Digitized Counterdiabatic Quantum Algorithms for Logistics Scheduling

Archismita Dalal [email protected] Kipu Quantum, Greifswalderstrasse 226, 10405 Berlin, Germany    Iraitz Montalban Kipu Quantum, Greifswalderstrasse 226, 10405 Berlin, Germany    Narendra N. Hegade Kipu Quantum, Greifswalderstrasse 226, 10405 Berlin, Germany    Alejandro Gomez Cadavid Kipu Quantum, Greifswalderstrasse 226, 10405 Berlin, Germany Department of Physics, University of the Basque Country UPV/EHU, Barrio Sarriena, s/n, 48940 Leioa, Biscay, Spain    Enrique Solano Kipu Quantum, Greifswalderstrasse 226, 10405 Berlin, Germany    Abhishek Awasthi [email protected] BASF Digital Solutions GmbH, Pfalzgrafenstrasse 1, 67061 Ludwigshafen am Rhein, Germany    Davide Vodola BASF Digital Solutions GmbH, Pfalzgrafenstrasse 1, 67061 Ludwigshafen am Rhein, Germany    Caitlin Jones BASF Digital Solutions GmbH, Pfalzgrafenstrasse 1, 67061 Ludwigshafen am Rhein, Germany    Horst Weiss BASF SE, Pfalzgrafenstrasse 1, 67061 Ludwigshafen am Rhein, Germany    Gernot Füchsel [email protected] Kipu Quantum, Greifswalderstrasse 226, 10405 Berlin, Germany
Abstract

We study a job shop scheduling problem for an automatized robot in a high-throughput laboratory and a travelling salesperson problem with recently proposed digitized counterdiabatic quantum optimization (DCQO) algorithms. In DCQO, we find the solution of an optimization problem via an adiabatic quantum dynamics, which is accelerated with counterdiabatic protocols. Thereafter, we digitize the global unitary to encode it in a digital quantum computer. For the job-shop scheduling problem, we aim at finding the optimal schedule for a robot executing a number of tasks under specific constraints, such that the total execution time of the process is minimized. For the traveling salesperson problem, the goal is to find the path that covers all cities and is associated with the shortest traveling distance. We consider both hybrid and pure versions of DCQO algorithms and benchmark the performance against digitized quantum annealing and the quantum approximate optimization algorithm (QAOA). In comparison to QAOA, the DCQO solution is improved by several orders of magnitude in success probability using the same number of two-qubit gates. Moreover, we experimentally implement our algorithms on superconducting and trapped-ion quantum processors. Our results demonstrate that circuit compression using counterdiabatic protocols is amenable to current NISQ hardware and can solve logistics scheduling problems, where other digital quantum algorithms show insufficient performance.

I Introduction

Quantum computing holds the promise of providing a significant advantage over classical computers for various industry-relevant problems, ranging from material simulation to optimization and machine learning [1]. In particular, the NP-hard class of optimization problems finds applications in finance, planning, and logistics [2]. Despite the theoretical speedup claims of quantum search algorithms [3], no quantum algorithm has yet demonstrated an advantage over classical optimization algorithms using noisy intermediate-scale quantum (NISQ) computers. On the other hand, claims of “quantum supremacy” and “quantum utility” have been established using NISQ devices for non-industry-relevant problems [4, 5, 6]. This gap in performance is mainly because the required depth of the quantum circuits for achieving an optimal solution exceeds the coherence time of a NISQ processor, and the imperfect quantum gate operations further mars the quality of the solution. To counteract this drawback, one not only needs high-performing processors but also seeks an efficient problem-to-circuit encoding. In this regard, we overcome the later challenge by using our circuit compression techniques. Specifically, we apply digitized counterdiabatic protocols to reduce the required circuit depth for tackling a combinatorial optimization problem on a gate-based quantum computer [7, 8], thereby demonstrating superior performance for various instances from industrial use cases.

The develo** field of digitized counterdiabatic quantum computing finds applications in various problems ranging from combinatorial optimization, many-body ground state preparation to protein folding [7, 8, 9, 10, 11, 12, 13, 14]. This paradigm combines the advantages of counterdiabatic (CD) protocols with the flexibility of digital quantum processors. CD protocols were first introduced in the analog quantum computing domain, where a CD Hamiltonian is added to the system’s Hamiltonian to speed up the evolution without making any undesired transitions between the instantaneous eigenstates [15, 16, 17]. In the digital domain, CD techniques have been demonstrated to enhance the performance of digitized quantum annealing and quantum approximate optimization algorithm [8, 11]. Unlike their analog counterparts, digital quantum computers grant the ability to realize arbitrary CD terms. This flexibility paves the way for diverse Hamiltonians and superior control. In this work we focus on digitized counterdiabatic quantum optimization (DCQO) [8] to solve logistics scheduling use cases. By modifying the annealing Hamiltonian with additional CD terms, rapid state evolution can be achieved without succumbing to non-adiabatic transitions. This not only reduces the required circuit depth but also minimizes the impact of gate errors and decoherence. Moreover, the hybrid version, namely hybrid-DCQO (h-DCQO), integrates classical optimization techniques with DCQO to further enhance the solution quality and robustness [13].

In this work, we tackle two specific use cases from the NP-hard optimization problem class, where quantum computing is expected to offer a promising avenue for finding optimal or near-optimal solutions faster than their classical counterpart. The first use case is an extension of a typical job-shop scheduling problem (JSSP) to an industrial high-throughput laboratory process, and the second is the travelling salesperson problem (TSP). For the JSSP use case considered in this work, the task is to determine the most efficient sequence of operations to be carried out by a robot, such that the different chemical samples are processed at different machines [18]. The complexity of these problems increases with the number of samples, machines used, and operational constraints, and a solution from a concrete classical algorithm requires an unfeasible growth in classical compute operations with problem size. Due to the large problem sizes of even the simplest non-trivial JSSP instances considered here and the size limitations of gate-based current NISQ devices, we restrict our study to a few ‘subproblems’ of smaller sizes. For TSP, where the task is to visit a specified number of cities in the least amount of time and return to the starting city, the required solution time also increases almost exponentially with the problem size, i.e. the number of cities involved [19]. Due to its more simplistic formulation, we are currently able to solve three-city and four-city TSP instances on gate-based NISQ hardware.

We employ DCQO and h-DCQO algorithms to obtain high-quality solutions for JSSP and TSP instances. Performing ideal noiseless simulations, we demonstrate superior performance of DCQO and h-DCQO over digitized quantum annealing (DQA) and quantum approximate optimization algorithm (QAOA), respectively, for all instances of JSSP and TSP studied in this work. Notably, the DCQO circuit yields the exact solution 3×\times× faster than the optimal QAOA circuit for one JSSP instance tested. Considering the variational training of QAOA and its multiple random initializations, DCQO is actually many orders of magnitude faster because of its pure quantum approach. We further decrease the depth of the DCQO circuits by restricting the evolution to the impulse regime, i.e. short evolution times, and discarding two-qubit gates with low-magnitude angles [13]. By adapting our original proposal for h-DCQO [13], here we construct different variants of the ansatz to serve the problem instance better. These techniques, along with efficient representation of the CD Hamiltonian in terms of hardware-dependent two-qubit gates, are essential towards successful implementations on NISQ hardware. We experimentally validate the DCQO and h-DCQO circuits using one 16-qubit JSSP instance, and 9-qubit and 16-qubit TSP instances. IonQ’s trapped-ion hardware successfully solves all three instances, whereas IBMQ’s superconducting circuit could only solve the 9-qubit instance. Thus, our results demonstrate that counterdiabatic techniques are better able to utilize NISQ hardware for small-scale optimization problems.

The work is organized as follows. We provide a brief background on the two use cases that we tacked in this work, namely, JSSP and TSP, in §II. Next in §III, we explain digitized counterdiabatic optimization algorithms, both pure quantum and hybrid quantum-classical versions, and the circuit-depth reduction techniques that we use for successfully solving optimization problems on NISQ hardware. The solutions obtained from using (h-)DCQO for JSSP and TSP are discussed in §IV A and B, respectively, along with their comparison against the solutions obtained from existing NISQ optimization algorithms. Finally, we conclude with a future scope and outlook in §V.

II Problem descriptions

In this section, we describe the two use cases in logistics scheduling that we tackle in this work. First we discuss the scheduling problem in a high-throughput laboratory process, and the challenges associated with solving its relevant instances on NISQ hardware. Next we provide a brief background on TSP and its existing quantum solutions.

II.1 Job-shop scheduling problem

Refer to caption
Figure 1: A simplified job-shop scheduling problem. The robot R carries a chemical sample between the four stations (rack, mixer, shaker, photo) for processing in a pre-defined order, i.e. rack \rightarrow mixer \rightarrow shaker \rightarrow photo booth. The path is shown by solid lines with decreasing thickness from the first to the last segment. Moreover, the samples can be transported back to the rack from mixer and shaker for storing in between the processing, which is shown by dashed lines.

The JSSP describes here an automatized process in a chemical laboratory and is schematically represented in Fig. 1. The goal is to find an optimal sequence of autonomous operations that minimizes the total time required to process a given number of chemical samples (nssubscript𝑛sn_{\text{s}}italic_n start_POSTSUBSCRIPT s end_POSTSUBSCRIPT) in a laboratory [18]. The setup here consists of four stations, namely rack, mixer, shaker and photo booth, and a single robot is tasked with moving samples between the stations one at a time, see Fig. 1. Each sample needs to be processed in a particular order: mixed, shaken and then photographed multiple times in pre-defined intervals. The rack stores all the samples in the beginning and at the end of its processing, as well as any time in between if needed. There are other constraints including pre-determined processing times at mixer and shaker, and required time-gap between two successive photographs for each sample. Thus, the goal is to find a sequence of instructions for the robot that minimizes the sum of processing times of each sample under multiple constraints.

In the simplest non-trivial setting, one can consider two samples, i.e. ns=2subscript𝑛s2n_{\text{s}}=2italic_n start_POSTSUBSCRIPT s end_POSTSUBSCRIPT = 2, one photograph for each sample and a time horizon (ntsubscript𝑛tn_{\text{t}}italic_n start_POSTSUBSCRIPT t end_POSTSUBSCRIPT) of 30 units. The total number of machines (nmsubscript𝑛mn_{\text{m}}italic_n start_POSTSUBSCRIPT m end_POSTSUBSCRIPT) used is 3, which are the rack, the shaker and one photo station. Based on a particular problem formulation [18], the number of binary variables required is 2nmnsnt2subscript𝑛msubscript𝑛ssubscript𝑛t2n_{\text{m}}n_{\text{s}}n_{\text{t}}2 italic_n start_POSTSUBSCRIPT m end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT s end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT t end_POSTSUBSCRIPT. Thus, the simplest problem setting results in a quadratic unconstrained binary optimization (QUBO) problem instance of size 360. Many of these variables can be eliminated by imposing the problem constraints. But a quantum optimization algorithm would still require hundreds of qubits to solve a non-trivial JSSP instance, which is beyond what current NISQ devices are capable of and an adaption of the problem size needs to be carried out.

