conditional-phase gate for superconducting qubits, New Journal of Physics 15, 115012 (2013).
  • Wu et al. [2020] X. Wu, S. L. Tomarken, N. A. Petersson, L. A. Martinez, Y. J. Rosen, and J. L. DuBois, High-fidelity software-defined quantum logic on a superconducting qudit, Physical Review Letters 125, 170502 (2020).
  • Ashhab et al. [2021] S. Ashhab, F. Yoshihara, T. Fuse, N. Yamamoto, A. Lupascu, and K. Semba, Speed limits for quantum gates with weakly anharmonic qubits, arXiv preprint arXiv:2109.01469  (2021).
  • Varbanov et al. [2020] B. M. Varbanov, F. Battistel, B. M. Tarasinski, V. P. Ostroukh, T. E. O’Brien, L. DiCarlo, and B. M. Terhal, Leakage detection for a transmon-based surface code, npj Quantum Information 6, 1 (2020).
  • Ghosh et al. [2013] J. Ghosh, A. G. Fowler, J. M. Martinis, and M. R. Geller, Understanding the effects of leakage in superconducting quantum-error-detection circuits, Physical Review A 88, 062329 (2013).
  • Fowler [2013] A. G. Fowler, Coping with qubit leakage in topological codes, Physical Review A 88, 042308 (2013).
  • Motzoi et al. [2009] F. Motzoi, J. M. Gambetta, P. Rebentrost, and F. K. Wilhelm, Simple pulses for elimination of leakage in weakly nonlinear qubits, Physical review letters 103, 110501 (2009).
  • Werninghaus et al. [2021] M. Werninghaus, D. J. Egger, F. Roy, S. Machnes, F. K. Wilhelm, and S. Filipp, Leakage reduction in fast superconducting qubit gates via optimal control, npj Quantum Information 7, 1 (2021).
  • Economou and Barnes [2015] S. E. Economou and E. Barnes, Analytical approach to swift nonleaky entangling gates in superconducting qubits, Physical Review B 91, 161405 (2015).
  • Deng et al. [2017] X.-H. Deng, E. Barnes, and S. E. Economou, Robustness of error-suppressing entangling gates in cavity-coupled transmon qubits, Physical Review B 96, 035441 (2017).
  • Peruzzo et al. [2014] A. Peruzzo, J. McClean, P. Shadbolt, M.-H. Yung, X.-Q. Zhou, P. J. Love, A. Aspuru-Guzik, and J. L. O’brien, A variational eigenvalue solver on a photonic quantum processor, Nature communications 5, 1 (2014).
  • Sharma et al. [2020] K. Sharma, S. Khatri, M. Cerezo, and P. J. Coles, Noise resilience of variational quantum compiling, New Journal of Physics 22, 043006 (2020).
  • McArdle et al. [2020] S. McArdle, S. Endo, A. Aspuru-Guzik, S. C. Benjamin, and X. Yuan, Quantum computational chemistry, Reviews of Modern Physics 92, 015003 (2020).
  • Farhi et al. [2014] E. Farhi, J. Goldstone, and S. Gutmann, A quantum approximate optimization algorithm, arXiv preprint arXiv:1411.4028  (2014).
  • Farhi and Harrow [2016] E. Farhi and A. W. Harrow, Quantum supremacy through the quantum approximate optimization algorithm, arXiv preprint arXiv:1602.07674  (2016).
  • Farhi and Neven [2018] E. Farhi and H. Neven, Classification with quantum neural networks on near term processors, arXiv preprint arXiv:1802.06002  (2018).
  • Hou et al. [2021] X. Hou, G. Zhou, Q. Li, S. Jin, and X. Wang, A universal duplication-free quantum neural network, arXiv preprint arXiv:2106.13211  (2021).
  • Schuld et al. [2014] M. Schuld, I. Sinayskiy, and F. Petruccione, The quest for a quantum neural network, Quantum Information Processing 13, 2567 (2014).
  • Jeswal and Chakraverty [2019] S. Jeswal and S. Chakraverty, Recent developments and applications in quantum neural network: a review, Archives of Computational Methods in Engineering 26, 793 (2019).
  • Meitei et al. [2021a] O. R. Meitei, B. T. Gard, G. S. Barron, D. P. Pappas, S. E. Economou, E. Barnes, and N. J. Mayhall, Gate-free state preparation for fast variational quantum eigensolver simulations, npj Quantum Information 7, 155 (2021a).
