Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: arXiv.org perpetual non-exclusive license
arXiv:2309.00597v2 [cs.CE] 08 Dec 2023

The QUATRO Application Suite: Quantum Computing for Models of Human Cognition

Raghavendra Pradyumna Pothukuchi1    Leon Lufkin1    Yu Jun Shen1*    Alejandro Simon1*   
Rome Thorstenson1
   Bernardo Eilert Trevisan1*    Michael Tu1    Mudi Yang1*    Ben Foxman1   
Viswanatha Srinivas Pothukuchi2
   Gunnar Epping4    Thi Ha Kyaw3    Bryant J Jongkees5   
Yongshan Ding1
   Jerome R Busemeyer4    Jonathan D Cohen6    Abhishek Bhattacharjee1 Equal contribution. Listed in alphabetical order of last name.
Equal contribution. Listed in alphabetical order of last name.
*Worked as Yale University undergraduate.
( 1Yale University 2Independent 3LG Electronics Toronto AI Lab 4Indiana University 5Leiden University 6Princeton University )
Abstract

Research progress in quantum computing has, thus far, focused on a narrow set of application domains. Expanding the suite of quantum application domains is vital for the discovery of new software toolchains and architectural abstractions. In this work, we unlock a new class of applications ripe for quantum computing research—computational cognitive modeling. Cognitive models are critical to understanding and replicating human intelligence. Our work connects computational cognitive models to quantum computer architectures for the first time. We release QUATRO 111QUATRO stands for quantum applications from cognitive modeling, and also means four, which is the number of cognitive model types we study., a collection of quantum computing applications from cognitive models. The development and execution of QUATRO shed light on gaps in the quantum computing stack that need to be closed to ease programming and drive performance. Among several contributions, we propose and study ideas pertaining to quantum cloud scheduling (using data from gate- and annealing-based quantum computers), parallelization, and more. In the long run, we expect our research to lay the groundwork for more versatile quantum computer systems in the future.

1 Introduction

Quantum computers are evolving from early experimental prototypes to more mature cloud platforms [1, 2, 3, 4, 5], but their success now hinges on expanding their utility to more application domains. Unfortunately, most deployed quantum computing applications have thus far focused on select domains like physics, chemistry, finance, and recently, machine learning [6]. This presents a chicken-and-egg problem. More general-purpose computing stacks would expand the benefits of quantum computing to more application domains. But, the paucity of new application domains has over-specialized quantum computing software to just a few application domains. New application domains are needed to shed light on more general-purpose quantum computing stacks of the future.

In this work, we take a step towards breaking this vicious cycle of over-specialization by presenting computational cognitive modeling as a new target application domain for quantum computing. Cognitive models explain how humans process information and make decisions, and have been foundational to psychology, psychiatry, artificial intelligence (AI), machine learning (ML), and even economics [7, 8, 9]. These models are now adequately computationally demanding and attracting the attention of computer systems practitioners [10], making them a natural candidate for quantum computing research. Models used for relational reasoning [11] and planning [12] rely on simultaneous constraint satisfaction kernels, which are known to be slow and scale poorly with model complexity.

Recently, cognitive scientists have also shown the prospect of using quantum probability theory to develop a unified and formal decision-making framework that explains human behavior difficult to be captured by existing classical alternatives, like biases, effects of question framing and order etc. [8, 13, 14]. Rigorous psychological studies with humans have shown evidence supporting the expressive power of quantum-theoretic frameworks and their improved modeling accuracy over existing methods [15, 16, 17, 18, 8, 19, 14, 13]. However, realizing the full potential of these methods requires developing and mapping larger models onto real quantum machines [8], a hitherto unexplored study.

Ours is the first work to map cognitive models to quantum hardware, offering a path to scale and augment them, and enable more human-like AI. We transcend prior work [20], which focuses only on small kernels or circuit components, to develop QUATRO, a suite of full, computationally hard classical and quantum-theoretic models, mapped to real quantum hardware.

Ours is also the first work to showcase algorithmic advances to realize constructs necessitated by cognitive models. Examples include constructs like subspace projections for quantum walks, implemented with new low-cost state-detection circuits. Although mapping cognitive models to quantum hardware drives the discovery of these constructs, they are broadly applicable including, e.g., quantum program assertions [21, 22].

Finally, by developing bridges between the cognitive sciences and quantum computing, we lay the foundations for new research in quantum computing software toolchains, programming languages, compiler support, control interfaces, reorganization of the quantum cloud, and more. These research questions go beyond the hardware-centric focus of existing work (e.g., error mitigation [23, 24, 25], reliability [26, 27, 28, 29, 30, 31], circuit synthesis [32, 33], and microarchitecture [34]). To undergird our results, we evaluate new parallelism-aware cloud scheduling approaches on real quantum systems. Our approach offers performance speedups via parallelization and improved system throughput, representing new axes for performance scaling in quantum clouds. Like our newly discovered algorithmic constructs, these optimizations are broadly useful beyond cognitive modeling.

In summary, our specific research contributions are:

  1. 1.

    We introduce QUATRO, end-to-end cognitive models as a new application domain for quantum computing. QUATRO is suitable for current and future quantum hardware.

  2. 2.

    We discover new circuit-level constructs on subspace projections, techniques that benefit domains beyond cognitive modeling, with a fraction of resources over prior work.

  3. 3.

    We identify shortcomings and opportunities for research innovation in the programming stack for quantum computers. Unlike all prior work, ours transcends low-level hardware and addresses higher-level aspects of the quantum stack.

  4. 4.

    We propose and analyze new quantum cloud organization using data from real quantum computers, showing 2.4×\times×–10×\times× speedup for co-designed applications.

Overall, our work expands quantum computing into a new high-impact application domain and sets the stage for building versatile quantum systems that are easier and faster to use.

2 Motivation

Virtuous cycles among application domains, algorithms, and computer systems have been central to the decades-long success of computing. Consider, e.g., the advances in GPU and tensor hardware that have unlocked increasingly complex deep neural networks in the last decade [35]. Unfortunately, quantum computing has yet to experience such rapid growth in application-systems co-design. Quantum applications are concentrated in physics [36], chemistry [37], finance [38, 39], machine learning [40], and optimization [41]. New applications are needed to ensure that the potential of quantum computing is more broadly harnessed by society [42, 6].

