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

Simulating nonnative cubic interactions on noisy quantum machines

Yuan Shi [email protected]    Alessandro R. Castelli    Xian Wu    Ilon Joseph    Vasily Geyko    Frank R. Graziani    Stephen B. Libby    Jeffrey B. Parker    Yaniv J. Rosen    Luis A. Martinez    Jonathan L DuBois Lawrence Livermore National Laboratory, Livermore, California 94551, USA
Abstract

As a milestone for general-purpose computing machines, we demonstrate that quantum processors can be programmed to efficiently simulate dynamics that are not native to the hardware. Moreover, on noisy devices without error correction, we show that simulation results are significantly improved when the quantum program is compiled using modular gates instead of a restricted set of standard gates. We demonstrate the general methodology by solving a cubic interaction problem, which appears in nonlinear optics, gauge theories, as well as plasma and fluid dynamics. To encode the nonnative Hamiltonian evolution, we decompose the Hilbert space into a direct sum of invariant subspaces in which the nonlinear problem is mapped to a finite-dimensional Hamiltonian simulation problem. In a three-states example, the resultant unitary evolution is realized by a product of \sim20 standard gates, using which \sim10 simulation steps can be carried out on state-of-the-art quantum hardware before results are corrupted by decoherence. In comparison, the simulation depth is improved by more than an order of magnitude when the unitary evolution is realized as a single cubic gate, which is compiled directly using optimal control. Alternatively, parametric gates may also be compiled by interpolating control pulses. Modular gates thus obtained provide high-fidelity building blocks for quantum Hamiltonian simulations.

Ideal quantum computers can in principle perform arbitrary unitary operations. However, a lengthy sequence of standard gates is required when the unitary operation results from Hamiltonian evolution that is not native to the hardware. In this case, implementing the operation becomes infeasible when the gate depth exceeds the coherence limit of noisy devices. While this issue may ultimately be resolved by quantum error correction [1, 2, 3, 4, 5, 6], in this letter, we show that efficient compilation plays a key role in realizing quantum computing. When compiling a quantum program using modular gates [7], the end performance of quantum hardware is significantly improved due to the reduction of gate depth. In particular, we show that for cubic interactions, which are essential for nonlinear optics [8], gauge theories [9], as well as plasma and fluid dynamics [10], modular gates enable useful Hamiltonian simulations on present-day noisy hardware, which usually lacks native cubic couplings [11, 12].

To close a major gap in the literature, we consider the following underexamined cubic Hamiltonian (=1\hbar=1) that governs the lowest-order nonlinear interactions in quantum optics and plasma dynamics:

H=igA1A2A3igA1A2A3,H=igA_{1}^{\dagger}A_{2}A_{3}-ig^{*}A_{1}A_{2}^{\dagger}A_{3}^{\dagger}, (1)

where gg is the coupling coefficient and [Aj,Al]=δjl[A_{j},A_{l}^{\dagger}]=\delta_{jl}. The resulting Heisenberg equations are special cases of three-wave equations

dtA1\displaystyle d_{t}A_{1} =\displaystyle= gA2A3,\displaystyle gA_{2}A_{3}, (2)
dtA2\displaystyle d_{t}A_{2} =\displaystyle= gA1A3,\displaystyle-g^{*}A_{1}A_{3}^{\dagger}, (3)
dtA3\displaystyle d_{t}A_{3} =\displaystyle= gA1A2,\displaystyle-g^{*}A_{1}A_{2}^{\dagger}, (4)

where dt=t+𝐯jd_{t}=\partial_{t}+\mathbf{v}_{j}\cdot\nabla is the convective derivative at the wave group velocity 𝐯j=ωj/𝐤j\mathbf{v}_{j}=\partial\omega_{j}/\partial\mathbf{k}_{j}. As example applications, Eqs. (2)-(4) are solved to (i) optimize the input seed pulse shape when designing nonlinear optical systems in the pump depletion regime [13, 14, 15]; (ii) determine the turbulence spectrum in the wave-kinetic approach to weak turbulence [16], and (iii) compute the energy delivered to fusion fuel when drive lasers exchange energy in plasmas [17, 18, 19]. More broadly speaking, cubic terms are common for nonlinear problems, for which cubic gates that enacts Eq. (1) serve as basic building blocks. Cubic gates, in conjunction with quartic gates that are usually natively available on hardware, allow one to simulate a majority of problems in physics.

In the usual approach to quantum emulation, terms in the hardware Hamiltonian H0=k=1mHkH_{0}=\sum_{k=1}^{m}H_{k} are turned on or off to generate the unitary evolution due a targeted H=k=1makHkH=\sum_{k=1}^{m}a_{k}H_{k}, where aka_{k} are real coefficients. Explicitly, using the Lie-Trotter-Suzuki expansion, exp(iHT)=limn[k=1mUk(akT/n)]n\exp(-iHT)=\lim_{n\rightarrow\infty}[\prod_{k=1}^{m}U_{k}(a_{k}T/n)]^{n}, where Uk(t)=exp(iHkt)U_{k}(t)=\exp(-iH_{k}t). Difficulty arises when HH contains terms that are not natively available in H0H_{0}. Although cubic couplings may be realized natively with special-purpose hardware [20, 21], direct emulations have two major limitations. First, qubits can only be in a superposition of |0|0\rangle and |1|1\rangle states, whereas a large amplitude wave, such as that of a laser, involves many photons. Therefore, directly representing wave states by qubit states is inefficient. Second, parameters of qubits may not be easily adjustable [22] to generate Eq. (1). Since quantum hardware likely contains additional terms in H0H_{0} that cannot be completely turned off, direct emulation is demanding of the hardware tunability and may not be realizable for problems of practical interest.