  • Magann et al. [2021] A. B. Magann, C. Arenz, M. D. Grace, T.-S. Ho, R. L. Kosut, J. R. McClean, H. A. Rabitz, and M. Sarovar, From pulses to circuits and back again: A quantum optimal control perspective on variational quantum algorithms, PRX Quantum 2, 010101 (2021).
  • Yang et al. [2017] Z.-C. Yang, A. Rahmani, A. Shabani, H. Neven, and C. Chamon, Optimizing variational quantum algorithms using pontryagin’s minimum principle, Physical Review X 7, 021027 (2017).
  • Choquette et al. [2021] A. Choquette, A. Di Paolo, P. K. Barkoutsos, D. Sénéchal, I. Tavernelli, and A. Blais, Quantum-optimal-control-inspired ansatz for variational quantum algorithms, Physical Review Research 3, 023092 (2021).
  • Larocca et al. [2021a] M. Larocca, P. Czarnik, K. Sharma, G. Muraleedharan, P. J. Coles, and M. Cerezo, Diagnosing barren plateaus with tools from quantum optimal control, arXiv preprint arXiv:2105.14377  (2021a).
  • Russell et al. [2017] B. Russell, H. Rabitz, and R.-B. Wu, Control landscapes are almost always trap free: A geometric assessment, Journal of Physics A: Mathematical and Theoretical 50, 205302 (2017).
  • Larocca et al. [2021b] M. Larocca, N. Ju, D. García-Martín, P. J. Coles, and M. Cerezo, Theory of overparametrization in quantum neural networks, arXiv preprint arXiv:2109.11676  (2021b).
  • Deffner and Campbell [2017] S. Deffner and S. Campbell, Quantum speed limits: from heisenberg’s uncertainty principle to optimal quantum control, Journal of Physics A: Mathematical and Theoretical 50, 453001 (2017).
  • Poggi [2020] P. M. Poggi, Analysis of lower bounds for quantum control times and their relation to the quantum speed limit, arXiv preprint arXiv:2002.11147  (2020).
  • Tibbetts et al. [2012] K. W. M. Tibbetts, C. Brif, M. D. Grace, A. Donovan, D. L. Hocker, T.-S. Ho, R.-B. Wu, and H. Rabitz, Exploring the tradeoff between fidelity and time optimal control of quantum unitary transformations, Physical Review A 86, 062309 (2012).
  • Brady et al. [2021] L. T. Brady, C. L. Baldwin, A. Bapat, Y. Kharkov, and A. V. Gorshkov, Optimal protocols in quantum annealing and quantum approximate optimization algorithm problems, Physical Review Letters 126, 070505 (2021).
  • Lin et al. [2019] C. Lin, Y. Wang, G. Kolesov, and U. Kalabić, Application of pontryagin’s minimum principle to grover’s quantum search problem, Physical Review A 100, 022327 (2019).
  • Meitei et al. [2021b] O. R. Meitei, B. T. Gard, G. S. Barron, D. P. Pappas, S. E. Economou, E. Barnes, and N. J. Mayhall, Ctrl-vqe code, https://github.com/mayhallgroup/ctrlq (2021b).
  • Sun et al. [2018] Q. Sun, T. C. Berkelbach, N. S. Blunt, G. H. Booth, S. Guo, Z. Li, J. Liu, J. D. McClain, E. R. Sayfutyarova, S. Sharma, et al., Pyscf: the python-based simulations of chemistry framework, Wiley Interdisciplinary Reviews: Computational Molecular Science 8, e1340 (2018).
  • Ku et al. [2017] H. S. Ku, J. L. Long, X. Wu, M. Bal, R. E. Lake, E. Barnes, S. E. Economou, and D. P. Pappas, Single qubit operations using microwave hyperbolic secant pulses, Phys. Rev. A 96, 042339 (2017).
  • Boscain et al. [2021] U. Boscain, M. Sigalotti, and D. Sugny, Introduction to the pontryagin maximum principle for quantum optimal control, PRX Quantum 2, 030203 (2021).
  • Note [1] Conventionally known as control Hamiltonian but we use the edited name to avoid confusion with drive Hamiltonian (H¯C\bar{H}_{C}).
  • Liberzon [2011] D. Liberzon, Calculus of variations and optimal control theory (Princeton university press, 2011).
  • Note [2] Here we use 100 pulse segments and 1000 Trotter steps. We tested smaller and larger Trotter steps and found no appreciable change in results.