We bridge the cognitive sciences with quantum computing as a step in enabling new families of quantum applications. Computational cognitive models describe how humans generate decisions and behavior [43, 44, 45]. Such insights have enabled breakthroughs in AI [7], ML [46, 47], and behavioral economics that resulted in multiple Nobel prizes (Simon in 1978; Kahneman in 2002). Cognitive models are being studied today with the anticipation of identifying theoretical mechanisms to explain complex human behavior, and unlock AI that is not captured by modern deep learning [48, 49].

State-of-the-art cognitive models are computationally challenging, and are a natural target for quantum computing research. Models addressing fundamental human capabilities like relational reasoning and problem-solving involve simultaneous constraint satisfaction, and are realized with graph isomorphism techniques [11], combinatorial optimization and planning [12], which are proven to be hard to scale classically.

Furthermore, there is a new class of cognitive models that employs quantum probability as a theoretical tool to describe human decision-making. These models leverage the richer axioms of quantum probability such as projections, non-commutativity, and superposition, to describe complex behavior that has resisted explanation from all existing classical alternatives and/or identify parsimonious models with exponentially fewer parameters [8, 15]. Below are two illustrative results.

As an example of the expressiveness of quantum cognitive models, consider a two-stage task where individuals choose to play or quit a gamble before each stage [50]. When the outcome of the first stage was known—regardless of whether it is a win or a loss, real-world data showed individuals generally preferred to play the gamble in the second stage. However, when the outcome was unknown, most individuals switched and prefer to quit, which violates the classic law of total probability, and so existing models cannot explain such “irrational” behavior.

Alternatively, a quantum cognitive model representing decision-making as a superposition over possible outcomes, predicts different results because of quantum interference effects that can generate classical violations of total probability [15].

Quantum cognition models also provide lower dimensional representations of cognition and thus greater parsimony compared to classical models. Consider a strategic game involving n𝑛nitalic_n players, each choosing one of K𝐾Kitalic_K possible actions [18]. Classically representing an individual’s beliefs about the actions of all players requires Knsuperscript𝐾𝑛K^{n}italic_K start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT joint probabilities, representing an exponential growth in model dimension. Instead, the same could be modeled with quantum probability using a single K𝐾Kitalic_K-dimensional vector space, evaluated with n𝑛nitalic_n different bases representing the player’s different view points of the game.

While there could be a future classical formalism that can match quantum cognition in expressiveness, model conciseness, and accuracy, quantum cognition provides some of the best empirically-validated tools today to explain complex behavior [13, 16, 8, 14, 51, 17, 18, 52, 19, 53, 54, 15]. With commercial quantum computers being available, the timing is ideal to explore the use of quantum computers for large-scale modeling of human-level cognition.

Recently, a compilation framework, Distill [10], has been proposed to accelerate classical cognitive models on classical systems like multicore CPUs and GPUs. However, there is value in exploring how computationally hard cognitive models, including quantum cognitive models, can be mapped to quantum computers, which we discuss next.

2.1 Why QUATRO Applications?

Our cognitive models are full-scale applications that introduce richer and newer constructs not found in architecture benchmarking kernels like SuperMarq [55]. Such constructs include Hamiltonian-based walks, absorbing boundaries requiring projections, walks on analog quantum machines, multi-eigenstate problems, non-unitary eigensolvers, and nonlinear dynamics. These constructs are useful beyond cognitive models, e.g., in condensed matter physics [56], or dynamics of electron systems [57], but have received scant attention from architects.

Historically, end-to-end applications enabled computer architects to discover innovation missed by simple benchmarks. For example, by offering computer architects insight into datacenter applications, CloudSuite [58] unearthed shortcomings of CPU microarchitecture missed by simpler benchmarking suites like SPEC (e.g., inefficiencies in instruction fetch). Analogously, mapping the QUATRO applications and executing them, as we will present in Sections 4 and 6, exposed many limitations of the quantum computing stack today inspiring new research. We had to innovate new state-detection circuits, nonlinear dynamics solvers, and multi-stage translators to map these applications, and discovered architecture-algorithm co-design opportunities like data-parallel RBM training, and parallel eigenstate computation to tap new axes of speedup. These are broadly applicable and advance quantum architecture.

Finally, QUATRO presents current cognitive models, factored for implementation on quantum platforms; enabling these to run on quantum computers promises substantial progress in scaling them to address more complex tasks than currently possible. It could help identify how a quantum formulation could provide a more accurate and/or coherent account of human cognition, such as the stochasticity of decision-making, and the effects of cognitive arousal, than has been achieved classically [18, 14]. It could also help answer how the brain solves seemingly computationally hard tasks, and whether mental events obey the same probability rules as physical events [18].

Table 1: The QUATRO applications: Selecting cognitive models for quantum implementation.
Model Type Purpose Significance Quantum methods we apply Relevance for systems

Quantum Walk [59, 60]

Quantum-theoretic

Decision-making (2-choice)

Accuracy, size, and flexibility advantage over classical walks

Quantum walk simulation (gate-based, annealing)

Presents Hamiltonian based walks, and with absorbing boundaries

Multi-Particle Multi-Well (MPMW) [13, 14]

Quantum-theoretic

Decision-making (multi-choice)

Accuracy, size and flexibility advantage. First unified model over many cognitive parameters

Eigensolution: SSVQE, and QITE (both gate-based)

Studies multi-eigenstate problem, non-unitary algorithms

Predator-Prey [61]

Classical, hard

Cognitive control

Captures attention, cognitive resource utilization and planning

Restricted Boltzmann Machine (RBM; annealing)

Introduces computationally hard simultaneous constraint satisfaction, and brain-inspired neural networks

Leaky Competing Accumulator (LCA) [44, 62]

Classical, hard

Decision-making and control

Biologically plausible model

Nonlinear dynamics (annealing)

Introduces simultaneous constraint satisfaction with nonlinear dynamics

Refer to caption
(a) Quantum walk on a 1-dimensional lattice for two alternatives. The walk is described by Hamiltonian, H𝐻Hitalic_H.
Refer to caption
(b) MPMW for one infinitely tall, finite width well. Its lowest three eigenvectors are shown.
Refer to caption
(c) Predator-Prey screen grid to model cognitive control/attention.
Refer to caption
(d) LCA for two inputs. f(.)1{}_{1}(.)start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT ( . ), f(.)2{}_{2}(.)start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT ( . ) are nonlinear functions.
Figure 1: Cognitive model prototypes for which we develop quantum implementations.