One of the most important tasks of this work is to test the DCQO and h-DCQO algorithm on simulators as well as real hardware. Therefore, in this work we take only parts of the QUBO matrix and build sub-QUBOs of size 16161616, which are implementable on current NISQ devices. These sub-QUBOs are selected in a manner that the connectivity between the pairs of variables is very high, which requires a well connected hardware to embed the problem Hamiltonian. After solving all these subproblems, the solution of the full QUBO problem is updated and the decomposition-composition process is repeated until some convergence criteria is met. This iterative technique is commonly known as the large neighborhood search in classical optimization [20]. In the quantum domain, D-Wave has long borrowed this idea to create their quantum-classical Hybrid Solver [21, 22], and more recently QAOA has been implemented in a similar setting [23]. In this work, we focus on solving ten of the subproblems from a JSSP instance on gate-based quantum devices, and analyze the performance of the studied quantum algorithms and the hardware, in achieving good results for densely connected QUBOs. We present simulation results for all these ten 16-qubit instances to compare DCQO and DQA. For one of the instances, we provide experimental results from both IonQ and IBMQ hardware. Additionally, we compare h-DCQO against QAOA for the ten instances.

II.2 Travelling salesperson problem

Refer to caption
Figure 2: A TSP instance with 10 cities A to J. The shortest round trip is shown using solid (gold) thick lines. An alternative round trip with greater distance is shown with dashed (black) thin lines.

TSP is a prototypical combinatorial optimization problem falling into the complexity class of NP-hard [24]. The task is to find the shortest round trip that a salesperson can take to cover a given number of cities, see Fig. 2. Besides planning and logistics, TSP finds applications in manufacturing and DNA sequencing [25, 26]. The scaling of classical-compute time is almost exponential with the number of cities, making TSP an ideal candidate for quantum computing. Although Grover’s quantum algorithm promises a quadratic speedup for search [3], this claim is far beyond reach due to noise and limited size of current quantum hardware. Thus we solve TSP instances requiring up to 16 qubits in this work.

We follow the one-hot encoding technique to map TSP into a QUBO problem [19]. As per this encoding, n2superscript𝑛2n^{2}italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bits are required to represent a n𝑛nitalic_n-city TSP. In a bitstring 𝐱𝐱\bf{x}bold_x of length n2superscript𝑛2n^{2}italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, xt,c=1subscript𝑥𝑡𝑐1x_{t,c}=1italic_x start_POSTSUBSCRIPT italic_t , italic_c end_POSTSUBSCRIPT = 1 if the salesperson visits city c𝑐citalic_c in the t𝑡titalic_t step of their path. For example, for n=4𝑛4n=4italic_n = 4 cities, a path 021302130\to 2\to 1\to 30 → 2 → 1 → 3 is represented by 1000 0010 0100 000110000010010000011000\ 0010\ 0100\ 00011000 0010 0100 0001. With a trivial map** of one bit to one qubit, we thus end up using Nq=n2subscript𝑁qsuperscript𝑛2N_{\text{q}}=n^{2}italic_N start_POSTSUBSCRIPT q end_POSTSUBSCRIPT = italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT qubits. Due to such a big overhead in the number of qubits, and consequently possible outcomes, most (2Nqn!2superscript2subscript𝑁q𝑛22^{N_{\text{q}}}-\frac{n!}{2}2 start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT q end_POSTSUBSCRIPT end_POSTSUPERSCRIPT - divide start_ARG italic_n ! end_ARG start_ARG 2 end_ARG) bitstrings are infeasible outcomes for TSP. This makes finding the shortest path even more challenging. Utilizing quantum algorithms for solving simplistic 3-city and 4-city TSPs not only helps us verifying whether constraints and objectives can be converged with these algorithms, but also makes implementation of TSP on NISQ hardware feasible.

Only a few quantum approaches have been mentioned in literature to solve TSP instances. In the analog quantum computing regime, D-Wave’s quantum annealing provides a straightforward approach, but can only solve up to 8-city TSP instances due to device noise and sparse qubit connectivity [27]. In the digital regime, quantum phase estimation [28] and QAOA [29] have been used to address these problems. A recent numerical study demonstrates the advantage of using a superior encoding, i.e. nlogn𝑛𝑛n\log nitalic_n roman_log italic_n instead of n2superscript𝑛2n^{2}italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, for solving 4-city TSP instances using noiseless simulations of QAOA [30]. Additionally, using different variants of QAOA circuit has shown superior performance in noiseless simulation of TSP instances with up to 16 qubits [31].

III Methods

A real-variable quadratic optimization problem can be approximately converted into a binary-variable quadratic optimization problem using real-to-binary encoding. Furthermore, any constraint can be added to the optimization cost function by appropriate Lagrange multiplier. This leads to a QUBO problem, where the goal is to obtain the bit-string x0subscript𝑥0x_{0}italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT that minimizes or maximizes the objective function [32]

f(x)=xTQx,𝑓𝑥superscript𝑥𝑇𝑄𝑥f(x)=x^{T}Qx,italic_f ( italic_x ) = italic_x start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_Q italic_x , (1)

for a problem-dependent real-valued symmetric matrix Q𝑄Qitalic_Q. This QUBO problem is equivalent to finding the ground state of an Ising spin-glass Hamiltonian [33]

Hp=i<jJijσziσzj+ihiσzi,subscript𝐻psubscript𝑖𝑗subscript𝐽𝑖𝑗superscriptsubscript𝜎𝑧𝑖superscriptsubscript𝜎𝑧𝑗subscript𝑖subscript𝑖superscriptsubscript𝜎𝑧𝑖H_{\text{p}}=\sum_{i<j}J_{ij}\sigma_{z}^{i}\sigma_{z}^{j}+\sum_{i}h_{i}\sigma_% {z}^{i},italic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i < italic_j end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT , (2)

where σzisuperscriptsubscript𝜎𝑧𝑖\sigma_{z}^{i}italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT denotes the Pauli Z matrix acting on the i𝑖iitalic_i-th spin, and the coefficients Jij=12Qijsubscript𝐽𝑖𝑗12subscript𝑄𝑖𝑗J_{ij}=\frac{1}{2}Q_{ij}italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_Q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT and hi=12Qiij12Qijsubscript𝑖12subscript𝑄𝑖𝑖subscript𝑗12subscript𝑄𝑖𝑗h_{i}=-\frac{1}{2}Q_{ii}-\sum_{j}\frac{1}{2}Q_{ij}italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_Q start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_Q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT. Problems encoded in Ising Hamiltonians are suitable for a quantum computing treatment and several quantum optimization algorithms have been developed including quantum annealing [34], variational quantum eigensolver [35] and quantum approximate optimization algorithm [36], which are briefly discussed in Appendix A.

In this section, we elaborate on the design and analysis of counterdiabatic algorithms, both the pure quantum implementation and the quantum-classical variational implementation, for solving combinatorial optimization problems. Additionally, we discuss methods for further reducing circuit depths to make DCQO circuits amenable to NISQ hardware. In particular, we discuss the concepts of impulse regime and gate-angle thresholding for DCQO circuit construction, and explain how the Pauli operations in the CD Hamiltonian can be efficiently represented on a given quantum hardware, specifically, on IBMQ and IonQ devices.

III.1 Digitized counterdiabatic quantum optimization

A rapid driving of the adiabatic Hamiltonian Had(t)subscript𝐻ad𝑡H_{\text{ad}}(t)italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT ( italic_t ) (cf. Eq. (12)) results in unwanted transitions between instantaneous eigenstates and, consequently, a sub-optimal solution of the optimization problem. In this regard, counterdiabatic driving is effected by introducing a Hamiltonian Hcd(t)subscript𝐻cd𝑡H_{\text{cd}}(t)italic_H start_POSTSUBSCRIPT cd end_POSTSUBSCRIPT ( italic_t ) as an extension to the adiabatic Hamiltonian by

H(t)=Had(t)+λ˙(t)Aλ(t).𝐻𝑡subscript𝐻ad𝑡˙𝜆𝑡subscript𝐴𝜆𝑡H(t)=H_{\text{ad}}(t)+\dot{\lambda}(t)A_{\lambda}(t).italic_H ( italic_t ) = italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT ( italic_t ) + over˙ start_ARG italic_λ end_ARG ( italic_t ) italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) . (3)

The adiabatic gauge potential Aλ(t)subscript𝐴𝜆𝑡A_{\lambda}(t)italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) is responsible for suppressing the non-adiabatic state-to-state transitions during the Hamiltonian evolution. Here, λ(t)𝜆𝑡\lambda(t)italic_λ ( italic_t ) is a desired schedule function, and λ˙(t)˙𝜆𝑡\dot{\lambda}(t)over˙ start_ARG italic_λ end_ARG ( italic_t ) its time-derivative. Using Eq. (3), one can formulate the evolution operator as

U(tf,t0)=τeit0tfH(t)𝑑t,𝑈subscript𝑡𝑓subscript𝑡0𝜏superscript𝑒𝑖superscriptsubscriptsubscript𝑡0subscript𝑡𝑓𝐻𝑡differential-d𝑡U(t_{f},t_{0})=\tau e^{-i\int_{t_{0}}^{t_{f}}H(t)dt},italic_U ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = italic_τ italic_e start_POSTSUPERSCRIPT - italic_i ∫ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_H ( italic_t ) italic_d italic_t end_POSTSUPERSCRIPT , (4)

with τ𝜏\tauitalic_τ being the time-ordering operator, t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT are the initial and final evolution time, respectively. For the short-time limit, where limλ˙(t)subscript˙𝜆𝑡\lim_{\dot{\lambda}(t)\to\infty}roman_lim start_POSTSUBSCRIPT over˙ start_ARG italic_λ end_ARG ( italic_t ) → ∞ end_POSTSUBSCRIPT, one can show the time evolution of H(t)𝐻𝑡H(t)italic_H ( italic_t ) is governed solely by the gauge potential so that Eq.(3) simplifies to H(t)=λ˙(t)Aλ(t)𝐻𝑡˙𝜆𝑡subscript𝐴𝜆𝑡H(t)=\dot{\lambda}(t)A_{\lambda}(t)italic_H ( italic_t ) = over˙ start_ARG italic_λ end_ARG ( italic_t ) italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ), see ref.[37]. In the following, we will always work in the impulse regime, that is in the short time limit, as this leads to shorter circuits in the digital implementation.