In comparison, simulation that encodes the interaction at software level with reusable modular gates is more versatile and allows quantum processors to operate as general-purpose computing machines, which do not need to be rebuilt to solve a different problem. In what follows, we construct an efficient algorithm in the action space to enact the nonnative cubic Hamiltonian evolution. Here, the term “action” originates from the action-angle variables of the slow-envelope approximation, which is commonly used to derive the three-wave equations. For Eq. (1), two independent action operators that commute with HH are

S2=n1+n3,\displaystyle S_{2}=n_{1}+n_{3}, (5)
S3=n1+n2,\displaystyle S_{3}=n_{1}+n_{2}, (6)

where nj=AjAjn_{j}=A_{j}^{\dagger}A_{j} is the number operator. The simultaneous eigenspace of HH, S2S_{2} and S3S_{3} has dimension D=min(s2,s3)+1D=\min(s_{2},s_{3})+1, where the integer sj0s_{j}\geq 0 is the eigenvalue of SjS_{j}. Without loss of generality, suppose s2s3s_{2}\leq s_{3}, which breaks the 232\leftrightarrow 3 symmetry. Then, in the Fock basis |n1,n2,n3|n_{1},n_{2},n_{3}\rangle, any state in the DD-dimensional subspace VV is spanned by

|ψ=j=0s2αj|s2j,s3s2+j,j,|\psi\rangle=\sum_{j=0}^{s_{2}}\alpha_{j}|s_{2}-j,s_{3}-s_{2}+j,j\rangle, (7)

where the expansion coefficients satisfy the normalization condition j|αj|2=1\sum_{j}|\alpha_{j}|^{2}=1. It is easy to check that Sj|ψ=sj|ψS_{j}|\psi\rangle=s_{j}|\psi\rangle. Moreover, it is important to recognize that VV is a closed subspace under the action of HH, namely, H|ψVH|\psi\rangle\in V whenever |ψV|\psi\rangle\in V. Therefore, the total Hilbert space can be decomposed into a direct sum of invariant subspaces, and the Hamiltonian dynamics is confined within the subspaces specified by initial conditions. Notice that using D=s2+1D=s_{2}+1 levels, the above mapping can efficiently represent s3s2s_{3}\gg s_{2} photons.

Moreover, using the action-space algorithm, the nonlinear three-wave problem is mapped to a finite-dimensional linear Hamiltonian simulation problem. In the Schrödinger picture, it|ψ=H|ψi\partial_{t}|\psi\rangle=H|\psi\rangle becomes a system of equations for αj(t)\alpha_{j}(t)

itαj=ighj+12αj+1ighj12αj1,\displaystyle i\partial_{t}\alpha_{j}=igh_{j+\frac{1}{2}}\alpha_{j+1}-ig^{*}h_{j-\frac{1}{2}}\alpha_{j-1}, (8)

where hj12=j(s2+1j)(s3s2+j)h_{j-\frac{1}{2}}=\sqrt{j(s_{2}+1-j)(s_{3}-s_{2}+j)} with h12=hD+12=0h_{-\frac{1}{2}}=h_{D+\frac{1}{2}}=0. In other words, HH is represented by a block tridiagonal matrix with zero diagonal elements. The above mapping utilizes the special form of HH, but similar algorithms may be constructed for other nonlinear problems when the Hamiltonian has special symmetries.

Once the expansion coefficients are solved for given initial conditions, observables of interest are evaluated by O(D)O(D) classical operations during post processing. For example, occupation numbers of the three waves are given by n1=j=0s2(s2j)|αj|2\langle n_{1}\rangle=\sum_{j=0}^{s_{2}}(s_{2}-j)|\alpha_{j}|^{2}, n2=j=0s2(s3s2+j)|αj|2\langle n_{2}\rangle=\sum_{j=0}^{s_{2}}(s_{3}-s_{2}+j)|\alpha_{j}|^{2}, and n3=j=0s2j|αj|2\langle n_{3}\rangle=\sum_{j=0}^{s_{2}}j|\alpha_{j}|^{2}. Higher-order cumulants, thereby all possible expectation values of interest, can be obtained similarly, and the quantum three-wave problem is then solved.

Before solving the Schrödinger equation, it is instructive to also analyze the cubic interactions in the Heisenberg picture. The number operators satisfy

t2n1\displaystyle\partial_{t}^{2}n_{1} =\displaystyle= t2n2=t2n3\displaystyle-\partial_{t}^{2}n_{2}=-\partial_{t}^{2}n_{3}
=\displaystyle= 2|g|2[s2s3(2s2+2s3+1)n1+3n12].\displaystyle 2|g|^{2}\big{[}s_{2}s_{3}-(2s_{2}+2s_{3}+1)n_{1}+3n_{1}^{2}\big{]}.

In comparison, the classical wave action Ij=|Aj|2I_{j}=|A_{j}|^{2}, where AjA_{j} is treated as a complex number, satisfies

t2I1\displaystyle\partial_{t}^{2}I_{1} =\displaystyle= t2I2=t2I3\displaystyle-\partial_{t}^{2}I_{2}=-\partial_{t}^{2}I_{3}
=\displaystyle= 2|g|2[s2s3(2s2+2s3)I1+3I12].\displaystyle 2|g|^{2}\big{[}s_{2}s_{3}-(2s_{2}+2s_{3})I_{1}+3I_{1}^{2}\big{]}.

When identifying IjnjI_{j}\simeq\langle n_{j}\rangle, the difference between quantum and classical systems is proportional to 3(n12n12)n13(\langle n_{1}^{2}\rangle-\langle n_{1}\rangle^{2})-\langle n_{1}\rangle. The term in the parenthesis is exactly zero for states with Poisson statistics, and is usually small compared to other terms when the state is a well-localized semi-classical state. The second term is comparatively small when s2s_{2} and s3s_{3} are large, whereby stimulated processes, which depend quadratically on the number of photons, dominate spontaneous emission, which depends linearly on the number of photons. When both terms are small, the system is in the semiclassical regime [23], where the quantum and classical solutions are similar. In the opposite regime, the behaviors of the two systems are disparate. For example, in the thermodynamic limit where occupation numbers satisfy the Boltzmann distribution |αj|2=1/D|\alpha_{j}|^{2}=1/D, the quantum system is stationary, whereas the classical system, which is solved by Jacobi elliptic functions [24, 25], is nonstationary.