3 Relevant Quantum Techniques

In this section, we briefly review important quantum computing concepts relevant to the development of QUATRO. More details can be be found in standard texts (e.g., [63, 64]).

We use the Dirac (bra-ket) notation [64] (¯h=1¯1\mathchar 22\mkern-7.0muh=1¯ italic_h = 1), where |xket𝑥\ket{x}| start_ARG italic_x end_ARG ⟩ denotes a column vector x𝑥xitalic_x used to represent states, x|bra𝑥\bra{x}⟨ start_ARG italic_x end_ARG | is a row vector, and block letters (e.g., H𝐻Hitalic_H) are operators. A quantum system is described by its Hamiltonian, H𝐻Hitalic_H. If the system’s initial state is |ψinitketsubscript𝜓𝑖𝑛𝑖𝑡\ket{\psi_{init}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_i italic_n italic_i italic_t end_POSTSUBSCRIPT end_ARG ⟩, then its state after time t𝑡titalic_t is, |ψnew=U|ψinitketsubscript𝜓𝑛𝑒𝑤𝑈ketsubscript𝜓𝑖𝑛𝑖𝑡\ket{\psi_{new}}=U\ket{\psi_{init}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_n italic_e italic_w end_POSTSUBSCRIPT end_ARG ⟩ = italic_U | start_ARG italic_ψ start_POSTSUBSCRIPT italic_i italic_n italic_i italic_t end_POSTSUBSCRIPT end_ARG ⟩, where U=eiHt𝑈superscript𝑒𝑖𝐻𝑡U=e^{-iHt}italic_U = italic_e start_POSTSUPERSCRIPT - italic_i italic_H italic_t end_POSTSUPERSCRIPT is the evolution operator.

Quantum walks: Quantum walks, which describe the evolution of a particle over some states [65, 66], are used in cognitive models for decision-making [53, 52]. They capture how individuals process inputs (stimuli) to navigate their internal preferences (states) under uncertainty to make decisions.

Quantum walks are typically implemented on gate-based computers (also called digital quantum computers) by compiling U𝑈Uitalic_U into gates [32, 67] and initializing the system to |ψinitketsubscript𝜓𝑖𝑛𝑖𝑡\ket{\psi_{init}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_i italic_n italic_i italic_t end_POSTSUBSCRIPT end_ARG ⟩. The term U𝑈Uitalic_U is usually directly specified. Many specifications are possible, with coin-walks on graphs [66] being a common approach. Cognitive models require, however, a different approach, with walks specified by H𝐻Hitalic_H [53, 52], and their execution on gate-based systems requires exponentiation to derive U𝑈Uitalic_U.

Variational algorithms: Using classical and quantum computation for optimization, variational algorithms are important near-term constructs to find the eigenstates of cognitive models.

Variational quantum eigensolvers (VQEs) [68, 69] and the Quantum Approximate Optimziation Algorithm (QAOA) [70] are two important variational algorithms that minimize cost functions, thereby finding the lowest eigenstate (or ground state). Cost functions can be the expectation of the Hamiltonian or custom-defined.

SSVQE (Subspace VQE) is a recently proposed technique to identify eigenstates higher than the ground state [71]. These higher states are required by quantum cognitive models [13, 14], unlike most other applications that only need the ground state.

Similar to VQE, SSVQE uses a parameterized quantum circuit (or ansatz) to generate candidate solutions for the optimization. Then, a classical optimizer evaluates the cost function for these states, and changes the circuit parameters to explore better states. However, SSVQE differs from VQE in computing the cost function since it must find excited states.

We use two relevant SSVQE variants [71]. One approach (SSVQE B) finds one excited state at a time, e.g., the kthth{}^{\mathrm{th}}start_FLOATSUPERSCRIPT roman_th end_FLOATSUPERSCRIPT state using a weighted cost function. At each iteration, the algorithm inputs k+1𝑘1k+1italic_k + 1 orthogonal vectors, {|ψj}j=0ksuperscriptsubscriptketsubscript𝜓𝑗𝑗0𝑘\{\ket{\psi_{j}}\}_{j=0}^{k}{ | start_ARG italic_ψ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ } start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT, into the ansatz (e.g., k+1𝑘1k+1italic_k + 1 vectors whose elements are all zeros except for a 1 in a different position each), and obtains states {|ϕj}j=0ksuperscriptsubscriptketsubscriptitalic-ϕ𝑗𝑗0𝑘\{\ket{\phi_{j}}\}_{j=0}^{k}{ | start_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩ } start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT, and their energy expectations ϕj|H|ϕjbrasubscriptitalic-ϕ𝑗𝐻ketsubscriptitalic-ϕ𝑗\bra{\phi_{j}}H\ket{\phi_{j}}⟨ start_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG | italic_H | start_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩. Then, a classical optimizer minimizes the cost function with weight w𝑤witalic_w, wϕk|H|ϕk+j=0k1ϕj|H|ϕj𝑤brasubscriptitalic-ϕ𝑘𝐻ketsubscriptitalic-ϕ𝑘superscriptsubscript𝑗0𝑘1brasubscriptitalic-ϕ𝑗𝐻ketsubscriptitalic-ϕ𝑗w\bra{\phi_{k}}H\ket{\phi_{k}}+\sum_{j=0}^{k-1}\bra{\phi_{j}}H\ket{\phi_{j}}italic_w ⟨ start_ARG italic_ϕ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG | italic_H | start_ARG italic_ϕ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG ⟩ + ∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT ⟨ start_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG | italic_H | start_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩, to obtain the kthth{}^{\mathrm{th}}start_FLOATSUPERSCRIPT roman_th end_FLOATSUPERSCRIPT eigenstate.

The other method (SSVQE C) also uses a weighted cost function but finds all eigenstates up to the kthsuperscript𝑘thk^{\mathrm{th}}italic_k start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT state simultaneously. The cost function is j=0kwjϕj|H|ϕjsuperscriptsubscript𝑗0𝑘subscript𝑤𝑗brasubscriptitalic-ϕ𝑗𝐻ketsubscriptitalic-ϕ𝑗\sum_{j=0}^{k}w_{j}\bra{\phi_{j}}H\ket{\phi_{j}}∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟨ start_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG | italic_H | start_ARG italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ⟩, where the weights wjsubscript𝑤𝑗w_{j}italic_w start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT decrease in value i.e., wj<wj1subscript𝑤𝑗subscript𝑤𝑗1w_{j}<w_{j-1}italic_w start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT < italic_w start_POSTSUBSCRIPT italic_j - 1 end_POSTSUBSCRIPT. As we will show, the two methods have different implementation tradeoffs.