The exact computation of Aλ(t)subscript𝐴𝜆𝑡A_{\lambda}(t)italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) is computationally expensive because it requires the knowledge of the instantaneous eigenspectrum of the adiabatic Hamiltonian, Had(t)subscript𝐻ad𝑡H_{\text{ad}}(t)italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT ( italic_t ), at any time t𝑡titalic_t [16]. Therefore, approximate techniques for constructing and realizing counterdiabaticity have been proposed [38, 39, 40, 41]. In particular, one can approximate the gauge potential using a series of nested commutators (NC) as

Aλ(l)(t)=ik=1lαk(t)[Had,[Had,[Had,2k1λHad]]],A_{\lambda}^{(l)}(t)=i\sum_{k=1}^{l}\alpha_{k}(t)\underbrace{[H_{\text{ad}},[H% _{\text{ad}},\dots[H_{\text{ad}},}_{2k-1}\partial_{\lambda}H_{\text{ad}}]]]\;,italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT ( italic_t ) = italic_i ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) under⏟ start_ARG [ italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT , [ italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT , … [ italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT , end_ARG start_POSTSUBSCRIPT 2 italic_k - 1 end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT ] ] ] , (5)

where l𝑙litalic_l sets the maximum order of approximation and αk(t)subscript𝛼𝑘𝑡\alpha_{k}(t)italic_α start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) is the CD coefficient corresponding to the k𝑘kitalic_kth-order approximation.

To implement counterdiabatic driving on a gate-based quantum computer, we discretize the total evolution time T𝑇Titalic_T of the full Hamiltonian Eq. (3) into N𝑁Nitalic_N Trotter steps, with each step being applied for a duration of Δt=TNΔ𝑡𝑇𝑁\Delta t=\frac{T}{N}roman_Δ italic_t = divide start_ARG italic_T end_ARG start_ARG italic_N end_ARG. By representing H(t)𝐻𝑡H(t)italic_H ( italic_t ) as linear combination of L𝐿Litalic_L k𝑘kitalic_k-local Pauli operators Hksubscript𝐻𝑘H_{k}italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, expressed as H(t)=k=1Lck(t)Hk𝐻𝑡superscriptsubscript𝑘1𝐿subscript𝑐𝑘𝑡subscript𝐻𝑘H(t)=\sum_{k=1}^{L}c_{k}(t)H_{k}italic_H ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, and employing the first-order Trotterization, the continuous-time dynamics can be approximated as

Udigital m=1Nk=1Lexp[iΔtck(mΔt)Hk].subscript𝑈digital superscriptsubscriptproduct𝑚1𝑁superscriptsubscriptproduct𝑘1𝐿𝑖Δ𝑡subscript𝑐𝑘𝑚Δ𝑡subscript𝐻𝑘U_{\text{digital }}\approx\prod_{m=1}^{N}\prod_{k=1}^{L}\exp\left[-i\,\Delta t% \,c_{k}(m\Delta t)\,H_{k}\right].italic_U start_POSTSUBSCRIPT digital end_POSTSUBSCRIPT ≈ ∏ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ∏ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT roman_exp [ - italic_i roman_Δ italic_t italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_m roman_Δ italic_t ) italic_H start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ] . (6)

As each matrix-exponential term in the above expression can be easily decomposed into a set of quantum gates, this technique enables realization of arbitrary counterdiabatic terms on a gate-based quantum computer. This further motivates the choice of DCQO over quantum annealing.

The implementation of DCQO on near-term quantum processors suffers from various constraints and challenges. The accuracy of any digitized (Trotterized) algorithm depends on the step-size parameter ΔtΔ𝑡\Delta troman_Δ italic_t, with smaller ΔtΔ𝑡\Delta troman_Δ italic_t and larger N𝑁Nitalic_N typically resulting in higher success probability. The accuracy and robustness of digitized adiabatic and digitized counterdiabatic algorithms have been previously studied [42]. Moreover, a higher order NC expansion, Eq.(5), creates many-body Pauli terms, whose implementation with only one- and two-qubit gates dramatically increases the circuit depth. Thus, in practice, we use only the first-order NC and a few Trotter steps to keep the DCQO circuit depth within the coherence limit of current quantum processors. This approach to Aλsubscript𝐴𝜆A_{\lambda}italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT does not guarantee full suppression of non-adiabatic transitions during the evolution, but reduces their probability to an extend that allows achieving good results on current NISQ devices, as will be shown below in the experimental sections.

One can analytically calculate the CD Hamiltonian Eq. (3) for the first order NC expansion [8]. In this approximation, the CD coefficient reads

α1(t)=14ihi2+i<jJij2γ(t),subscript𝛼1𝑡14subscript𝑖superscriptsubscript𝑖2subscript𝑖𝑗superscriptsubscript𝐽𝑖𝑗2𝛾𝑡\displaystyle\alpha_{1}(t)=-\frac{1}{4}\frac{\sum_{i}h_{i}^{2}+\sum_{i<j}J_{ij% }^{2}}{\gamma(t)}\,,italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) = - divide start_ARG 1 end_ARG start_ARG 4 end_ARG divide start_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_i < italic_j end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_γ ( italic_t ) end_ARG , (7)

where

γ(t)=(1λ(t))2(ihi2+4ijJij2)+λ2(t)[ihi4+ijJij4+6ijhi2Jij2+6i<j<k(Jij2Jik2+Jij2Jjk2+Jik2Jjk2)],𝛾𝑡superscript1𝜆𝑡2subscript𝑖superscriptsubscript𝑖24subscript𝑖𝑗superscriptsubscript𝐽𝑖𝑗2superscript𝜆2𝑡delimited-[]subscript𝑖superscriptsubscript𝑖4subscript𝑖𝑗superscriptsubscript𝐽𝑖𝑗46subscript𝑖𝑗superscriptsubscript𝑖2superscriptsubscript𝐽𝑖𝑗26subscript𝑖𝑗𝑘superscriptsubscript𝐽𝑖𝑗2superscriptsubscript𝐽𝑖𝑘2superscriptsubscript𝐽𝑖𝑗2superscriptsubscript𝐽𝑗𝑘2superscriptsubscript𝐽𝑖𝑘2superscriptsubscript𝐽𝑗𝑘2\begin{split}\gamma(t)=&(1-\lambda(t))^{2}\left(\sum_{i}h_{i}^{2}+4\sum_{i\neq j% }J_{ij}^{2}\right)\\ &+\lambda^{2}(t)\bigg{[}\sum_{i}h_{i}^{4}+\sum_{i\neq j}J_{ij}^{4}+6\sum_{i% \neq j}h_{i}^{2}J_{ij}^{2}\\ &+6\sum_{i<j<k}\left(J_{ij}^{2}J_{ik}^{2}+J_{ij}^{2}J_{jk}^{2}+J_{ik}^{2}J_{jk% }^{2}\right)\bigg{]},\end{split}start_ROW start_CELL italic_γ ( italic_t ) = end_CELL start_CELL ( 1 - italic_λ ( italic_t ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 ∑ start_POSTSUBSCRIPT italic_i ≠ italic_j end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) [ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_i ≠ italic_j end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + 6 ∑ start_POSTSUBSCRIPT italic_i ≠ italic_j end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + 6 ∑ start_POSTSUBSCRIPT italic_i < italic_j < italic_k end_POSTSUBSCRIPT ( italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_J start_POSTSUBSCRIPT italic_i italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_J start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ] , end_CELL end_ROW (8)

and the approximated gauge potential is represented as

Aλ(1)(t)=2α1(t)[ihiσyi+i<jJij(σyiσzj+σziσyj)].superscriptsubscript𝐴𝜆1𝑡2subscript𝛼1𝑡delimited-[]subscript𝑖subscript𝑖superscriptsubscript𝜎𝑦𝑖subscript𝑖𝑗subscript𝐽𝑖𝑗superscriptsubscript𝜎𝑦𝑖superscriptsubscript𝜎𝑧𝑗superscriptsubscript𝜎𝑧𝑖superscriptsubscript𝜎𝑦𝑗A_{\lambda}^{(1)}(t)=-2\alpha_{1}(t)\left[\sum_{i}h_{i}\sigma_{y}^{i}+\sum_{i<% j}J_{ij}\left(\sigma_{y}^{i}\sigma_{z}^{j}+\sigma_{z}^{i}\sigma_{y}^{j}\right)% \right].italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_t ) = - 2 italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) [ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_i < italic_j end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ) ] . (9)

Using Eq.(9) and a sinusoidal scheduling function λ(t)=sin2(π2sin2(πt2T))𝜆𝑡superscript2𝜋2superscript2𝜋𝑡2𝑇\lambda(t)=\sin^{2}\left(\frac{\pi}{2}\sin^{2}\left(\frac{\pi t}{2T}\right)\right)italic_λ ( italic_t ) = roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_π end_ARG start_ARG 2 end_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_π italic_t end_ARG start_ARG 2 italic_T end_ARG ) ), the digitized evolution of the CD Hamiltonian is given by the unitary operator

Ucdsubscript𝑈cd\displaystyle U_{\text{cd}}italic_U start_POSTSUBSCRIPT cd end_POSTSUBSCRIPT \displaystyle\approx m=1Nexp(i2πsin(πmΔtT)sin(πsin2(πmΔt2T))4TΔtAλ(1)(mΔt))superscriptsubscriptproduct𝑚1𝑁𝑖2𝜋𝜋𝑚Δ𝑡𝑇𝜋superscript2𝜋𝑚Δ𝑡2𝑇4𝑇Δ𝑡superscriptsubscript𝐴𝜆1𝑚Δ𝑡\displaystyle\prod_{m=1}^{N}\exp\left(-i\frac{2\pi\sin(\pi m\frac{\Delta t}{T}% )\sin(\pi\sin^{2}(\pi m\frac{\Delta t}{2T}))}{4T}\Delta tA_{\lambda}^{(1)}(m% \Delta t)\right)∏ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_exp ( - italic_i divide start_ARG 2 italic_π roman_sin ( italic_π italic_m divide start_ARG roman_Δ italic_t end_ARG start_ARG italic_T end_ARG ) roman_sin ( italic_π roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_π italic_m divide start_ARG roman_Δ italic_t end_ARG start_ARG 2 italic_T end_ARG ) ) end_ARG start_ARG 4 italic_T end_ARG roman_Δ italic_t italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_m roman_Δ italic_t ) ) (10)
=\displaystyle== m=1Nexp(iπsin(πmN)sin(πsin2(πm2N))2NAλ(1)(mΔt)).superscriptsubscriptproduct𝑚1𝑁𝑖𝜋𝜋𝑚𝑁𝜋superscript2𝜋𝑚2𝑁2𝑁superscriptsubscript𝐴𝜆1𝑚Δ𝑡\displaystyle\prod_{m=1}^{N}\exp\left(-i\frac{\pi\sin(\pi\frac{m}{N})\sin(\pi% \sin^{2}(\pi\frac{m}{2N}))}{2N}A_{\lambda}^{(1)}(m\Delta t)\right).∏ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_exp ( - italic_i divide start_ARG italic_π roman_sin ( italic_π divide start_ARG italic_m end_ARG start_ARG italic_N end_ARG ) roman_sin ( italic_π roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_π divide start_ARG italic_m end_ARG start_ARG 2 italic_N end_ARG ) ) end_ARG start_ARG 2 italic_N end_ARG italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_m roman_Δ italic_t ) ) .