In the following, we demonstrate the algorithm using three levels, or equivalently, the |00|00\rangle, |01|01\rangle, and |10|10\rangle states of two qubits. The allowable dimension is D3D\leq 3, and we take s2=2s_{2}=2 and s3=s2s_{3}=s\geq 2. Then, a natural mapping is |2,s2,0=(1,0,0)T|2,s-2,0\rangle=(1,0,0)^{\text{T}}, |1,s1,1=(0,1,0)T|1,s-1,1\rangle=(0,1,0)^{\text{T}}, and |0,s,2=(0,0,1)T|0,s,2\rangle=(0,0,1)^{\text{T}}. The normalized Hamiltonian h=H/|g|h=H/|g|, when restricted to the invariant subspace, is

h(θ,s)=(0eiθ2(s1)0eiθ2(s1)0eiθ2s0eiθ2s0),h(\theta,s)=\left(\begin{array}[]{ccc}0&e^{i\theta}\sqrt{2(s-1)}&0\\ e^{-i\theta}\sqrt{2(s-1)}&0&e^{i\theta}\sqrt{2s}\\ 0&e^{-i\theta}\sqrt{2s}&0\end{array}\right), (11)

where exp(iθ)=ig/|g|\exp(i\theta)=ig/|g|. It is worth emphasizing that the general case is associated with a D×DD\times D matrix, where D2nD\simeq 2^{n} with nn being the number of qubits needed. In other words, DD needs not be three, but the Hamiltonian matrix is always tridiagonal. For the three-level problem, Eq. (11) is exponentiated analytically 111See Supplemental Material for the unitary matrix of the three-level problem. to determine the unitary time-evolution operator U(Δτ,θ,s)=exp[ih(θ,s)Δτ]U(\Delta\tau,\theta,s)=\exp[-ih(\theta,s)\Delta\tau]. This unitary matrix, for a given Δτ=|g|Δt\Delta\tau=|g|\Delta t, is an input for quantum compilers. Notice that the quantum hardware may not be able to track U(τ)U(\tau) continuously, because the cubic Hamiltonian has odd parity whereas the native Hamiltonian may not be. Nevertheless, at discrete time steps, U(Δτ)U(\Delta\tau) can be realized just like other complex unitary operators.

Refer to caption
Figure 1: (a) Optimized control pulse ϵ(t)\epsilon(t) for achieving the unitary operator U(0.1,π/2,2)U(0.1,\pi/2,2) on a transmon qudit. (b) The Fourier transform ϵ^(f)\hat{\epsilon}(f) shows low-frequency modulations and peaks at harmonics of f01f12f_{01}-f_{12}0.23\approx 0.23 GHz, where fijf_{ij} is the qudit |i|j|i\rangle\leftrightarrow|j\rangle transition frequency. (c) Experimentally estimated occupations |αj|2|\alpha_{j}|^{2} (dots) during the control pulse application match solutions to the master equation (lines). At the end of the pulse application, the targeted transition probabilities PP (arrows) are attained.

To compile the three-level unitary operator into a hardware executable form, the standard approach approximates it using a sequence of universal gates. As an example of state-of-the-art quantum cloud services, we implement the unitary operator on Aspen-4-2Q-A of Rigetti Computing [27, 28]. The probabilistic Quil compiler converts U(Δτ,θ,s)U(\Delta\tau,\theta,s) to a sequence of 20\sim 20 native gates, including single-qubit Pauli gates and two CZ gates 222See Supplemental Material for an example the Quil program..

As a more hardware-efficient compilation approach, we realize the unitary operator as a single gate using optimal control: Instead of preparing control pulses for standard gates and then using these gates to approximate a targeted operation, we directly optimize a control pulse for the targeted operation. The device we use is a transmon placed inside a 3D superconducting aluminum microwave cavity, whose details are described in [30]. In addition to enhancing the coherence time, this many-qudit single-readout architecture reduces the number of control lines required to access larger computational spaces, and allows simultaneous control of multiple subsystems without relying on their crosstalk [31, 32, 33, 34, 35, 36]. The control Hamiltonian is [37]

Hc(t)j(cj+cj)(ϵjeiΩjt+ϵjeiΩjt),H_{c}(t)\simeq\sum_{j}(c_{j}+c^{\dagger}_{j})\Big{(}\epsilon_{j}e^{-i\Omega_{j}t}+\epsilon_{j}^{*}e^{i\Omega_{j}t}\Big{)}, (12)

where the complex-valued ϵj(t)\epsilon_{j}(t) is the slowly-varying envelope of a microwave field whose carrier frequency is Ωj\Omega_{j}, and cjc_{j} is its control operator 333See Supplemental Material for an explicit expression of the control operator.. The time dependence of ϵj(t)\epsilon_{j}(t) may be engineered such that

𝒯ei0T𝑑t[H0+Hc(t)]eiHΔt,\mathcal{T}e^{-i\int_{0}^{T}dt[H_{0}+H_{c}(t)]}\simeq e^{-iH\Delta t}, (13)

where 𝒯\mathcal{T} is the time-ordering operator, TT is the length of the control pulse, and H0H_{0} is the hardware Hamiltonian. The control is not unique but can be implemented using direct digital synthesis [39, 40] and designed using numerical optimization [41, 42, 43, 44] for small problems. For larger problems, the tridiagonal Hamiltonian may be decomposed as H=k=1mHkH=\sum_{k=1}^{m}H_{k}, where each HkH_{k} is block diagonal. Each block in HkH_{k}, which is small enough and decoupled from other blocks at the hardware level, can now be enacted using numerically optimized control pulse prepared for the subsystem. Subsequently, the full Hamiltonian evolution due to HH can be approximated using the Lie-Trotter-Suzuki expansion.