Quantum imaginary time evolution (QITE): Finding eigenvalues of cognitive models with variational algorithms requires identifying a good ansatz and optimizer configuration. Unfortunately, this is hard [72]. Therefore, we also consider an alternative approach, QITE, to obtain eigenvalues [73].

QITE works by evolving a quantum state with the time parameter as t=iτ𝑡𝑖𝜏t=-i\tauitalic_t = - italic_i italic_τ. Then, the state is given by eHτ|ψinitsuperscript𝑒𝐻𝜏ketsubscript𝜓𝑖𝑛𝑖𝑡e^{-H\tau}\ket{\psi_{init}}italic_e start_POSTSUPERSCRIPT - italic_H italic_τ end_POSTSUPERSCRIPT | start_ARG italic_ψ start_POSTSUBSCRIPT italic_i italic_n italic_i italic_t end_POSTSUBSCRIPT end_ARG ⟩, which represents state decay. If τ𝜏\tauitalic_τ is long, all higher energy components vanish, and the system decays into the groundstate. To obtain excited states, we use the Quantum Lanczos method that uses several QITE states with different τ𝜏\tauitalic_τ values [73].

QITE belongs to the class of non-unitary algorithms that can provide quantum speedup, but require additional steps to run on quantum computers because the hardware is unitary [74, 75, 76]. Such methods have not been studied in prior systems research.

Quantum annealing: Cognitive models using simultaneous constraint satisfaction [77, 44] and quantum walks can be cast as optimization problems solved by quantum annealing. In quantum annealing [78], the system begins in the ground state of a simple Hamiltonian, which is slowly changed to the Hamiltonian whose minimum is to be found. If the rate of change is slow enough, the system ends up in the ground state of the target Hamiltonian, giving the solution we seek [79, 80, 81, 78, 82, 83].

Quantum annealers represent analog quantum computers, which are programmed directly with the Hamiltonian [3, 84, 85, 78] unlike digital versions. This makes them attractive for cognitive models, which are commonly specified with the Hamiltonian, but such systems haven’t received adequate attention in quantum computer architecture research.

Present quantum annealers take as input Ising Hamiltonians, which are of the form H=i,jaijσziσzj+ibiσzi𝐻subscript𝑖𝑗subscript𝑎𝑖𝑗superscriptsubscript𝜎𝑧𝑖superscriptsubscript𝜎𝑧𝑗subscript𝑖subscript𝑏𝑖superscriptsubscript𝜎𝑧𝑖H=\sum_{\begin{subarray}{c}i,j\end{subarray}}a_{ij}{\sigma_{z}}^{i}\sigma_{z}^% {j}+\sum_{i}b_{i}\sigma_{z}^{i}italic_H = ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL italic_i , italic_j end_CELL end_ROW end_ARG end_POSTSUBSCRIPT italic_a 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_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT, where σzisuperscriptsubscript𝜎𝑧𝑖\sigma_{z}^{i}italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT is the Pauli Z𝑍Zitalic_Z spin operator [78] whose values can be +1 or -1, and aijsubscript𝑎𝑖𝑗a_{ij}italic_a start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT, bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are scalars and i𝑖iitalic_i, j𝑗jitalic_j refer to the qubits. General Hamiltonians, including those in cognitive models, additionally contain Pauli X𝑋Xitalic_X (σxsubscript𝜎𝑥\sigma_{x}italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT), Pauli Y𝑌Yitalic_Y (σysubscript𝜎𝑦\sigma_{y}italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT) operators. Annealing such Hamiltonians requires manual reformulation to map the applications to the hardware [86].

4 QUATRO: Quantum Cognitive Applications

We identify important cognitive models that can benefit from quantum execution, and which also introduce new constructs that have not been studied in quantum systems research. Table 1 lists these models, the probability theory they use, significance to cognitive science, the techniques we use to map them to quantum hardware, and the new features they present for systems research. We present our implementations next.

4.1 Quantum Walk

Quantum Walk captures 2-choice decision-making (e.g., yes/no tasks) using the biased random walk of a particle on a 1-dimensional lattice of states [59, 60], as shown in Figure 1a. The states represent an individual’s preference levels for the choices, and the model is given by the Hamiltonian H𝐻Hitalic_H. This model is flexible to describe many aspects of decision-making such as uncertainty, perspective; has better accuracy; and is more compact than classical walks [15, 52].

There are two variants of Quantum Walk [53], one where the walk evolves over all states including the boundary states for a fixed number of steps (called reflecting boundaries), and another where the walk evolves only until it doesn’t reach a boundary (called absorbing boundaries). The former represents decision-making responses to a prompt, and the latter captures the first time an individual arrives at a decision.

Reflecting boundaries (gate-based): The probability of a decision state (e.g., |d=|00ket𝑑ket00\ket{d}=\ket{0\ldots 0}| start_ARG italic_d end_ARG ⟩ = | start_ARG 0 … 0 end_ARG ⟩) after N𝑁Nitalic_N timesteps is given by MUN|ψinit22subscriptsuperscriptnorm𝑀superscript𝑈𝑁ketsubscript𝜓𝑖𝑛𝑖𝑡22||MU^{N}\ket{\psi_{init}}||^{2}_{2}| | italic_M italic_U start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT | start_ARG italic_ψ start_POSTSUBSCRIPT italic_i italic_n italic_i italic_t end_POSTSUBSCRIPT end_ARG ⟩ | | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, where M=|dd|𝑀ket𝑑bra𝑑M=\ket{d}\bra{d}italic_M = | start_ARG italic_d end_ARG ⟩ ⟨ start_ARG italic_d end_ARG | is the measurement operator [53]. Realizing this walk on a digital quantum computer is straightforward, using the circuit construction in Figure 2a for an example 3-qubit (8-state lattice) model.