The above equation shows that the evolution of Hcdsubscript𝐻cdH_{\text{cd}}italic_H start_POSTSUBSCRIPT cd end_POSTSUBSCRIPT remains independent of T𝑇Titalic_T if N𝑁Nitalic_N is fixed; this is independent from the expansion order of the gauge potential, Eq.(5). For an ideal simulation of DCQO, as given by Ucdsubscript𝑈cdU_{\text{cd}}italic_U start_POSTSUBSCRIPT cd end_POSTSUBSCRIPT, the success probability depends on N𝑁Nitalic_N, T𝑇Titalic_T (or ΔtΔ𝑡\Delta troman_Δ italic_t) and the number of qubits Nqsubscript𝑁qN_{\text{q}}italic_N start_POSTSUBSCRIPT q end_POSTSUBSCRIPT. As previously demonstrated, the success probability drops exponentially with increasing Nqsubscript𝑁qN_{\text{q}}italic_N start_POSTSUBSCRIPT q end_POSTSUBSCRIPT, but with a polynomial speedup as compared to digitized quantum annealing [8].

III.2 Hybrid-DCQO

Refer to caption
Figure 3: A p𝑝pitalic_p-layer hybrid-DCQO circuit assigns two parameters per layer l𝑙litalic_l, namely αlsubscript𝛼𝑙\alpha_{l}italic_α start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and βlsubscript𝛽𝑙\beta_{l}italic_β start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT for one-body and two-body terms in the CD Hamiltonian, respectively. The light orange block represents the single-qubit evolution exp(ihiαlσyi)𝑖subscript𝑖subscript𝛼𝑙superscriptsubscript𝜎𝑦𝑖\exp(-ih_{i}\alpha_{l}\sigma_{y}^{i})roman_exp ( - italic_i italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ). The light yellow block represents the combined evolution exp(iβlJijσziσyj)exp(iβlJijσyiσzj)𝑖subscript𝛽𝑙subscript𝐽𝑖𝑗subscriptsuperscript𝜎𝑖𝑧subscriptsuperscript𝜎𝑗𝑦𝑖subscript𝛽𝑙subscript𝐽𝑖𝑗subscriptsuperscript𝜎𝑖𝑦subscriptsuperscript𝜎𝑗𝑧\exp(-i\beta_{l}J_{ij}\sigma^{i}_{z}\sigma^{j}_{y})\exp(-i\beta_{l}J_{ij}% \sigma^{i}_{y}\sigma^{j}_{z})roman_exp ( - italic_i italic_β start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) roman_exp ( - italic_i italic_β start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) on two qubits i𝑖iitalic_i and j𝑗jitalic_j. The circuit is initialized by an uniform superposition of the computational basis. The parameters are updated in an iterative manner based on the classical optimization of the cost function α,β|Hp|α,βquantum-operator-product𝛼𝛽subscript𝐻p𝛼𝛽\langle{\mathbf{\alpha},\mathbf{\beta}}|H_{\text{p}}|{\mathbf{\alpha},\mathbf{% \beta}}\rangle⟨ italic_α , italic_β | italic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT | italic_α , italic_β ⟩.

The classical-quantum hybrid version of the DCQO algorithm is called hybrid-DCQO (h-DCQO) [9], which is a variational quantum algorithm based on a CD-inspired ansatz selection. The development of this algorithm is motivated by the observation that the success probability of the pure approach, i.e. DCQO, exponentially decreases with increasing number of qubits. In this algorithm, the CD coefficients αk(t)subscript𝛼𝑘𝑡\alpha_{k}(t)italic_α start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) from Eq. (5) are taken to be time-independent free parameters that can be variationally trained, as opposed to explicitly calculating them as done in the pure quantum approach. In an earlier version of the hybrid algorithm, the coefficients of the full Hamiltonian of Eq. (3) were parameterized, leading to digitized-counterdiabatic-QAOA (DC-QAOA) that numerically demonstrates superior performance over QAOA for random Ising spin models, MaxCut problem and p-spin models [11]. In comparison to DC-QAOA, the low-depth ansatz in the h-DCQO algorithm is composed of only the CD term, making it amenable to current quantum hardware. In particular, the h-DCQO algorithm has outperformed state-of-the-art algorithms for protein folding on both trapped ions and superconducting circuits [9], for portfolio optimization on trapped ions [13] and for p-spin model on superconducting circuits [14].

We now briefly describe the h-DCQO algorithm. As shown in Fig. 3, a simple h-DCQO ansatz is parameterized by α𝛼\mathbb{\alpha}italic_α and β𝛽\mathbb{\beta}italic_β, which correspond to the one- and two-body Hamiltonian terms, respectively, of the first order NC expansion, see Eq. (9). These parameters are then variationally updated to minimize the expectation value of the problem Hamiltonian from Eq. (2). Although a two-parameter h-DCQO ansatz ensures equal number of parameters for both QAOA and h-DCQO ansatze, the MS gate count for one layer of h-DCQO is twice than that of QAOA. In spite of this increased gate count per layer, our h-DCQO algorithm outperforms QAOA in portfolio optimization problems by reducing the required number of layers on trapped ions for achieving a target performance [13]. On the other hand, the CX gate counts for the QAOA and h-DCQO can be kept equal with efficient decomposition strategy, as will be shown below. In addition to the two-parameter h-DCQO ansatz, we also use an ansatz with (Nq+1)subscript𝑁q1(N_{\text{q}}+1)( italic_N start_POSTSUBSCRIPT q end_POSTSUBSCRIPT + 1 )-parameters in the current work, where Nqsubscript𝑁qN_{\text{q}}italic_N start_POSTSUBSCRIPT q end_POSTSUBSCRIPT one-body terms are independently parameterized.

III.3 Circuit-depth reduction techniques

Refer to caption

(a)
Refer to caption Refer to caption
         (b)                                                              (c)

Figure 4: Impulse regime analysis for a random 10-qubit Ising spin-glass model. (a) Hamiltonian coefficients vs. Trotter step for the three operational regimes. T=0.005,0.25,10𝑇0.0050.2510T=0.005,~{}0.25,~{}10italic_T = 0.005 , 0.25 , 10 for impulse, intermediate and adiabatic regimes, respectively, and N=20𝑁20N=20italic_N = 20. Notice the different y-axis scales of the three plots. (b) Dependence of success probability on evolution time T𝑇Titalic_T, with N=20𝑁20N=20italic_N = 20, for three different evolutions. The inset magnifies the impulse regime where the performance of only CD Hamiltonian matches that of the total Hamiltonian. (c) Dependence of CD coefficient on Trotter step, with the dashed line separating significant from insignificant contribution.

The evolution of the CD-enhanced Hamiltonian can be broadly categorized into three regimes, namely impulse, intermediate, and adiabatic regimes, based on the evolution time T𝑇Titalic_T [13]. In this work, we choose the impulse regime where the scheduling function changes more rapidly than the magnitude of the function itself, i.e. |λ(t)||λ˙(t)|much-less-than𝜆𝑡˙𝜆𝑡|\lambda(t)|\ll|\dot{\lambda}(t)|| italic_λ ( italic_t ) | ≪ | over˙ start_ARG italic_λ end_ARG ( italic_t ) |, for most of the evolution. This property makes the CD term Aλ(t)subscript𝐴𝜆𝑡A_{\lambda}(t)italic_A start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) more dominant than the adiabatic Hamiltonian Had(t)subscript𝐻ad𝑡H_{\text{ad}}(t)italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT ( italic_t ), and hence we can safely ignore the contribution of Had(t)subscript𝐻ad𝑡H_{\text{ad}}(t)italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT ( italic_t ) to the dynamics in the DCQO circuit. Further, the DCQO circuit can be constructed from N2𝑁2N-2italic_N - 2 Trotter steps in the impulse regime, because CD coefficients are zero for the initial step at t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and the final step at t0+Tsubscript𝑡0𝑇t_{0}+Titalic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_T.

In Fig. 4, we results from numerical DCQO simulations for a 10-qubit random Ising spin-glass model to demonstrate the applicability of different approximations. In Fig. 4(a), we plot the magnitudes of the CD Hamiltonian (violet color) and the adiabatic Hamiltonians (grey color) for different evolution times T𝑇Titalic_T ranging from 0.005 to 1 while kee** the number of Trotter steps fixed. In the impulse regime (T𝑇Titalic_T = 0.005), the maximum value of the CD Hamiltonian is roughly 40 times larger than that of the adiabatic Hamiltonian. With increasing T𝑇Titalic_T, this behavior reverses. In the intermediate regime (T𝑇Titalic_T = 0.25), the maxima of both Hamiltonians are similarly large, while in the adiabatic regime (T𝑇Titalic_T = 1) the adiabatic Hamiltonian becomes the dominant part of the total Hamiltonian Eq. (3). This is also evident from Fig. 4(b), where we we draw the success probability PSsubscript𝑃𝑆P_{S}italic_P start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT of finding the optimal solution for the three operational regimes for three different types of evolutions, namely i) evolution with adiabatic Hamiltonian only (in grey), ii) evolution with adiabatic and CD Hamiltonian (in gold), and evolution with only CD Hamiltonian. As expected, the “only CD"- approach yields a constant value over all T𝑇Titalic_T, for reasons explained above. In the impulse regime, the success probability is as large as the one obtained from the full Hamiltonian comprising of both, adiabatic and CD Hamiltian. The pure adiabatic approach yields PS0subscript𝑃𝑆0P_{S}\approx 0italic_P start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ≈ 0 as expected in the short-time limit. In the intermediate regime, the success probability increases for the adiabatic evolution and adiabatic + CD evolution, with the latter being significantly more effective in finding the optimal solution at lower times T𝑇Titalic_T. In the adiabatic regime, both perform similarly good. Finally, in Fig. 4(c) we draw the value of the first-order CD coefficient as function of the number of Trotter steps. Its value is zero for the initial and final Trotter step, which justifies its neglect in the DCQO circuit construction or even only the use of those dominant CD terms that are larger than a threshold value, here indicated as dashed line at 0.005. In this way, we can shorten the DCQO circuit to effectively half the required depth while thereby achieving similarly accurate results.

In addition to eliminating a few unitaries from the Trotterized CD evolution based on their corresponding Hamiltonian coefficients, we perform further circuit-depth reduction based on a pre-defined threshold on gate angles. If the angle associated with a single- or two-qubit rotation is smaller than a chosen threshold, we omit the corresponding gate from the DCQO circuit, resulting in a shorter-depth ‘DCQO (gc)’ circuit. This approximation is valid because the DCQO algorithm would then only apply the gates required to modify the state of the system up to a certain resolution set by the threshold. The smallest angle that IonQ hardware can accurately identify is 0.00628 [43], whereas for IBMQ, the value is not publicly available. Thus we choose a threshold of 0.1 in this work.

Further, we use a threshold based on the problem Hamiltonian Eq. (2) coefficients for constructing the h-DCQO ansatz. Given an Ising Hamiltonian whose coefficients hisubscript𝑖h_{i}italic_h start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and Jijsubscript𝐽𝑖𝑗J_{ij}italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT are normalized to lie between 11-1- 1 and 1111, we can discard some coupling terms whose Jijsubscript𝐽𝑖𝑗J_{ij}italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT are below certain threshold, without loosing much information about the Hamiltonian. This threshold value is problem dependent and we have used different values in this work. This extra step of two-qubit gate reduction is more relevant for IBMQ implementation because of the sparse connectivity of the hardware and the swap overhead.