As a test case, we construct a control pulse to achieve U(0.1,π/2,2)U(0.1,\pi/2,2). Based on a thorough characterization that provides parameters of the native Hamiltonian [30], the control pulse is generated in the |0|1|0\rangle\leftrightarrow|1\rangle rotating frame using optimize_pulse_unitary in QuTIP [45, 46], which calls an optimizer that uses the gradient ascent algorithms to iteratively update the control pulse to minimize the fidelity error. The optimization uses zero as the initial guess, and is constrained by |ϵ|<0.03|\epsilon|<0.03. Two forbidden levels beyond the D=3D=3 levels are included in the optimization to suppress possible state leakages at the end of the control pulse. The pulse duration (T=150T=150 ns) is much shorter than the coherence time but long enough to allow the lowest three levels of the transmon, which has an anharmonicity of 0.23\sim 0.23 GHz, to be separately addressed. The control pulse ϵ(t)\epsilon(t) is shown in Fig. 1(a), and the absolute value of its Fourier transform ϵ^(f)\hat{\epsilon}(f) is shown in Fig. 1(b). The optimization takes 10\sim 10 iterations to reach convergence with an error target of 10510^{-5}.

When applying to the hardware, optimized control pulses are (i) concatenated in the rotating frame, (ii) transformed to the lab frame by mixing with a carrier wave at the qudit |0|1|0\rangle\leftrightarrow|1\rangle transition frequency f014.1f_{01}\approx 4.1 GHz, and (iii) compensated for spectral filtering effects of the hardware [30]. The final waveform is synthesized using an arbitrary waveform generator at 32-GHz sampling rate, and is sent to the qudit inside a dilution refrigerator through a series of attenuators and a band-pass filter. Occupations of the three levels are measured using dispersive readout [31, 47, 30]. The measured in-phase and quadrature signals are used to classify the states of the qudit, and the classification errors are partially corrected using a confusion matrix [30], which may cause unphysical occupations slightly above 1 or below 0. The estimated occupations during the pulse application are shown in Fig. 1(c) when the qudit is initialized in the ground state, which has minimal state preparation error. By taking into account hardware-specific decay and dephasing 444See Supplemental Material for the noise model based on the Lindblad master equation.. the measurement results (dots) are well explained by numerical solutions (lines) of the Lindblad master equation [49, 50, 51, 52]. At the end of the control pulse, the intended unitary operator is realized, and the targeted transition probabilities (arrows) are attained. Using a modified process tomography [30], we estimate the process matrix of the cubic gate, which yields an average gate fidelity of 99.3%.

Refer to caption
Figure 2: Occupations of |0|0\rangle, |1|1\rangle, and |2|2\rangle states after repeating U(0.2,π/2,2)U(0.2,\pi/2,2) for NN times. By compiling UU as a single gate (blue), the simulation depth is improved by more than an order of magnitude compared to the standard compilation approach (cyan), which approximates each UU by 20\sim 20 native gates. Deviations from the exact results (orange) are explained by a noisy model based on the Lindblad master equation (black).

To compare the performance of the two compilation approaches, we repeatedly apply the unitary operator to simulate Hamiltonian time evolution. With the quantum processors initialized in the ground state, we read out occupations after repeating U(0.2,π/2,2)U(0.2,\pi/2,2) for NN times, and compare the results with the exact solutions (Fig. 2, orange). Using a modular cubic gate (T=80T=80 ns), the experimental results (blue) follow the exact solutions up to N100N\gtrsim 100, and the deviations are explained by the master equation solutions (black) using our noise model [48]. In comparison, using a sequence of standard gates, the experimental results (cyan) track the exact solutions only up to N10N\sim 10, which fails to capture even a single period of the nonlinear oscillation. However, it is worth emphasizing that the hardware performance is comparable, and both quantum processors can perform 100\sim 100 gates with high fidelity. The difference in results mostly comes from how the unitary operator is realized: Using the standard compilation approach, each UU is realized by 20\sim 20 native gates, whereas using the modular compilation approach, UU is realized as a single gate. By reducing the gate depth, the achievable simulation depth is thus significantly increased, which in this case qualitatively changes whether a useful Hamiltonian simulation is feasible or not.

Refer to caption
Figure 3: Interpolated control pulses drive high-fidelity cubic gates for a wide range of ss parameters with the qudit initialized in |0|0\rangle, |1|1\rangle, and |2|2\rangle states. The interpolation only uses optimized control pulses at s=2s=2 and s=s=\infty at fixed τ2s=0.2\tau\sqrt{2s}=0.2 and θ=π/2\theta=\pi/2.

While offering better performance, modular gates are not more costly to compile on classical computers. The standard compilation approach first prepares a universal set of gates that are executable on quantum hardware, and then approximates a targeted D×DD\times D unitary operator UU using these native gates. Apart from the cost of preparing the gate set, compiling an arbitrary UU requires O(D2)=O(22n)O(D^{2})=O(2^{2n}) native gates [53]. Exponential speedup is possible only when U=exp(iHt)U=\exp(-iHt) and HH is native to the hardware. When HH is not native, sophisticated quantum algorithms [54, 55, 56, 57, 58] still have favorable complexity scaling, but are not readily implementable on hardware. In comparison, directly compiling UU as a modular gate using optimal control requires O(D2)O(D^{2}) classical steps as well. However, once the control pulse is obtained, the unitary operator is enacted much more efficiently as a single pulse. Then, compared with classical computers, which always take O(D2)O(D^{2}) operations to compute UψU\psi, quantum processors may be used as special-purpose hardware accelerators for matrix multiplications. Moreover, the overhead of modular compilation may be amortized when the same UU is needed repeatedly or when a family of UU can be compiled without expensive numerical optimization.