Refer to caption

(a) Reflecting.
Refer to caption
(b) Absorbing.
Figure 2: Realizing an 8-state Quantum Walk on gate-based systems. H is the Hadamard gate.

Absorbing boundaries (gate-based): This walk evolves only in the non-boundary states, and requires projection after each timestep to contain its evolution. Thus, the probability for a decision after N𝑁Nitalic_N timesteps is given by MU(PU)N1|ψinit22subscriptsuperscriptnorm𝑀𝑈superscript𝑃𝑈𝑁1ketsubscript𝜓𝑖𝑛𝑖𝑡22||MU(PU)^{N-1}\ket{\psi_{init}}||^{2}_{2}| | italic_M italic_U ( italic_P italic_U ) start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT | start_ARG italic_ψ start_POSTSUBSCRIPT italic_i italic_n italic_i italic_t end_POSTSUBSCRIPT end_ARG ⟩ | | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, where P𝑃Pitalic_P is the projector for the non-boundary states.

Realizing this walk on existing hardware is difficult because the hardware doesn’t support partial projections. The only approach is post-selection, i.e., running the system many times and discarding the trials where the system was found in the boundaries. This requires a non-disruptive, runtime mechanism to identify when the system touches the boundaries.

Unfortunately, prior state-detection/checking circuits [87, 21, 22] are ill-suited for our purpose. We cannot use projection-based methods [87] since they change the state. Among alternatives, one approach [21] detects only certain states, such as those with an even number of ones. In our case, a boundary state can have any number of ones. The other, swap-based assertions [22], can check for approximate state membership but doubles the number of qubits to swap and restore the state.

To overcome these problems, we developed a new circuit to detect the boundary states with only one ancilla qubit. Figure 2b shows the circuit with our proposed state detector for an example 8-state walk. The ancilla qubit is shown at the bottom. For any input state i=07ai|i|0superscriptsubscript𝑖07tensor-productsubscript𝑎𝑖ket𝑖ket0\sum_{i=0}^{7}a_{i}\ket{i}\otimes\ket{0}∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | start_ARG italic_i end_ARG ⟩ ⊗ | start_ARG 0 end_ARG ⟩, the detector’s output is (a0|0+a7|7)|0+(i=16ai|i)|1tensor-productsubscript𝑎0ket0subscript𝑎7ket7ket0tensor-productsuperscriptsubscript𝑖16subscript𝑎𝑖ket𝑖ket1(a_{0}\ket{0}+a_{7}\ket{7})\otimes\ket{0}+(\sum_{i=1}^{6}a_{i}\ket{i})\otimes% \ket{1}( italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_ARG 0 end_ARG ⟩ + italic_a start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT | start_ARG 7 end_ARG ⟩ ) ⊗ | start_ARG 0 end_ARG ⟩ + ( ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | start_ARG italic_i end_ARG ⟩ ) ⊗ | start_ARG 1 end_ARG ⟩ i.e., the ancilla is in |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ if and only if the main qubits are not in the boundary states. Therefore, post-selecting on this condition is sufficient to measure the correct probabilities. The detector is easily extended to larger systems and other states by varying the CNOT and Toffoli gates. This is the first generic absorbing boundary walk design on existing quantum computers.

Annealing walk: Digital quantum computers are programmed with U𝑈Uitalic_U, which requires exponentiating the H𝐻Hitalic_H given by the cognitive model. To avoid this expensive step, we study Quantum Walk on analog computers like quantum annealers [3]. Unfortunately, existing annealers accept only Ising Hamiltonians [3], which have diagonal elements and can be decomposed into pairwise Pauli-Z terms (σzsubscript𝜎𝑧\sigma_{z}italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT). The Quantum Walk Hamiltonian is not diagonal, resulting in additional Pauli-X (σxsubscript𝜎𝑥\sigma_{x}italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT) and Pauli-Y (σysubscript𝜎𝑦\sigma_{y}italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT) terms, and can have more than two Pauli operator interactions. Therefore, we need a new approach.

We build on prior work [86] that mapped molecular Hamiltonians to annealers, to map our Quantum Walk model. The multi-stage translation we use is shown in Figure 3. We begin by decomposing the N-qubit Quantum Walk Hamiltonian into a sum of Pauli interactions. Next, we create a larger rN-qubit system and map these interactions to the larger system using new operators (e.g., X) that only use Pauli-Z operations. Finally, we convert the Pauli-Z operations in the new system to binary variables, and quadratize them i.e., expand the Hamiltonian such that each term has at most two variables [88]. The result, HBsubscript𝐻𝐵H_{B}italic_H start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT, can be annealed with existing machines, and the original ground state can be recovered from the annealer’s output. Increasing the value of r𝑟ritalic_r results in more accurate estimates.

Refer to caption
Figure 3: Realizing Quantum Walk on annealing systems.

Even though the annealing version of Quantum Walk only gives the ground state instead of full dynamics, it is still useful for cognitive scientists. Mapping this model also highlights the challenges in using Hamiltonian-based systems today.

4.2 Multi-Particle Multi-Well (MPMW)

The MPMW is a quantum cognitive model for multi-choice decision-making [13, 14]. It uses a 1-dimensional landscape of potential wells, one for each choice. The well’s parameters like its height, width, and separation from neighbors, correspond to the cognitive parameters of attention, internal representation and concept similarity, respectively. Figure 1b shows the simplest MPMW model with one infinite height well [14]. The model uses particles with different energies, representing an individual’s cognitive arousal, to describe how individuals integrate information. Each particle enters the landscape serially and its final position is measured. When a particle falls within a well, the information towards that choice is incremented by one bit. This process repeats until a definitive choice is made.

The MPMW is the first formal, unified model to integrate several cognitive parameters, and outperforms classical alternatives in accuracy, model conciseness, and expressiveness [14].

In MPMW, the particle’s position is determined by the eigenstates of the landscape’s Hamiltonian, and finding them is the key computational step. We begin by deriving the Hamiltonian from the Schrödinger equation of the particle in the landscape. In this equation, we discretize the position dimension, and use the finite difference method for the derivative to obtain the Hamiltonian operators, whose eigenstates we then find.

We identify two approaches for eigensolving: SSVQE [71], which is variational, and QITE/Quantum Lanczos [73], which is not. We select these methods because they demand different design and execution support on quantum hardware.