We can further compress the depth of the DCQO circuit on an IBMQ hardware by expressing the two-qubit operations in terms of CX gates. Although the evolution of each two-body term, i.e. YZ𝑌𝑍YZitalic_Y italic_Z and ZY𝑍𝑌ZYitalic_Z italic_Y, requires two CX gates, we can efficiently express their combined evolution using only two instead of four CX gates. This decomposition can be easily implemented by first constructing a circuit where the rotation operators RYZsubscript𝑅𝑌𝑍R_{YZ}italic_R start_POSTSUBSCRIPT italic_Y italic_Z end_POSTSUBSCRIPT and RZYsubscript𝑅𝑍𝑌R_{ZY}italic_R start_POSTSUBSCRIPT italic_Z italic_Y end_POSTSUBSCRIPT are placed consecutively for each pair of interacting qubits, and then use a generic transpiler to find the efficient circuit. There are implications of reordering Pauli terms of a Hamiltonian in digitized quantum computation due to the error introduced from finite number of Trotter steps during the digitization process. The ordering of the unitary operations in the circuit representation determines the Trotter error, and, in principle, there exists an optimal ordering that minimizes the Trotter error for a given Hamiltonian [44]. The default ordering used by qiskit is lexicographical, where the two Pauli terms IYZ𝐼𝑌𝑍IYZitalic_I italic_Y italic_Z and IZY𝐼𝑍𝑌IZYitalic_I italic_Z italic_Y acting on the same pair of qubits do not appear consecutively, and make it hard for any generic transpiler to work efficiently. Thus we need to shuffle the total Hamiltonian at each Trotter step to enforce the above condition. The resultant circuit is in the following called DCQO-(ro), with ro signifying the re-ordering technique.

For IonQ implementation, we make use of its native two-qubit Mølmer-Sørensen (MS) gate. The evolution of each of the two-body Hamiltonian terms can be expressed with only one MS gate. In particular, the entangling gate RYY(θ)subscript𝑅𝑌𝑌𝜃R_{YY}(\theta)italic_R start_POSTSUBSCRIPT italic_Y italic_Y end_POSTSUBSCRIPT ( italic_θ ) can be implemented using MS gate as

RYY(θ)={MS(π/2,π/2,θ)θπ2(GPi(π/2)GPi(π/2))×MS(3π/2,π/2,πθ),π2<θπ(GPi(π/2)GPi(π/2))×MS(π/2,π/2,θπ),π<θ3π2MS(3π/2,π/2,2πθ),θ>3π2,RZY(θ)=(GPi2(π)I)RYY(θ)(GPi2(0)I),RYZ(θ)=(IGPi2(π))RYY(θ)(IGPi2(0)),formulae-sequencesubscript𝑅𝑌𝑌𝜃cases𝑀𝑆𝜋2𝜋2𝜃𝜃𝜋2tensor-product𝐺𝑃𝑖𝜋2𝐺𝑃𝑖𝜋2𝑀𝑆3𝜋2𝜋2𝜋𝜃𝜋2𝜃𝜋tensor-product𝐺𝑃𝑖𝜋2𝐺𝑃𝑖𝜋2𝑀𝑆𝜋2𝜋2𝜃𝜋𝜋𝜃3𝜋2𝑀𝑆3𝜋2𝜋22𝜋𝜃𝜃3𝜋2subscript𝑅𝑍𝑌𝜃tensor-product𝐺𝑃𝑖2𝜋𝐼subscript𝑅𝑌𝑌𝜃tensor-product𝐺𝑃𝑖20𝐼subscript𝑅𝑌𝑍𝜃tensor-product𝐼𝐺𝑃𝑖2𝜋subscript𝑅𝑌𝑌𝜃tensor-product𝐼𝐺𝑃𝑖20\begin{split}R_{YY}(\theta)&=\begin{cases}\!MS(\pi/2,\pi/2,\theta)&\theta\leq% \frac{\pi}{2}\\ (GPi(\pi/2)\otimes GPi(\pi/2))\times MS(3\pi/2,\pi/2,\pi-\theta),&\frac{\pi}{2% }<\theta\leq\pi\\ (GPi(\pi/2)\otimes GPi(\pi/2))\times MS(\pi/2,\pi/2,\theta-\pi),&\pi<\theta% \leq\frac{3\pi}{2}\\ MS(3\pi/2,\pi/2,2\pi-\theta),&\theta>\frac{3\pi}{2},\end{cases}\\ R_{ZY}(\theta)&=(GPi2(\pi)\otimes I)R_{YY}(\theta)(GPi2(0)\otimes I),\\ R_{YZ}(\theta)&=(I\otimes GPi2(\pi))R_{YY}(\theta)(I\otimes GPi2(0))\;,\end{split}start_ROW start_CELL italic_R start_POSTSUBSCRIPT italic_Y italic_Y end_POSTSUBSCRIPT ( italic_θ ) end_CELL start_CELL = { start_ROW start_CELL italic_M italic_S ( italic_π / 2 , italic_π / 2 , italic_θ ) end_CELL start_CELL italic_θ ≤ divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_CELL end_ROW start_ROW start_CELL ( italic_G italic_P italic_i ( italic_π / 2 ) ⊗ italic_G italic_P italic_i ( italic_π / 2 ) ) × italic_M italic_S ( 3 italic_π / 2 , italic_π / 2 , italic_π - italic_θ ) , end_CELL start_CELL divide start_ARG italic_π end_ARG start_ARG 2 end_ARG < italic_θ ≤ italic_π end_CELL end_ROW start_ROW start_CELL ( italic_G italic_P italic_i ( italic_π / 2 ) ⊗ italic_G italic_P italic_i ( italic_π / 2 ) ) × italic_M italic_S ( italic_π / 2 , italic_π / 2 , italic_θ - italic_π ) , end_CELL start_CELL italic_π < italic_θ ≤ divide start_ARG 3 italic_π end_ARG start_ARG 2 end_ARG end_CELL end_ROW start_ROW start_CELL italic_M italic_S ( 3 italic_π / 2 , italic_π / 2 , 2 italic_π - italic_θ ) , end_CELL start_CELL italic_θ > divide start_ARG 3 italic_π end_ARG start_ARG 2 end_ARG , end_CELL end_ROW end_CELL end_ROW start_ROW start_CELL italic_R start_POSTSUBSCRIPT italic_Z italic_Y end_POSTSUBSCRIPT ( italic_θ ) end_CELL start_CELL = ( italic_G italic_P italic_i 2 ( italic_π ) ⊗ italic_I ) italic_R start_POSTSUBSCRIPT italic_Y italic_Y end_POSTSUBSCRIPT ( italic_θ ) ( italic_G italic_P italic_i 2 ( 0 ) ⊗ italic_I ) , end_CELL end_ROW start_ROW start_CELL italic_R start_POSTSUBSCRIPT italic_Y italic_Z end_POSTSUBSCRIPT ( italic_θ ) end_CELL start_CELL = ( italic_I ⊗ italic_G italic_P italic_i 2 ( italic_π ) ) italic_R start_POSTSUBSCRIPT italic_Y italic_Y end_POSTSUBSCRIPT ( italic_θ ) ( italic_I ⊗ italic_G italic_P italic_i 2 ( 0 ) ) , end_CELL end_ROW (11)

where GPi𝐺𝑃𝑖GPiitalic_G italic_P italic_i and GPi2𝐺𝑃𝑖2GPi2italic_G italic_P italic_i 2 are native single-qubit gates. The above entangling operation can then in turn be used to construct the relevant operations for the DCQO circuit. Thus, we can represent the combined evolution of YZ𝑌𝑍YZitalic_Y italic_Z and ZY𝑍𝑌ZYitalic_Z italic_Y using only two native two-qubit gates, on both superconducting circuits and trapped ion hardware.

IV Results

In this section, we discuss quantum solutions to the logistics scheduling use cases introduced in sec. II, namely job-shop scheduling and travelling salesperson problems. We apply both pure quantum and hybrid quantum-classical counterdiabatic algorithms to solve different instances of these use cases, and additionally compare those solutions against the ones obtained using DQA and QAOA. Finally, we report hardware implementation results obtained from two different gate-based quantum processors, namely IonQ’s trapped ions and IBMQ’s superconducting circuits. In this work, we use the typical performance metrics, namely success probability (SP) and approximation ratio (AR), because we can compute the exact solutions of the tackle problem sizes. SP is the probability of finding the ground-state bitstrings, i.e. the bitstrings corresponding to the exact solutions, from all possible bitstrings in the output distribution. Whereas, AR measures the fraction of near-minimal bitstrings, and it is calculated as ratio between the average energy of the output distribution and the ground-state energy.

IV.1 (h-)DCQO for JSSP

Refer to caption
Refer to caption
Figure 5: Ideal simulation results comparing counterdiabatic algorithms against existing NISQ algorithms for solving 10 subproblems of JSSP. (a) DCQO circuits with T=0.1𝑇0.1T=0.1italic_T = 0.1 and N=4𝑁4N=4italic_N = 4 has the same number (1440absent1440\approx 1440≈ 1440) of CX gates as DQA circuits with T=1.0𝑇1.0T=1.0italic_T = 1.0 and N=6𝑁6N=6italic_N = 6. (b) With one layer, each of the h-DCQO and QAOA circuits has effectively 240 CX gates.

Due to the intractability of the full job-shop scheduling problem (sec. II A), we address only ten subproblems of size 16. These QUBO instances are fully-connected, making them challenging to be solved using NISQ hardware. For these instances, we first compare DCQO (h-DCQO) against DQA (QAOA) using noiseless simulations. In Fig. 5(a), we plot success probability and approximation ratios for the ten different instances considered as obtained from DCQO. We observe that with an equal circuit depth, DCQO has a SP (AR) of at least 100×\times× (2×2\times2 ×) greater than that of DQA on all instances. In particular, for QUBO ID 5, DQA fails to find the ground state. Due to such extremely low SPs and high CX counts, DQA is expected to fail in achieving the optimal solutions on NISQ hardware for these 16-qubit instances. Moreover for QUBO instances 1, 2, 7 and 9, DQA yields negative ARs, implying extremely poor solutions. On the other hand, even after omitting about 80% CX gates using the gate-cutoff technique as described in sec. III.3, DCQO is still able to find the ground states with SP exceeding 0.1% in noiseless simulations. In the following paragraph we show that DCQO, with gate cutoff, also performs well on NISQ hardware. In Fig 5(b), we observe that warm-started h-DCQO, described in sec. III.2, always yields better results, in terms of both SP and AR, than a single run of randomly-initialized QAOA. Although one can perform multiple random initialization of QAOA to search for a better result, our experience is that QAOA always falls behind h-DCQO.