In fact, high-fidelity parametric gates may be compiled cheaply using interpolation. For example, consider a one-parameter family of the cubic Hamiltonian h(s)=2s[(1ξ)K(2)+(ξ1/2)K()]/(11/2)h(s)=\sqrt{2s}[(1-\xi)K(2)+(\xi-1/\sqrt{2})K(\infty)]/(1-1/\sqrt{2}), where ξ(s)=11/s\xi(s)=\sqrt{1-1/s}, and the nonzero elements of the symmetric 3×33\times 3 matrix KK are K12(s)=ξK_{12}(s)=\xi and K23(s)=1K_{23}(s)=1. Motivated by this form, the interpolated control pulse for s2s\geq 2 at each time slice is taken to be ϵI(s)=[(1ξ)ϵO(2)+(ξ1/2)ϵO()]/(11/2)\epsilon_{\text{I}}(s)=[(1-\xi)\epsilon_{\text{O}}(2)+(\xi-1/\sqrt{2})\epsilon_{\text{O}}(\infty)]/(1-1/\sqrt{2}), where ϵO(s)\epsilon_{\text{O}}(s) denotes the optimized pulse and τ2s=0.2\tau\sqrt{2s}=0.2 is held a constant. In other words, we interpolate the control pulse using the same formula for interpolating the Hamiltonian. To test this scheme, we numerically solve the master equation to obtain the gate fidelity, which is defined by [53] F(ρ,σ)=trρ1/2σρ1/2F(\rho,\sigma)=\text{tr}\sqrt{\rho^{1/2}\sigma\rho^{1/2}}, where ρ\rho and σ\sigma are the density matrices attained using optimized and interpolated pulses after one gate application. To optimize control pulses for ρ(s)\rho(s), ϵO(2)\epsilon_{\text{O}}(2) is used as the initial guess. Although this compilation shortcut is not proven to work in general, the interpolated pulses are able to drive high-fidelity results for the three-level problem (Fig. 3).

In summary, we demonstrate that nonnative cubic interactions can be programmed entirely at the software level using an efficient action-space algorithm. Moreover, modular compilation reduces the requisite gate depth and qualitatively improves performance on noisy hardware. The resultant high-fidelity gates provide necessary building blocks for Hamiltonian simulations of a large class of nonlinear problems.

Acknowledgements.
We would like to thank Rigetti Computing for providing access to their 16Q Aspen-4 processor, where we used lattice Aspen-4-2Q-A. We thank Max D. Porter for verifying numerical results in Figs. 1 and 2. Y. S. would like to thank Eric T. Holland and Hong Qin for helpful discussions. This work was performed under the auspices of US Department of Energy (DOE) by LLNL under Contract DE-AC52-07NA27344. The experimental work was supported by the DOE Office of Fusion Energy Sciences “Quantum Leap for Fusion Energy Sciences” under project FWP-SCW1680, and the theory work was supported by LLNL-LDRD under Project 19-FS-072. The QuDIT hardware was funded under the National Nuclear Security Administration (NNSA) Advanced Simulation and Computing (ASC) “Beyond Moore’s Law” quantum program under NA-ASC-127R-16 and US DOE, Office of Science, Office of Advanced Scientific Computing Research, Quantum Testbed Pathfinder Program under Award 2017-LLNL-SCW1631. Y. S. was supported by the Lawrence Fellowship through LLNL-LDRD under Project 19-ERD-038.