SSVQE: We select two variants, SSVQE B and SSVQE C (Section 3) that present different algorithmic-implementation tradeoffs. SSVQE B runs a simpler optimization to find one eigenstate, but must be run multiple times to find multiple eigenstates. Instead, SSVQE C runs a single but complex optimization to find all eigenstates. While SSVQE C is more algorithmically efficient, SSVQE B is embarrassingly parallel and each instance is faster than SSVQE C. This presents a new opportunity for speedup that could be tapped with an efficient system architecture. This axis of parallelism is also complementary to existing methods that parallelize the expectation calculation of Pauli operators within a variational step [89, 90], or those that parallelize runs for averaging [91].

QITE/Quantum Lanczos: Variational algorithms like SSVQE require selecting a suitable ansatz and optimizer, which is generally difficult [72]. Hence, we pick an alternative non-variational algorithm, QITE [73], for eigensolving. QITE is in the class of non-unitary quantum algorithms [92, 75, 76], which haven’t been studied in systems design. Being non-unitary, QITE is unsupported natively on quantum computers, which are unitary. Therefore, we follow prior work [73] to determine unitary operators that result in the same state evolution as with QITE. At each timestep, these operators are obtained by solving a linear system of equations with Pauli-operator expectations, which in turn, are measured from a quantum computer.

To find the excited states, we use the Quantum Lanczos algorithm base on QITE [73]. This approach obtains different state vectors evolved with QITE to construct a vector subspace. Then, it runs a classical Lanczos iteration on this subspace to orthogonalize the states into excited eigenstates.

4.3 Predator-Prey

Predator-Prey is used to model cognitive control [61], which governs other mental processes including decision-making. The model describes a player playing a game shown in Figure 1c. The game has a screen grid with three entities: an agent, which is the player’s screen icon, a prey, and a predator. The player must move the agent to capture the prey and avoid being caught by the predator. The model captures how the player allocates attention to the screen icons to discern their position and uses those perceived positions to make a move.

Identifying the optimal attention levels allocable to each icon, and deciding movement is a simultaneous constraint satisfaction problem. We model it with a quantum restricted Boltzmann machine (RBM) that we propose [77], shown in Figure 4. We choose an RBM because it aligns with the simultaneous interactions found in the brain, and also for its deep connection with cognitive modeling [93, 94]. Our RBM has a visible layer with nodes for the true position of the on-screen entities, attention, perceived positions, and the movement direction.

Refer to caption
Figure 4: Quantum RBM network for Predator-Prey.

We map our RBM to a quantum annealer since it is a good fit for the RBM’s Ising interactions. The standard approach to run the RBM uses 2 anneals for inference (visible to hidden, and back) and 3 for training (+1 to update hidden nodes). We additionally propose two new methods that benefit from different system architectures. One approach (RBMeffeff{}_{\textrm{eff}}start_FLOATSUBSCRIPT eff end_FLOATSUBSCRIPT) leverages compute-parallelism by combining updates to both hidden and visible layers. This method is resource efficient, requiring only 1 anneal for inference and 2 for training, but has more complexity. Another approach (RBMparKparK{}_{\textrm{parK}}start_FLOATSUBSCRIPT parK end_FLOATSUBSCRIPT) leverages data-parallelism where we anneal for K𝐾Kitalic_K different samples simultaneously. This is inspired by asynchronous ML training algorithms [95], where the network weights are updated once after K𝐾Kitalic_K samples, instead of K𝐾Kitalic_K sequential updates.

4.4 Leaky Competing Accumulator (LCA)

The LCA is a biologically inspired model for multi-choice decision-making and cognitive control [44, 62]. Figure 1d shows the LCA for two choices. Each choice has an accumulator x(t)𝑥𝑡x(t)italic_x ( italic_t ), which integrates the values of the input i(t)𝑖𝑡i(t)italic_i ( italic_t ) corresponding to that choice at timestep t𝑡titalic_t. The accumulator loses or leaks its value (by a factor λ𝜆\lambdaitalic_λ), and is inhibited by the output of the other choice (by a factor β𝛽\betaitalic_β), i.e., x1(t)=i1(t)+(1λ)x1(t1)βf2(t)subscript𝑥1𝑡subscript𝑖1𝑡1𝜆subscript𝑥1𝑡1𝛽subscript𝑓2𝑡x_{1}(t)=i_{1}(t)+(1-\lambda)\cdot x_{1}(t-1)-\beta f_{2}(t)italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) = italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) + ( 1 - italic_λ ) ⋅ italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t - 1 ) - italic_β italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ), where f()𝑓f()italic_f ( ) is a nonlinear function such as a sigmoid that acts on the accumulator’s output.

We select the LCA because it is widely used, and its mathematical behavior has been well-characterized [62]. Moreover, the LCA is a simultaneous constraint satisfaction model like Ising models, with the additional complexity of stateful dynamics or memory—occurring due to the accumulator, and nonlinearity—due to the activation function. These aspects have been crucial in its utility as a model, but they also create challenges in running it on existing machines.

We map the LCA to quantum annealers since they can solve Ising problems. We first linearize f()𝑓f()italic_f ( ) into the format ax+b𝑎𝑥𝑏ax+bitalic_a italic_x + italic_b using its Taylor expansion. Next, to obtain the value of f()𝑓f()italic_f ( ) for a given value of x𝑥xitalic_x, we use the annealing cost function, (fx(ax+b))2superscriptsubscript𝑓𝑥𝑎𝑥𝑏2(f_{x}-(ax+b))^{2}( italic_f start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT - ( italic_a italic_x + italic_b ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, following prior work that used annealing for prime factorization [96]. Minimizing this cost function would yield fxsubscript𝑓𝑥f_{x}italic_f start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT to be f(x)absent𝑓𝑥\approx f(x)≈ italic_f ( italic_x ). Then, we replace x𝑥xitalic_x with its previous value (x(t1)𝑥𝑡1x(t-1)italic_x ( italic_t - 1 )), and the input. This gives a cost function with fxsubscript𝑓𝑥f_{x}italic_f start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, x(t1)𝑥𝑡1x(t-1)italic_x ( italic_t - 1 ), and the inputs i(t)𝑖𝑡i(t)italic_i ( italic_t ). Next, we expand each variable with multiple qubits to represent floating point numbers. Finally, we quadratize the cost function before annealing.