Refer to caption
Figure 6: DCQO results for a 16-qubit subproblem of JSSP. The plot shows only the top few probabilities from the whole probability distribution. For T=0.002𝑇0.002T=0.002italic_T = 0.002, N=2𝑁2N=2italic_N = 2 and a gate-cutoff threshold of 0.1 (§III.3), the DCQO circuit has 142 MS gates. The IonQ emulator uses the noise model of Aria-2, which is also the hardware used for our experiment.

Among the ten subproblems, we notice that the last instance is ‘easy to solve’ in a sense that it has the highest SP among the ten instances studied. We further test the performance of DCQO for this instance both on a noisy simulator and a real quantum hardware to understand the impact of noise on DCQO’s performance and gauge the usability of DCQO for a 16-qubit optimization problem. In Fig. 6, we plot the bitstring spectrum obtained from noiseless and noisy DCQO simulations, as well as DCQO hardware experiments on IonQ. We notice that DCQO can identify the true ground state in a noisy simulation, although the success probability is reduced to one-quarter as compared to noiseless simulations. Using a total of 5000 measurements and noise mitigation, particularly debiasing, IonQ’s Aria-2 yields the exact solution of the QUBO problem with 3.5% probability and the solution bitstring being distinctly occupied with respect to other bitstrings. This performance is possible due to the all-to-all qubit connectivity of the hardware, along with its high two-qubit gate fidelity and coherence time. In contrast to this, the sparse qubit connectivity of ibm_brisbane requires twice as much two-qubit gates in the transpiled circuit. This massive increase in the two-qubit gate count, along with lower gate fidelities of IBMQ’s hardware, leads to very low quality in the solution for this problem instance.

Now we zoom in on the performance comparison between h-DCQO and QAOA for the last instance in Fig. 5(b). Here we report the best performance after iterating over 20 random initializations of QAOA circuit. In ideal simulation, h-DCQO is able to achieve 1.6×1.6\times1.6 × (3.5×\times×) higher AR (SP) than QAOA, see Fig. 7. By constructing the variational ansatz using CD terms only, we are able to obtain low-energy bitstrings with higher probability; this leads to both higher AR and SP for h-DCQO algorithm. To understand the impact of noise, we then run the final circuit, with optimal parameters, on IonQ’s processor. Although the hardware noise, even after error mitigation, lowers the quality of the noiseless results, h-DCQO circuit still achieves superior performance than the ideal QAOA circuit. In particular, the experiment yields a SP of 2.8% and an AR of 63%. Moreover, we notice that the SP of DCQO is 3.4×\times× more than that of QAOA, making DCQO superior. This is particularly interesting for NISQ applications because DCQO is a pure quantum algorithm and thus does not suffer from the challenges of variational quantum algorithms.

Refer to caption
Refer to caption
Figure 7: Comparing h-DCQO (ideal simulation) and QAOA (ideal simulation and IonQ implementation) for a 16-qubit subproblem of JSSP. Both QAOA and h-DCQO circuits have 240 MS gates. IonQ Aria-1 is used for the experiment and only the final circuit with optimized parameters is run on the hardware. (a) Energy distribution obtained from the final quantum circuit with optimized parameters. ARs are 50%, 80% and 63% for QAOA (ideal), h-DCQO (ideal) and h-DCQO (hardware), respectively. (b) Probability distribution obtained from the final quantum circuit with optimized parameters. SPs are 1.2%, 5.4% and 2.8% for QAOA (ideal), h-DCQO (ideal) and h-DCQO (hardware), respectively.

IV.2 (h-)DCQO for TSP

Having already established the superiority of counterdiabatic algorithms, i.e. DCQO and h-DCQO, over other NISQ algorithms, i.e. DQA and QAOA, for JSSP instances, we now present experimental results on using counterdiabatic algorithms for TSP instances. Although a three-city TSP instance is trivial as all possible paths are the shortest-distance paths, its map** as a 9-qubit Ising Hamiltonian makes it moderately challenging to be solved on NISQ hardware. In Fig. 8 the exact bitstring spectrum as well as spectra from DCQO ideal simulations, and DCQO hardware experiments carried out on IBMQ and IonQ hardware are plotted. One can see that a noiseless simulation of DCQO is able to identify all six degenerate ground states, i.e. all possible solutions of the three-city TSP, with high probability and just two Trotter steps. On the other hand, DQA, even with 11×11\times11 × more two-qubit gates, yields random outcomes and thus fails to solve this problem instance. A successful result from ibmq_guadalupe is possible for this TSP instance because of the low number of two-qubit gates in the transpiled circuit. Though the success probabilities and the quality of the solution are overall somewhat lower than the ones obtained from IonQ. For the latter experiments, the degeneracy is clearly restored. Although preserving degeneracy is not a requirement for optimization problems, it can become useful when studying perturbations to an open system resulting in a breakdown of degeneracy.

Refer to caption
Figure 8: DCQO results for a 9-qubit TSP instance. For T=0.2𝑇0.2T=0.2italic_T = 0.2, N=2𝑁2N=2italic_N = 2 and a gate-cutoff threshold of 0.1 (sec. III.3), the ideal DCQO circuit has 36 CX gates. Whereas the transpiled DCQO circuits run on IonQ Aria-2 and IBMQ ibmq_guadalupe have 36 MS gates and 66 CX gates, respectively.

We adapt the existing h-DCQO ansatz , see sec. III.2, for solving the 3-city and 4-city TSP instances. In particular, we construct an ansatz by parameterizing all one-body terms independently and all two-body terms together with a single parameter; this leads to Nq+1subscript𝑁q1N_{\text{q}}+1italic_N start_POSTSUBSCRIPT q end_POSTSUBSCRIPT + 1 trainable parameters for a Nqsubscript𝑁qN_{\text{q}}italic_N start_POSTSUBSCRIPT q end_POSTSUBSCRIPT-qubit circuit. Additionally, we omit all YZ𝑌𝑍YZitalic_Y italic_Z Pauli terms from the ansatz, which not only halves the required number of two-qubit (MS) gates but also results in a higher success probability. For the 9-qubit TSP instance, we warm-start the parameters using a DCQO circuit. The variational training of the ten parameters took about 120 iterations for the simulated expectation value of the problem Hamiltonian Eq. (2) to converge to the exact ground-state energy with high precision. The final quantum circuit, with the optimal parameters, yields only one out of six of the degenerate ground states with a success probability of nearly 71%, as seen in Fig. 9(a). This result is valid because all degenerate ground states in TSP encode the same travel path. Upon running the final circuit on IBMQ’s hardware, we are able to extract the desired solution with about 15% probability. This is possible due to a low CX count of 100 in the circuit and the already high success probability in ideal simulations.

Refer to caption
Refer to caption
Figure 9: Hybrid DCQO results for TSP instances. The h-DCQO ansatz has one layer consisting of Y𝑌Yitalic_Y and ZY𝑍𝑌ZYitalic_Z italic_Y terms, with a coefficient threshold of 0.1. The hybrid optimization is performed on a noiseless simulation and only the final circuit with optimized parameters is run on the hardware. (a) 3-city TSP instance. IBMQ’s ibmq_guadalupe is used for the experiment and the transpiled circuit has 100 CX gates. (b) 4-city TSP instance. IonQ Aria-1 is used for the experiment and the transpiled circuit has 48 MS gates.

For the 4-city TSP, i.e. with 16 qubits, DCQO is unable to discern the eight degenerate solutions with significant probabilities as it could do for the 3-city TSP. On the other hand, h-DCQO successfully finds a solution with a very high probability, see Fig. 9(b). For this problem instance, the h-DCQO ansatz has 17 variational parameters, which we train by scanning over 20 different random initializations. Warm starting h-DCQO fails in this case because the chances of getting stuck in a local minima increase significantly with more parameters. From our hardware implementations, we observe that IonQ hardware, with only 5000 shots and error mitigation, yields the shortest-distance path with a high probability of 45%. However, ibm_brisbane fails for this instance because the error from the 180 two-qubit gates of the transpiled circuit results in a random output distribution.

V Conclusions and outlook

We have used our DCQO algorithm and its hybrid version to solve 9- and 16-qubit Ising spin glass problems on current quantum hardware. The problem instances are taken from physically motivated and industrially relevant combinatorial optimization problems, namely the Traveling Salesperson Problem (TSP) and a Job Shop Scheduling Problem (JSSP). Using ideal simulations, we show that DCQO consistently outperforms DQA and QAOA in terms of success probability and required number of two-qubit gates. The main challenge was to go beyond simple ideal simulation. Realizing a hardware implementation of an all-to-all connected problem, even for the problem sizes addressed here, quickly approaches the limits of currently available NISQ devices. Therefore, we implemented additional circuit modifications using several circuit reduction strategies, including gate-cutoff and gate-reordering, to further reduce the depth of our compressed circuits. This is followed by an efficient transpilation for the given hardware, resulting in a successful implementation of our algorithms on NISQ devices.

We use both superconducting circuits and trapped ions to solve the use case instances. For the TSP, we successfully solve a 3-city problem instance on an IBM superconducting machine, as well as 3-city and 4-city instances on an IonQ trapped-ion machine. We also investigated the simplest form of a non-trivial JSSP for a given laboratory application, which would require a QUBO representation with 360 variables. Since such a problem is still too large to tackled with current quantum hardware, requiring access to a 360-qubit machine with high two-qubit gate fidelity and dense qubit connectivity, we solve only subproblems of size 16, requiring 16 qubits. We show that h-DCQO outperforms QAOA in terms of success probability and approximation ratio for all 16-qubit instances tested here. Due to resource limitations, we perform the variational training classically using a noiseless simulator and run only the final circuit on the quantum hardware.

We have identified the algorithmic and the implementation challenges of our (hybrid) DCQO algorithm. The success probability obtained from DCQO decreases exponentially with increasing number of qubits, making it currently infeasible for solving Ising problems beyond 20 qubits. On the other hand, h-DCQO can handle larger size problems due to the flexibility in the construction of the ansatz. In this algorithm, we combine the advantages of variational training of parameterized quantum circuit with counterdiabatic driving. Athough hybrid methods may be able to tackle larger problems and yield high approximation ratio, there is no guarantee of high success probability since the algorithm may suffer from the appearance of barren plateaus and local minima.

We can test the performance of other variants of counterdiabatic optimization algorithms for logistics scheduling use cases. We recently proposed the bias-field DCQO, which demonstrates an empirical scaling advantage, and use it to successfully solve spin-glass problem instances with up to 100 qubits on NISQ hardware [45]. In this approach, we incorporate specific information about the target state into the initial Hamiltonian, achieved by introducing an inhomogeneous longitudinal field along with the standard transverse field. On the other hand, it would be both interesting and challenging to investigate the performance of counterdiabatic optimization algorithms that include higher-order NC terms, as compared to DCQO, which has only first-order terms. Finally, we can also make use of our digital-analog counterdiabatic optimization algorithms to address industrially relevant use cases [46].