References

  • Cory et al. [1998] D. G. Cory, M. D. Price, W. Maas, E. Knill, R. Laflamme, W. H. Zurek, T. F. Havel, and S. S. Somaroo, Experimental quantum error correction, Phys. Rev. Lett. 81, 2152 (1998).
  • Chiaverini et al. [2004] J. Chiaverini, D. Leibfried, T. Schaetz, M. D. Barrett, R. B. Blakestad, J. Britton, W. M. Itano, J. D. Jost, E. Knill, C. Langer, R. Ozeri, and D. J. Wineland, Realization of quantum error correction, Nature 432, 602 (2004).
  • Terhal [2015] B. M. Terhal, Quantum error correction for quantum memories, Rev. Mod. Phys. 87, 307 (2015).
  • Ofek et al. [2016] N. Ofek, A. Petrenko, R. Heeres, P. Reinhold, Z. Leghtas, B. Vlastakis, Y. Liu, L. Frunzio, S. Girvin, L. Jiang, et al., Extending the lifetime of a quantum bit with error correction in superconducting circuits, Nature 536, 441 (2016).
  • Reiter et al. [2017] F. Reiter, A. S. Sørensen, P. Zoller, and C. Muschik, Dissipative quantum error correction and application to quantum sensing with trapped ions, Nat. Commun. 8, 1822 (2017).
  • 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).
  • Shi et al. [2019] Y. Shi, N. Leung, P. Gokhale, Z. Rossi, D. I. Schuster, H. Hoffmann, and F. T. Chong, Optimized compilation of aggregated instructions for realistic quantum computers, in Proceedings of the Twenty-Fourth International Conference on Architectural Support for Programming Languages and Operating Systems (2019) pp. 1031–1044.
  • Bloembergen [1996] N. Bloembergen, Nonlinear Optics (World Scientific, 1996).
  • Yang and Mills [1954] C.-N. Yang and R. L. Mills, Conservation of isotopic spin and isotopic gauge invariance, Phys. Rev. 96, 191 (1954).
  • Davidson [1972] R. Davidson, Methods in Nonlinear Plasma Theory (Academic Press, 1972).
  • Häffner et al. [2008] H. Häffner, C. F. Roos, and R. Blatt, Quantum computing with trapped ions, Phys. Rep. 469, 155 (2008).
  • Clarke and Wilhelm [2008] J. Clarke and F. K. Wilhelm, Superconducting quantum bits, Nature 453, 1031 (2008).
  • Frantz and Nodvik [1963] L. M. Frantz and J. S. Nodvik, Theory of pulse propagation in a laser amplifier, J. Appl. Phys. 34, 2346 (1963).
  • Ahn et al. [2003] J. Ahn, A. Efimov, R. Averitt, and A. Taylor, Terahertz waveform synthesis via optical rectification of shaped ultrafast laser pulses, Opt. Express 11, 2486 (2003).
  • Brunton et al. [2012] G. Brunton, G. Erbert, D. Browning, and E. Tse, The shaping of a national ignition campaign pulsed waveform, Fusion Eng. Des. 87, 1940 (2012).
  • Zakharov et al. [2012] V. E. Zakharov, V. S. L’vov, and G. Falkovich, Kolmogorov spectra of turbulence I: Wave turbulence (Springer Science & Business Media, 2012).
  • Moody et al. [2012] J. D. Moody, P. Michel, L. Divol, R. L. Berger, E. Bond, D. K. Bradley, D. A. Callahan, E. L. Dewald, S. Dixit, M. J. Edwards, et al., Multistep redirection by cross-beam power transfer of ultrahigh-power lasers in a plasma, Nature Physics 8, 344 (2012).
  • Myatt et al. [2014] J. F. Myatt, J. Zhang, R. W. Short, A. V. Maximov, W. Seka, D. H. Froula, D. H. Edgell, D. T. Michel, I. V. Igumenshchev, D. E. Hinkel, et al., Multiple-beam laser–plasma interactions in inertial confinement fusion, Phys. Plasmas 21, 055501 (2014).
  • Shi et al. [2017] Y. Shi, H. Qin, and N. J. Fisch, Three-wave scattering in magnetized plasmas: From cold fluid to quantized lagrangian, Phys. Rev. E 96, 023204 (2017).
  • Bergeal et al. [2010a] N. Bergeal, F. Schackert, M. Metcalfe, R. Vijay, V. Manucharyan, L. Frunzio, D. Prober, R. Schoelkopf, S. Girvin, and M. Devoret, Phase-preserving amplification near the quantum limit with a josephson ring modulator, Nature 465, 64 (2010a).
  • Bergeal et al. [2010b] N. Bergeal, R. Vijay, V. Manucharyan, I. Siddiqi, R. Schoelkopf, S. Girvin, and M. Devoret, Analog information processing at the quantum limit with a josephson ring modulator, Nat. Phys. 6, 296 (2010b).
  • Chen et al. [2014] Y. Chen, C. Neill, P. Roushan, N. Leung, M. Fang, R. Barends, J. Kelly, B. Campbell, Z. Chen, B. Chiaro, A. Dunsworth, E. Jeffrey, A. Megrant, J. Y. Mutus, P. J. J. O’Malley, C. M. Quintana, D. Sank, A. Vainsencher, J. Wenner, T. C. White, M. R. Geller, A. N. Cleland, and J. M. Martinis, Qubit architecture with high coherence and fast tunable coupling, Phys. Rev. Lett. 113, 220502 (2014).
  • Jaynes and Cummings [1963] E. T. Jaynes and F. W. Cummings, Comparison of quantum and semiclassical radiation theories with application to the beam maser, Proc. IEEE 51, 89 (1963).
  • Jurkus and Robson [1960] A. Jurkus and P. Robson, Saturation effects in a travelling-wave parametric amplifier, Proceedings of the IEE-Part B: Electronic and Communication Engineering 107, 119 (1960).
  • Armstrong et al. [1962] J. A. Armstrong, N. Bloembergen, J. Ducuing, and P. S. Pershan, Interactions between light waves in a nonlinear dielectric, Phys. Rev. 127, 1918 (1962).
  • Note [1] See Supplemental Material for the unitary matrix of the three-level problem.
  • Karalekas et al. [2020] P. J. Karalekas, N. A. Tezak, E. C. Peterson, C. A. Ryan, M. P. da Silva, and R. S. Smith, A quantum-classical cloud platform optimized for variational hybrid algorithms, Quantum Sci. Technol. 5, 024003 (2020).
  • Smith et al. [2016] R. S. Smith, M. J. Curtis, and W. J. Zeng, A practical quantum instruction set architecture (2016), arXiv:1608.03355 [quant-ph] .
  • Note [2] See Supplemental Material for an example the Quil program.
  • 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, Phys. Rev. Lett. 125, 170502 (2020).
  • Blais et al. [2004] A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin, and R. J. Schoelkopf, Cavity quantum electrodynamics for superconducting electrical circuits: An architecture for quantum computation, Phys. Rev. A 69, 062320 (2004).
  • Wallraff et al. [2004] A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang, J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, Strong coupling of a single photon to a superconducting qubit using circuit quantum electrodynamics, Nature 431, 162 (2004).
  • Koch et al. [2007] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Charge-insensitive qubit design derived from the Cooper pair box, Phys. Rev. A 76, 042319 (2007).
  • Schreier et al. [2008] J. A. Schreier, A. A. Houck, J. Koch, D. I. Schuster, B. R. Johnson, J. M. Chow, J. M. Gambetta, J. Majer, L. Frunzio, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Suppressing charge noise decoherence in superconducting charge qubits, Phys. Rev. B 77, 180502 (2008).
  • Paik et al. [2011] H. Paik, D. I. Schuster, L. S. Bishop, G. Kirchmair, G. Catelani, A. P. Sears, B. R. Johnson, M. J. Reagor, L. Frunzio, L. I. Glazman, S. M. Girvin, M. H. Devoret, and R. J. Schoelkopf, Observation of high coherence in josephson junction qubits measured in a three-dimensional circuit QED architecture, Phys. Rev. Lett. 107, 240501 (2011).
  • Rigetti et al. [2012] C. Rigetti, J. M. Gambetta, S. Poletto, B. L. T. Plourde, J. M. Chow, A. D. Córcoles, J. A. Smolin, S. T. Merkel, J. R. Rozen, G. A. Keefe, M. B. Rothwell, M. B. Ketchen, and M. Steffen, Superconducting qubit in a waveguide cavity with a coherence time approaching 0.1 ms, Phys. Rev. B 86, 100506 (2012).
  • Gerry et al. [2005] C. Gerry, P. Knight, and P. L. Knight, Introductory quantum optics (Cambridge university press, 2005).
  • Note [3] See Supplemental Material for an explicit expression of the control operator.
  • Raftery et al. [2017] J. Raftery, A. Vrajitoarea, G. Zhang, Z. Leng, S. Srinivasan, and A. Houck, Direct digital synthesis of microwave waveforms for quantum computing, arXiv:1703.00942  (2017).
  • Heeres et al. [2017] R. W. Heeres, P. Reinhold, N. Ofek, L. Frunzio, L. Jiang, M. H. Devoret, and R. J. Schoelkopf, Implementing a universal gate set on a logical qubit encoded in an oscillator, Nat. Commun. 8, 94 (2017).
  • Khaneja et al. [2005] N. Khaneja, T. Reiss, C. Kehlet, T. Schulte-Herbrüggen, and S. J. Glaser, Optimal control of coupled spin dynamics: design of NMR pulse sequences by gradient ascent algorithms, J. Magn. Reson. 172, 296 (2005).
  • De Fouquieres et al. [2011] P. De Fouquieres, S. Schirmer, S. Glaser, and I. Kuprov, Second order gradient ascent pulse engineering, J. Magn. Reson. 212, 412 (2011).
  • Lloyd and Montangero [2014] S. Lloyd and S. Montangero, Information theoretical analysis of quantum optimal control, Phys. Rev. Lett. 113, 010502 (2014).
  • 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:2001.01013  (2020).
  • Johansson et al. [2013] J. Johansson, P. Nation, and F. Nori, Qutip 2: A python framework for the dynamics of open quantum systems, Comput. Phys. Commun. 184, 1234 (2013).
  • Johansson et al. [2012] J. R. Johansson, P. Nation, and F. Nori, Qutip: An open-source python framework for the dynamics of open quantum systems, Comp. Phys. Commun. 183, 1760 (2012).
  • Boissonneault et al. [2009] M. Boissonneault, J. M. Gambetta, and A. Blais, Dispersive regime of circuit QED: Photon-dependent qubit dephasing and relaxation rates, Phys. Rev. A 79, 013819 (2009).
  • Note [4] See Supplemental Material for the noise model based on the Lindblad master equation.
  • Louisell [1973] W. H. Louisell, Quantum statistical properties of radiation (Wiley New York, 1973).
  • Lindblad [1976] G. Lindblad, On the generators of quantum dynamical semigroups, Commun. Math. Phys. 48, 119 (1976).
  • Gorini et al. [1976] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan, Completely positive dynamical semigroups of n-level systems, J. Math. Phys. 17, 821 (1976).
  • Schlosshauer [2019] M. Schlosshauer, The quantum-to-classical transition and decoherence, arXiv:1404.2635  (2019).
  • Nielsen and Chuang [2010] M. A. Nielsen and I. Chuang, Quantum computation and quantum information (Cambridge University Press, 2010).
  • Childs et al. [2003] A. M. Childs, R. Cleve, E. Deotto, E. Farhi, S. Gutmann, and D. A. Spielman, Exponential algorithmic speedup by a quantum walk, in Proceedings of the thirty-fifth annual ACM symposium on Theory of computing (2003) pp. 59–68.
  • Szegedy [2004] M. Szegedy, Quantum speed-up of markov chain based algorithms, in 45th Annual IEEE Symposium on Foundations of Computer Science (IEEE, 2004) pp. 32–41.
  • Childs et al. [2013] A. M. Childs, D. Gosset, and Z. Webb, Universal computation by multiparticle quantum walk, Science 339, 791 (2013).
  • Berry et al. [2015] D. W. Berry, A. M. Childs, and R. Kothari, Hamiltonian simulation with nearly optimal dependence on all parameters, in 2015 IEEE 56th Annual Symposium on Foundations of Computer Science (IEEE, 2015) pp. 792–809.
  • Low and Chuang [2017] G. H. Low and I. L. Chuang, Optimal hamiltonian simulation by quantum signal processing, Phys. Rev. Lett. 118, 010501 (2017).
  • Note [5] Rigetti QCS is continuously upgraded and routinely calibrated. For data reported in the main text, the QCS was accessed on January 7th, 2020 at 12:30 PM.