We explored different alternatives to run the LCA over multiple timesteps. One approach is to anneal for a single timestep at once, where the outputs from the previous timestep are used as inputs for the next anneal. Since existing annealers do not offer easy variable initialization, we use additional terms in our cost function to realize it. This method has the overhead of reading and re-initializing qubits at each step.

Another approach is to unroll the LCA dynamics for K𝐾Kitalic_K steps, so that one anneal would return the output for all K𝐾Kitalic_K steps. This amortizes the reading/initialization overheads but results in a much larger Hamiltonian than what is reliably run on current systems. The last approach uses the technique of Feynman’s clock, where a new problem is formulated that includes a timestep register along with the LCA variables [97]. This method introduces complex variables that are unsupported presently. As a balance, we anneal for K𝐾Kitalic_K steps at once.

5 Experimental Setup

Refer to caption
(a) Timestep 0.
Refer to caption
(b) Timestep 1.
Refer to caption
(c) Timestep 2.
Refer to caption
(d) Timestep 3.
Refer to caption
(e) Timestep 4.
Figure 5: Reflecting boundary Quantum Walk with statevector simulation.
Refer to caption
(a) Timestep 0.
Refer to caption
(b) Timestep 1.
Refer to caption
(c) Timestep 2.
Refer to caption
(d) Timestep 3.
Refer to caption
(e) Timestep 4.
Figure 6: Reflecting boundary Quantum Walk on IBM Perth.

Platforms: We choose publicly available hardware platforms with mature software to evalaute QUATRO. For our gate-based applications, we use IBM systems (IBM Perth/Quito, which have 5 superconducting transmon qubits), and for annealing, we use D-Wave’s Advantage_system4.1 solver, which has 5, 627 qubits connected in a Pegasus graph [98].

Cognitive model parameters: QUATRO can scale to the large sizes we target, but existing hardware cannot fit them. Therefore, we run smaller versions. Table 2 shows the minimum model sizes we’d like to evaluate, the actual sizes that we could fit on hardware, and the implementation parameters.

Table 2: Model and implementation parameters.
Model parameters
Model

Ideal minimum

Actual

Implementation parameters
Quantum Walk 100 states, 1000 walks

8 states, 1 walk

Gates. Qubits: 3 (ref), 4 (abs)

4 states, 1 walk

Annealing (r𝑟ritalic_r:2–9, 60% pausing)

MPMW 10 dynamic wells, 1000 positions, 10 eigenstates 1 static well, 4 positions, 3 eigenstates

SSVQE (COBYLA/SPSA)

QITE (step size = 0.2, steps = 135)

Predator-Prey

1000×\times×1000 grid, 100 levels, 10 icons

6×\times×6 grid, 12 levels, 3 icons

RBM (112 node; data: 100 train, 100 test), 60% pausing

LCA

10 units, multilayer

2 units, 1 layer

5-step unroll, 6-bit floats, 60% pausing

A large Quantum Walk uses >100 states, and multiple walks to model interacting decisions. The smaller version we evaluate uses a single walk with 8 states for gate systems and 4 states for annealing. Annealing has fewer states since the actual number of qubits needed is much higher (\geqrN; Section 4). We pause anneals [99] to mitigate thermalization and noisy outputs.

MPMW scales to multiple wells and well-sizes, several eigenstates, and fine spatial resolution in the horizontal dimension. On actual hardware, we find 3 eigenstates for the simplest MPMW model, which has one infinite-height well. In finding eigenvalues with SSVQE, we use the SPSA optimizer [67] in simulations as it is more accurate, but use COBYLA [67] on actual hardware since it requires fewer executions.

Predator-Prey could scale to large screen grids, many attention levels, and icons. For the smaller version we run, our RBM has 56 visible and 56 hidden nodes. The nodes take binary values, but their biases and the weights for the edges can have real values. Since we do not need the perceived positions explicitly, we use a single set of nodes for the true and perceived positions, encoded with one-hot vectors.

The RBM is trained with contrastive divergence [100]. We obtain training data by identifying the best decision from an exhaustive search without lookahead. We train on 100 randomly chosen screens over 30 epochs, and use another 100 as test. While this data may seem small, evaluating it consumed hours of anneals and thousands of dollars. We use one-hot encoding to represent the screen grid coordinates, and allow the agent to move one step in any of the 8 cardinal and ordinal directions.

A large LCA model would have multiple layers, each with several LCA units to model cognitive control and decision-making. In our experiments, we use a single layer with 2 units. We use 6 qubits to represent float values. Since the LCA has multiple outputs over multiple timesteps, we quantify accuracy using the mean local relative error, which is the average error of a method f¯¯𝑓\overline{f}over¯ start_ARG italic_f end_ARG relative to exact LCA (f𝑓fitalic_f), when they run on the same inputs i𝑖iitalic_i and previous values (x1¯¯subscript𝑥1\overline{x_{-1}}over¯ start_ARG italic_x start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG) from the method. It is given by, avg(|f(i,x1)f(i,x1¯)f(i,x1)|)avg(|\frac{f(i,{x_{-1}})-\overline{f(i,{x_{-1}}})}{f(i,{x_{-1}})}|)italic_a italic_v italic_g ( | divide start_ARG italic_f ( italic_i , italic_x start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ) - over¯ start_ARG italic_f ( italic_i , italic_x start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG ) end_ARG start_ARG italic_f ( italic_i , italic_x start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ) end_ARG | ).

6 Evaluation

6.1 Quantum Walk

Gate-based implementation: Figures 5 and 6 show the probabilities of the states at various timesteps in the reflecting boundary Quantum Walk using statevector simulations, and as measured from quantum hardware (IBM Perth), respectively. The simulations align closely with our analytical calculations, but the results from quantum hardware deviate significantly, especially from timestep 2. The absorbing boundaries walk, shown in Figures 10 (simulations) and 11 (measured) is worse affected. Much of the distribution is lost even by timestep 2.

The cause of poor hardware performance is noise from running large circuits. Figure 7 shows the gate count, depth, and execution time of the walks. While these models use only 3 or 4 qubits, the gate depths are in the hundreds, resulting in large noise. We also simulate the absorbing boundary walk with 1- and 2-qubit gate depolarizing noise to study its sensitivity. Figure 8 shows the results at timestep 4 for different noise probabilities. There is noticeable distortion (vs Figure 10e) at p𝑝pitalic_p=0.001, suggesting significant need for fault tolerance [101], but near-accurate result for p𝑝pitalic_p=0.0001, indicating promise.