On the application side, possible next steps are as follows. First, one could apply our optimization algorithms to solve larger TSP instances by using an encoding scaling of nlog(n)𝑛𝑛n\log(n)italic_n roman_log ( italic_n ), i.e. using only nlog(n)𝑛𝑛n\log(n)italic_n roman_log ( italic_n ) qubits instead of the currently used n2superscript𝑛2n^{2}italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT qubits for a n𝑛nitalic_n-city TSP instance. Second, one could explore our quantum optimization algorithms for a more-involved logistics use case, such as the vehicle routing problem. This has previously been tackled using quantum annealing and variational quantum algorithms [47]. Using an improved bit-to-qubit encoding [48], our DCQO algorithm has the potential to solve large instances without having to deal with the large measurement overhead of variational optimization methods. Third, we can use the large neighborhood search algorithm [20] in conjunction with DCQO to solve a simple, non-trivial JSSP instance. The performance of this hybrid digital-digital optimization algorithm can be compared to D-Wave’s Hybrid Solver, i.e. a hybrid digital-analog algorithm, for the robot scheduling problem [18].

Appendix A Quantum optimization

In the following, we briefly outline quantum algorithms used to solving combinatorial optimization problems.

A.1 Adiabatic quantum optimization

Since adiabatic quantum computing (AQC) with non-stoquastic Hamiltonians is equivalent with the circuit model of quantum computing [49], any computational problem can in principle be treated with AQC. Therefore, AQC represents an innovative approach to tackle optimization problems using analog quantum processors [50], an approach known as adiabatic quantum optimization (AQO). Central to AQC is the adiabatic theorem [51] which states that a physical system remains in an instantaneous eigenstate of its time-dependent Hamiltonian if a perturbation acts slow enough and the system exhibits a gap between the respective state and the rest of the Hamiltonian spectrum. In ideal AQC, a quantum system being in a ground state of a given initial Hamiltonian Hisubscript𝐻iH_{\text{i}}italic_H start_POSTSUBSCRIPT i end_POSTSUBSCRIPT, which is easy to prepare on a quantum device, is driven to the ground state of a final Hamiltonian Hpsubscript𝐻pH_{\text{p}}italic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT, which encodes the solution of the computational problem. The corresponding adiabatic dynamics is governed by a time-dependent adiabatic Hamiltonian defined as:

Had(t)=(1λ(t))Hi+λ(t)Hp,subscript𝐻ad𝑡1𝜆𝑡subscript𝐻i𝜆𝑡subscript𝐻pH_{\text{ad}}(t)=\left(1-\lambda(t)\right)H_{\text{i}}+\lambda(t)H_{\text{p}},italic_H start_POSTSUBSCRIPT ad end_POSTSUBSCRIPT ( italic_t ) = ( 1 - italic_λ ( italic_t ) ) italic_H start_POSTSUBSCRIPT i end_POSTSUBSCRIPT + italic_λ ( italic_t ) italic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT , (12)

where the scheduling function λ(t)[0,1]𝜆𝑡01\lambda(t)\in[0,1]italic_λ ( italic_t ) ∈ [ 0 , 1 ] is continuously differentiable, and it guides the transition from the initial to the problem Hamiltonian. A common approach to Hisubscript𝐻𝑖H_{i}italic_H start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in AQO is a transverse field

Hi=iσix,subscript𝐻isubscript𝑖superscriptsubscript𝜎𝑖𝑥H_{\text{i}}=-\sum_{i}\sigma_{i}^{x},italic_H start_POSTSUBSCRIPT i end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT , (13)

where σxsuperscript𝜎𝑥\sigma^{x}italic_σ start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT is a Pauli matrix in Dirac notation, and whose ground state is a uniform superposition state in the computational basis. For the optimization problems treated in this work, we assume for the problem Hamiltonian Hpsubscript𝐻pH_{\text{p}}italic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT an Ising spin-glass model according to Eq. (2). A transverse field appears to be a proper choice here because then Hisubscript𝐻iH_{\text{i}}italic_H start_POSTSUBSCRIPT i end_POSTSUBSCRIPT and Hpsubscript𝐻pH_{\text{p}}italic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT do not commute. The implementation challenge of AQO is to make the dynamics slow enough to avoid unwanted transitions to eigenstates different from the ground state while the adiabatic evolution is carried out in finite time. This leads to the concept of finite-time AQO which is typically referred to as quantum annealing.

A.2 Quantum annealing

Quantum annealing is a meta-heuristic optimisation procedure that aims at solving a combinatorial optimisation problem using properties of quantum physics, and not requiring adiabatic evolution and universality [34]. This technique has been commercialized by D-Wave Systems, which owns the state-of-the-art quantum annealer with 5000+ qubits. The quantum annealing process is contingent on the preservation of quantum coherence. A prolonged annealing process, in comparison to the coherence time of the qubits, can diminish the quantum attributes of the system. Besides noisy qubits, one of the other inherent limitations of these annealers is the restricted connectivity between the qubits. This architectural constraint impedes the direct implementation of Hpsubscript𝐻pH_{\text{p}}italic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT, often compelling the use of embedding techniques, which comes at the cost of additional qubit requirement. Moreover, it is challenging to implement a non-stochastic problem Hamiltonian on these annealers. Thus, although D-Wave annealers are powerful for sampling near-optimal solutions for QUBO problems, it is not suitable for universal computation, i.e. not any computational problem is treatable on these devices. In contrast to that, digitized adiabatic quantum computing supports universality and error correction, and has been successfully implemented on a superconducting circuit with nine qubits [52].

A.3 Variational quantum algorithms

Due to the above-mentioned challenges of analog quantum annealing and flexibility of gate-based quantum processors, quantum optimization algorithms are being developed for NISQ hardware. The two popular NISQ algorithms for combinatorial optimization are quantum approximate optimization algorithm (QAOA) and variational quantum eigensolver (VQE) [53]. Both are hybrid quantum-classical algorithms in a sense that a solution is reached variationally by making a classical and quantum computer work together in tandem. The two main components are parameterized quantum circuit, which acts typically on an equal superposition initial state to construct a parameterized quantum state, and classical optimizer. The expectation value of the problem Hamiltonian Eq. (2) is estimated quantumly for the parameterized state, whose parameters are updated in an iterative manner using the classical optimizer such that the expectation value is minimized. The parameterized quantum circuit in QAOA has alternating layers of unitaries constructed from the problem Hamiltonian and a mixer Hamiltonian Eq. (13), respectively [36]. A successful implementation of QAOA on digital hardware would require deep circuits, and hence infeasible on NISQ devices. On the other hand, VQE for optimization employ single- and two-qubit gates available on a given hardware to construct the parameterized circuits [35]. This hardware-efficient VQE, although easily implementable on NISQ hardware, does not guarantee optimal solution.