Appendix A Unitary evolution of three levels

For the three-level problem, the Hamiltonian matrix [Eq. (11)] can be exponentiated analytically. The Schrödinger time evolution of the three levels is determined by the unitary matrix U=exp(ihτ)U=\exp(-ih\tau), which is given explicitly by

U(τ,θ,s)=((s1)cosλτ+s2s1ieiθs12s1sinλτe2iθs(s1)2s1(cosλτ1)ieiθs12s1sinλτcosλτieiθs2s1sinλτe2iθs(s1)2s1(cosλτ1)ieiθs2s1sinλτscosλτ+s12s1),U(\tau,\theta,s)=\left(\begin{array}[]{ccc}\frac{(s-1)\cos\lambda\tau+s}{2s-1}&-ie^{i\theta}\sqrt{\frac{s-1}{2s-1}}\sin\lambda\tau&e^{2i\theta}\frac{\sqrt{s(s-1)}}{2s-1}(\cos\lambda\tau-1)\\ -ie^{-i\theta}\sqrt{\frac{s-1}{2s-1}}\sin\lambda\tau&\cos\lambda\tau&-ie^{i\theta}\sqrt{\frac{s}{2s-1}}\sin\lambda\tau\\ e^{-2i\theta}\frac{\sqrt{s(s-1)}}{2s-1}(\cos\lambda\tau-1)&-ie^{-i\theta}\sqrt{\frac{s}{2s-1}}\sin\lambda\tau&\frac{s\cos\lambda\tau+s-1}{2s-1}\end{array}\right), (S.1)

where λ=2(2s1)\lambda=\sqrt{2(2s-1)} is the positive eigenvalue of the Hamiltonian. The above 3×33\times 3 unitary matrix can be embedded into a two-qubits system by acting it on the |00|00\rangle, |01|01\rangle, and |10|10\rangle states, while leaving the |11|11\rangle state invariant.

The above unitary matrix is an input for both the standard and the modular compilation approaches. Although it may seem that once the unitary matrix is known, the problem would have already been solved, it is worth noting that the cubic gate may be a subproblem of some more complex problems. For example, in radiation hydrodynamics, lasers couple via three-wave interactions for given plasma conditions, while plasma conditions evolve due to laser energy deposition. The self consistent equations may be solved using a splitting algorithm, whose solutions may be written as UNVNU1V1U0V0U_{N}V_{N}\dots U_{1}V_{1}U_{0}V_{0}, where UU’s are cubic gates and VV’s are gates that advance the plasma conditions. In problems of this type, knowing UU is an intermediate step towards solving the entire problem, and quantum computing may be used to effectively carry out the matrix multiplications by applying precompiled gates.

Appendix B Realizing cubic gates on Rigetti QCS

As an example of state-of-the-art quantum cloud services (QCS), Rigetti Computing [27, 28] offers a native gate set consists of single-qubit rotations Rx(θ),Rz(θ)\textsf{R}_{x}(\theta),\textsf{R}_{z}(\theta), and two-qubit CZ gates. The Quil compiler uses these gates to approximate other unitary operators using non-deterministic algorithms [28], which efficiently generate approximations for a given error tolerance. In particular, the unitary matrix U(Δτ,θ,s)U(\Delta\tau,\theta,s) can be declared as a user-defined gate via the DEFGATE directive in the pyQuil library [28]. After routing to two adjacent qubits on the quantum hardware, the probabilistic compiler typically converts this cubic gate to a sequence of 20\sim 20 native gates, including two CZ gates (Fig. S1). When directly repeating the gate in pyQuil, the compiler multiplies [U(Δτ)]N=U(NΔτ)[U(\Delta\tau)]^{N}=U(N\Delta\tau) and compiles for U(NΔτ)U(N\Delta\tau) instead, so the hardware performance is independent of NN. This default simplification, which offloads the burden of computation to classical computers, is suppressed by placing the gate sequence for U(Δτ)U(\Delta\tau) within PRAGMA PRESERVE_BLOCK and PRAGMA END_PRESERVE_BLOCK. In this way, [U(Δτ)]N[U(\Delta\tau)]^{N} is realized on Rigetti’s Aspen-4-2Q-A by applying a precompiled U(Δτ)U(\Delta\tau) for NN times, and the results 555Rigetti QCS is continuously upgraded and routinely calibrated. For data reported in the main text, the QCS was accessed on January 7th, 2020 at 12:30 PM. are shown in Fig. 2 of the main text.

Refer to caption
Figure S1: An example Quil program that implements U(0.2,π/2,2)U(0.2,\pi/2,2) on Rigetti’s Aspen-4. In this example, the cubic gate is approximated by 17 native gates. This gate sequence can be repeated NN times to realize UNU^{N} before the final readout.

Appendix C Noise modeling using Lindblad master equation

Realistic quantum computers are open systems, and coupling to the environment is inevitable especially during control and readout operations. The state of a quantum processor, which is mixed with the environment, may be characterized by its density matrix. Assuming processes in the environment are stationary and Markovian, then the time evolution of the density matrix may be described by the Lindblad master equation [49, 50, 51, 52]

tρ=i[ρ,H0+Hc(t)]+j(LjρLj12{LjLj,ρ}),\partial_{t}\rho=i[\rho,H_{0}+H_{c}(t)]+\sum_{j}\big{(}L_{j}\rho L_{j}^{\dagger}-\frac{1}{2}\{L_{j}^{\dagger}L_{j},\rho\}\big{)}, (S.2)

where we have used the unit =1\hbar=1. The first term is the unitary evolution due to the bare Hamiltonian H0H_{0} of the quantum hardware, as well as the control Hamiltonian Hc(t)H_{c}(t) due to the application of the control pulse. In the second term, Lindblad operators LjL_{j} are used to model dissipative processes due to couplings with the environment. The above Lindblad master equation is solved numerically using the built-in function mesolve in QuTIP [45, 46].

To determine our hardware-specific Hamiltonians and Lindbladians, the transmon qudit is modeled using the Cooper pair box model whose parameters are measured experimentally [30]. The measurements are made for the lowest three levels of the qudit and then extrapolated to higher levels. First, the transition frequencies and the effective drive Hamiltonian are measured using Rabi spectroscopy. In the lab frame and in the transmon eigenbasis, keeping the lowest five levels, we approximate

H0(00000025.7580000050.0990000072.8480000093.828),c(01.000000001.372000001.618000001.78100000),H_{0}\simeq\left(\begin{array}[]{ccccc}0&0&0&0&0\\ 0&25.758&0&0&0\\ 0&0&50.099&0&0\\ 0&0&0&72.848&0\\ 0&0&0&0&93.828\end{array}\right),\quad c\simeq\left(\begin{array}[]{ccccc}0&1.000&0&0&0\\ 0&0&-1.372&0&0\\ 0&0&0&-1.618&0\\ 0&0&0&0&1.781\\ 0&0&0&0&0\end{array}\right), (S.3)

where angular frequencies are in units of rad/ns. Second, two Lindbladians are included to model decay and dephasing. The Lindblad operator L1aL_{1}\sim a describes successive decays to lower levels, whose nonzero matrix elements are L1(j,j+1)=1/T1(j+1,j)L_{1}(j,j+1)=1/\sqrt{T_{1}(j+1,j)}. The Lindblad operator L2aaL_{2}\sim a^{\dagger}a describes dephasing with respect to lower levels, whose nonzero matrix elements are L2(j,j)=1/T2(j,j1)L_{2}(j,j)=1/\sqrt{T_{2}^{*}(j,j-1)}. The T1T_{1} decay time is measured by readout delays, and the T2T_{2}^{*} dephasing time is measured using Ramsey spectroscopy. Keeping the lowest five levels, we approximate

L1(00.004000000.006000000.007000000.00900000),L2(0000000.005000000.014000000.045000000.000).L_{1}\simeq\left(\begin{array}[]{ccccc}0&0.004&0&0&0\\ 0&0&0.006&0&0\\ 0&0&0&0.007&0\\ 0&0&0&0&0.009\\ 0&0&0&0&0\end{array}\right),\quad L_{2}\simeq\left(\begin{array}[]{ccccc}0&0&0&0&0\\ 0&0.005&0&0&0\\ 0&0&0.014&0&0\\ 0&0&0&0.045&0\\ 0&0&0&0&0.000\end{array}\right). (S.4)

Parameters in the Hamiltonians and the Lindbladians may drift over time and change after each cool down. The above parameters were obtained from the Quantum Design and Integration Testbed (QuDIT) calibration that was performed immediately before the experimental runs that produced the data reported in the main text.