References

  • Bauer et al. [2020] B. Bauer, S. Bravyi, M. Motta, and G. K.-L. Chan, Quantum algorithms for quantum chemistry and quantum materials science, Chemical Reviews 120, 12685 (2020).
  • Herman et al. [2023] D. Herman, C. Googin, X. Liu, Y. Sun, A. Galda, I. Safro, M. Pistoia, and Y. Alexeev, Quantum computing for finance, Nature Reviews Physics 5, 450 (2023).
  • Grover [1996] L. K. Grover, A fast quantum mechanical algorithm for database search, arXiv e-prints , quant-ph/9605043 (1996)arXiv:quant-ph/9605043 [quant-ph] .
  • Arute et al. [2019] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, R. Biswas, S. Boixo, F. G. S. L. Brandao, D. A. Buell, B. Burkett, Y. Chen, Z. Chen, B. Chiaro, R. Collins, W. Courtney, A. Dunsworth, E. Farhi, B. Foxen, A. Fowler, C. Gidney, M. Giustina, R. Graff, K. Guerin, S. Habegger, M. P. Harrigan, M. J. Hartmann, A. Ho, M. Hoffmann, T. Huang, T. S. Humble, S. V. Isakov, E. Jeffrey, Z. Jiang, D. Kafri, K. Kechedzhi, J. Kelly, P. V. Klimov, S. Knysh, A. Korotkov, F. Kostritsa, D. Landhuis, M. Lindmark, E. Lucero, D. Lyakh, S. Mandrà, J. R. McClean, M. McEwen, A. Megrant, X. Mi, K. Michielsen, M. Mohseni, J. Mutus, O. Naaman, M. Neeley, C. Neill, M. Y. Niu, E. Ostby, A. Petukhov, J. C. Platt, C. Quintana, E. G. Rieffel, P. Roushan, N. C. Rubin, D. Sank, K. J. Satzinger, V. Smelyanskiy, K. J. Sung, M. D. Trevithick, A. Vainsencher, B. Villalonga, T. White, Z. J. Yao, P. Yeh, A. Zalcman, H. Neven, and J. M. Martinis, 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, P. Hu, X.-Y. Yang, W.-J. Zhang, H. Li, Y. Li, X. Jiang, L. Gan, G. Yang, L. You, Z. Wang, L. Li, N.-L. Liu, C.-Y. Lu, and J.-W. Pan, Quantum computational advantage using photons, Science 370, 1460 (2020)https://www.science.org/doi/pdf/10.1126/science.abe8770 .
  • Kim et al. [2023] Y. Kim, A. Eddins, S. Anand, K. X. Wei, E. van den Berg, S. Rosenblatt, H. Nayfeh, Y. Wu, M. Zaletel, K. Temme, and A. Kandala, Evidence for the utility of quantum computing before fault tolerance, Nature 618, 500 (2023).
  • Hegade et al. [2021] N. N. Hegade, K. Paul, Y. Ding, M. Sanz, F. Albarrán-Arriagada, E. Solano, and X. Chen, Shortcuts to adiabaticity in digitized adiabatic quantum computing, Physical Review Applied 15, 024038 (2021).
  • Hegade et al. [2022] N. N. Hegade, X. Chen, and E. Solano, Digitized counterdiabatic quantum optimization, Physical Review Research 4, L042030 (2022).
  • Chandarana et al. [2023] P. Chandarana, N. N. Hegade, I. Montalban, E. Solano, and X. Chen, Digitized counterdiabatic quantum algorithm for protein folding, Phys. Rev. Appl. 20, 014024 (2023).
  • Chandarana et al. [2022a] P. Chandarana, P. S. Vieites, N. N. Hegade, E. Solano, Y. Ban, and X. Chen, Meta-learning digitized-counterdiabatic quantum optimization, arXiv preprint arXiv:2206.09966  (2022a).
  • Chandarana et al. [2022b] P. Chandarana, N. N. Hegade, K. Paul, F. Albarrán-Arriagada, E. Solano, A. Del Campo, and X. Chen, Digitized-counterdiabatic quantum approximate optimization algorithm, Physical Review Research 4, 013141 (2022b).
  • Hegade and Solano [2023] N. N. Hegade and E. Solano, Digitized-counterdiabatic quantum factorization, arXiv preprint arXiv:2301.11005  (2023).
  • Gomez Cadavid et al. [2023] A. Gomez Cadavid, I. Montalban, A. Dalal, E. Solano, and N. N. Hegade, Efficient DCQO Algorithm within the Impulse Regime for Portfolio Optimization, arXiv e-prints , arXiv:2308.15475 (2023)arXiv:2308.15475 [quant-ph] .
  • Guan et al. [2023] H. Guan, F. Zhou, F. Albarrán-Arriagada, X. Chen, E. Solano, N. N. Hegade, and H.-L. Huang, Single-Layer Digitized-Counterdiabatic Quantum Optimization for p𝑝pitalic_p-spin Models, arXiv e-prints , arXiv:2311.06682 (2023)arXiv:2311.06682 [quant-ph] .
  • Demirplak and Rice [2003] M. Demirplak and S. A. Rice, Adiabatic population transfer with control fields, The Journal of Physical Chemistry A 107, 9937 (2003).
  • Berry [2009] M. V. Berry, Transitionless quantum driving, Journal of Physics A: Mathematical and Theoretical 42, 365303 (2009).
  • del Campo [2013] A. del Campo, Shortcuts to adiabaticity by counterdiabatic driving, Physical review letters 111, 100502 (2013).
  • Leib et al. [2023] D. Leib, T. Seidel, S. Jäger, R. Heese, C. Jones, A. Awasthi, A. Niederle, and M. Bortz, An optimization case study for solving a transport robot scheduling problem on quantum-hybrid and quantum-inspired hardware, Scientific Reports 13, 18743 (2023).
  • Lucas [2014] A. Lucas, Ising formulations of many NP problems, Frontiers in Physics 2, 5 (2014)arXiv:1302.5843 [cond-mat.stat-mech] .
  • Pisinger and Ropke [2019] D. Pisinger and S. Ropke, Large neighborhood search, in Handbook of Metaheuristics, edited by M. Gendreau and J.-Y. Potvin (Springer International Publishing, Cham, 2019) pp. 99–127.
  • Rosenberg et al. [2016] G. Rosenberg, M. Vazifeh, B. Woods, and E. Haber, Building an iterative heuristic solver for a quantum annealer, Computational Optimization and Applications 65, 845 (2016).
  • Booth et al. [2017] M. Booth, S. P. Reinhardt, and A. Roy, Partitioning optimization problems for hybrid classical/quantum execution technical report (2017).
  • Ponce et al. [2023] M. Ponce, R. Herrman, P. C. Lotshaw, S. Powers, G. Siopsis, T. Humble, and J. Ostrowski, Graph decomposition techniques for solving combinatorial optimization problems with variational quantum algorithms, arXiv e-prints , arXiv:2306.00494 (2023)arXiv:2306.00494 [quant-ph] .
  • Osaba et al. [2020] E. Osaba, X.-S. Yang, and J. Del Ser, Chapter 9 - Traveling salesman problem: a perspective review of recent research and new results with bio-inspired metaheuristics, in Nature-Inspired Computation and Swarm Intelligence (Academic Press, 2020) pp. 135–164.
  • Mandal and Kumar Deva Sarma [2022] A. K. Mandal and P. Kumar Deva Sarma, Novel applications of ant colony optimization with the traveling salesman problem in dna sequence optimization, in 2022 IEEE 2nd International Symposium on Sustainable Energy, Signal Processing and Cyber Security (iSSSC) (2022) pp. 1–6.
  • Nałęcz-Charkiewicz and Nowak [2022] K. Nałęcz-Charkiewicz and R. M. Nowak, Algorithm for DNA sequence assembly by quantum annealing, BMC Bioinformatics 23, 122 (2022).
  • Jain [2021] S. Jain, Solving the traveling salesman problem on the d-wave quantum computer, Frontiers in Physics 910.3389/fphy.2021.760783 (2021).
  • Srinivasan et al. [2018] K. Srinivasan, S. Satyajit, B. K. Behera, and P. K. Panigrahi, Efficient quantum algorithm for solving travelling salesman problem: An IBM quantum experience, arXiv e-prints , arXiv:1805.10928 (2018)arXiv:1805.10928 [quant-ph] .
  • Khumalo et al. [2021] M. T. Khumalo, H. A. Chieza, K. Prag, and M. Woolway, An investigation of IBM Quantum Computing device performance on Combinatorial Optimisation Problems, arXiv e-prints , arXiv:2107.03638 (2021)arXiv:2107.03638 [quant-ph] .
  • Ramezani et al. [2024] M. Ramezani, S. Salami, M. Shokhmkar, M. Moradi, and A. Bahrampour, Reducing the Number of Qubits from n2superscript𝑛2n^{2}italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to nlog2(n)𝑛subscript2𝑛n\log_{2}(n)italic_n roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_n ) to Solve the Traveling Salesman Problem with Quantum Computers: A Proposal for Demonstrating Quantum Supremacy in the NISQ Era, arXiv e-prints , arXiv:2402.18530 (2024)arXiv:2402.18530 [quant-ph] .
  • Qian et al. [2023] W. Qian, R. A. M. Basili, M. M. Eshaghian-Wilner, A. Khokhar, G. Luecke, and J. P. Vary, Comparative Study of Variations in Quantum Approximate Optimization Algorithms for the Traveling Salesman Problem, Entropy 25, 1238 (2023)arXiv:2307.07243 [quant-ph] .
  • Glover et al. [2018] F. Glover, G. Kochenberger, and Y. Du, A Tutorial on Formulating and Using QUBO Models, arXiv e-prints , arXiv:1811.11538 (2018)arXiv:1811.11538 [cs.DS] .
  • Bian et al. [2010] Z. Bian, F. A. Chudak, W. G. Macready, and G. Rose, The ising model : teaching an old problem new tricks (2010).
  • Kadowaki and Nishimori [1998] T. Kadowaki and H. Nishimori, Quantum annealing in the transverse Ising model, Phys. Rev. E 58, 5355 (1998).
  • Moll et al. [2018] N. Moll, P. Barkoutsos, L. S. Bishop, J. M. Chow, A. Cross, D. J. Egger, S. Filipp, A. Fuhrer, J. M. Gambetta, M. Ganzhorn, A. Kandala, A. Mezzacapo, P. Müller, W. Riess, G. Salis, J. Smolin, I. Tavernelli, and K. Temme, Quantum optimization using variational algorithms on near-term quantum devices, Quantum Science and Technology 3, 030503 (2018).
  • Farhi et al. [2014] E. Farhi, J. Goldstone, and S. Gutmann, A Quantum Approximate Optimization Algorithm, arXiv e-prints , arXiv:1411.4028 (2014)arXiv:1411.4028 [quant-ph] .
  • Kolodrubetz et al. [2017] M. Kolodrubetz, D. Sels, P. Mehta, and A. Polkovnikov, Geometry and non-adiabatic response in quantum and classical systems, Physics Reports 697, 1 (2017), geometry and non-adiabatic response in quantum and classical systems.
  • Sels and Polkovnikov [2017] D. Sels and A. Polkovnikov, Minimizing irreversible losses in quantum systems by local counterdiabatic driving, Proceedings of the National Academy of Sciences 114, E3909 (2017).
  • Claeys et al. [2019] P. W. Claeys, M. Pandey, D. Sels, and A. Polkovnikov, Floquet-engineering counterdiabatic protocols in quantum many-body systems, Physical Review Letters 123, 090602 (2019).
  • Hatomura and Takahashi [2021] T. Hatomura and K. Takahashi, Controlling and exploring quantum systems by algebraic expression of adiabatic gauge potential, Physical Review A 103, 012220 (2021).
  • Takahashi and del Campo [2023] K. Takahashi and A. del Campo, Shortcuts to adiabaticity in krylov space, arXiv preprint arXiv:2302.05460  (2023).
  • Kovalsky et al. [2023] L. K. Kovalsky, F. A. Calderon-Vargas, M. D. Grace, A. B. Magann, J. B. Larsen, A. D. Baczewski, and M. Sarovar, Self-healing of trotter error in digital adiabatic state preparation, Physical Review Letters 131, 060602 (2023).
  • [43] Getting started with native gates, https://ionq.com/docs/getting-started-with-native-gates, accessed: 2023-08-22.
  • Tranter et al. [2019] A. Tranter, P. J. Love, F. Mintert, N. Wiebe, and P. V. Coveney, Ordering of Trotterization: Impact on Errors in Quantum Simulation of Electronic Structure, Entropy 21, 1218 (2019)arXiv:1912.07555 [quant-ph] .
  • Gomez Cadavid et al. [2024] A. Gomez Cadavid, A. Dalal, A. Simen, E. Solano, and N. N. Hegade, Bias-field digitized counterdiabatic quantum optimization, arXiv e-prints , arXiv:2405.13898 (2024), arXiv:2405.13898 [quant-ph] .
  • Kumar et al. [2024] S. Kumar, N. N. Hegade, A. Gomez Cadavid, M. H. de Oliveira, E. Solano, and F. Albarrán-Arriagada, Digital-Analog Counterdiabatic Quantum Optimization with Trapped Ions, arXiv e-prints , arXiv:2405.01447 (2024)arXiv:2405.01447 [quant-ph] .
  • Osaba et al. [2022] E. Osaba, E. Villar-Rodriguez, and I. Oregi, A systematic literature review of quantum computing for routing problems, IEEE Access 10, 55805 (2022).
  • Leonidas et al. [2023] I. D. Leonidas, A. Dukakis, B. Tan, and D. G. Angelakis, Qubit efficient quantum algorithms for the vehicle routing problem on NISQ processors, arXiv e-prints , arXiv:2306.08507 (2023)arXiv:2306.08507 [quant-ph] .
  • Aharonov et al. [2008] D. Aharonov, W. van Dam, J. Kempe, Z. Landau, S. Lloyd, and O. Regev, Adiabatic quantum computation is equivalent to standard quantum computation, SIAM Review 50, 755 (2008)https://doi.org/10.1137/080734479 .
  • Farhi et al. [2001] E. Farhi, J. Goldstone, S. Gutmann, J. Lapan, A. Lundgren, and D. Preda, A quantum adiabatic evolution algorithm applied to random instances of an np-complete problem, Science 292, 472 (2001)https://www.science.org/doi/pdf/10.1126/science.1057726 .
  • Farhi et al. [2000] E. Farhi, J. Goldstone, S. Gutmann, and M. Sipser, Quantum computation by adiabatic evolution, arXiv:quant-ph/0001106  (2000).
  • Barends et al. [2016] R. Barends, A. Shabani, L. Lamata, J. Kelly, A. Mezzacapo, U. L. Heras, R. Babbush, A. G. Fowler, B. Campbell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, E. Jeffrey, E. Lucero, A. Megrant, J. Y. Mutus, M. Neeley, C. Neill, P. J. J. O’Malley, C. Quintana, P. Roushan, D. Sank, A. Vainsencher, J. Wenner, T. C. White, E. Solano, H. Neven, and J. M. Martinis, Digitized adiabatic quantum computing with a superconducting circuit, Nature 534, 222 (2016).
  • Symons et al. [2023] B. C. B. Symons, D. Galvin, E. Sahin, V. Alexandrov, and S. Mensa, A practitioner’s guide to quantum algorithms for optimisation problems, Journal of Physics A Mathematical General 56, 453001 (2023)arXiv:2305.07323 [quant-ph] .