Noise mitigation is vital for Quantum Walk models, even more than building qubit capacity.

Refer to caption
(a) Gate count.
Refer to caption
(b) Circuit depth.
Refer to caption
(c) Execution time.
Figure 7: Gate count, depth, and performance of Quantum Walk.
Refer to caption
(a) p=0.01𝑝0.01p=0.01italic_p = 0.01
Refer to caption
(b) p=0.001𝑝0.001p=0.001italic_p = 0.001
Refer to caption
(c) p=0.0001𝑝0.0001p=0.0001italic_p = 0.0001
Figure 8: Timestep 4 of absorbing walk with depolarizing noise.

Annealing implementation: Figure 9 shows the results for finding the ground state of a 4-state Quantum Walk using annealing. We show the results for both simulated annealing (SA) and quantum annealing with pausing (QA-P). Figure 9a shows the absolute error of the measured eigenvalue (the exact value is -7.22). We expect the error to decrease as the qubit repetition (r𝑟ritalic_r) is increased, and we find this to be true for SA. With QA-P, however, the error increases after an initial decrease. This is because bigger r𝑟ritalic_r values result in more qubits (Figure 9b) and complex interaction, worsening noise.

Refer to caption
(a) Accuracy.
Refer to caption
(b) Size.
Refer to caption
(c) Execution time.
Figure 9: Annealing for a 4-state Quantum Walk.

Figure 9c shows the execution time of SA and QA-P. SA’s execution time grows linearly with r𝑟ritalic_r while QA-P is relatively flat. Even though the error for QA-P is higher at large r𝑟ritalic_r, the trend for small r𝑟ritalic_r (where its accuracy is comparable to SA) suggests that QA’s execution time can scale better.

QA suffers from noise with large Quantum Walk models. With small models, QA’s performance scales better over SA. There’s also an accuracy-time tradeoff to explore in annealers.

6.2 MPMW

Refer to caption
(a) Timestep 0.
Refer to caption
(b) Timestep 1.
Refer to caption
(c) Timestep 2.
Refer to caption
(d) Timestep 3.
Refer to caption
(e) Timestep 4.
Figure 10: Absorbing boundary Quantum Walk with statevector simulation (Total probability <<<1 due to projection).
Refer to caption
(a) Timestep 0.
Refer to caption
(b) Timestep 1.
Refer to caption
(c) Timestep 2.
Refer to caption
(d) Timestep 3.
Refer to caption
(e) Timestep 4.
Figure 11: Absorbing boundary Quantum Walk on IBM Perth (Total probability <<<1 due to projection).

SSVQE: Identifying a suitable ansatz for SSVQE was nontrivial. We simulated 24 choices from prior work [72, 69, 102] to solve for MPMW’s ground state using standard VQE, and Figure 12 shows their performance. The best choice was (Ansatz 1), which is “Circuit 14” from [72] that uses RYsubscript𝑅𝑌R_{Y}italic_R start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT and controlled RXsubscript𝑅𝑋R_{X}italic_R start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT gates. We also used circuit search methods from quantum ML [103], but they did not perform competitively.

Refer to caption
Figure 12: Ansatz selection with VQE (the exact value is shown).

For optimizers, we evaluated both SPSA and COBYLA in our simulations, and found SPSA to be better. However, SPSA required 2–10×\times× more circuit evaluations, which we could not run on real hardware in reasonable time for convergence. So, we only used COBYLA for quantum hardware.

We run the SSVQE B and SSVQE C algorithms to find the 0thth{}^{\textrm{th}}start_FLOATSUPERSCRIPT th end_FLOATSUPERSCRIPT, 1stst{}^{\textrm{st}}start_FLOATSUPERSCRIPT st end_FLOATSUPERSCRIPT and 2ndnd{}^{\textrm{nd}}start_FLOATSUPERSCRIPT nd end_FLOATSUPERSCRIPT eigenstates of MPMW, with exact values as 1.72, 6.21, and 11.78, respectively. Figure 13 shows the progression of SSVQE B2, which only finds the 2nd𝑛𝑑{}^{nd}start_FLOATSUPERSCRIPT italic_n italic_d end_FLOATSUPERSCRIPT eigenstate. The measured value (10.1) has a steady state error (14%). Other runs (SSVQE B0/B1/C) are similar.

Refer to caption
(a) Simulated.
Refer to caption
(b) Measured.
Figure 13: Simulated and measured SSVQE B2 for MPMW.

Figure 14 shows the execution time of SSVQE. SSVQE uses many circuit evaluations to compute the Pauli expectations, repeated for each eigenstate, causing long execution times. The evaluations increase with the SPSA optimizer.

Refer to caption
(a) SPSA.
Refer to caption
(b) COBYLA.
Refer to caption
(c) COBYLA (IBM Quito).
Figure 14: Execution time of the SSVQE algorithm for MPMW

Figure 14 also shows that each SSVQE B instance takes lesser time than SSVQE C, which finds all eigenvalues at once. Thus, if these multiple instances could be run in parallel, the time to solution will be faster. Unfortunately, existing quantum clouds do not support such execution.

Little support for ansatz and optimizer selection. There is embarassing parallelism to be exploited in SSVQE.

QITE/Quantum Lanczos: Figure 15 shows their performance for MPMW. Figure 15a shows convergence of the ground state but also significant hardware noise. Figure 15b shows the error in the calculated eigenvalues. The ground state error is low and is better than what SSVQE achieved. The excited states have a steady state error due to our unitary approximation, but the error is only slightly worse than with SSVQE. Importantly, however, QITE is much faster than SSVQE, despite being nonunitary. It required just 164.8 s in simulations (vs 12,000 s for SSVQE C with SPSA for comparable accuracy). This is because it has fewer evaluations, and doesn’t tune an ansatz. Moreover, QITE/Quantum Lanczos do not require an ansatz and optimizer search, which is hard. Unfortunately, these methods haven’t received adequate attention in systems.

Refer to caption
(a) Convergence.
Refer to caption
(b) Accuracy of eigenstates.
Figure 15: QITE and Quantum Lanczos for MPMW.

6.3 Predator-Prey

Figure 16 shows a few sequences of moves generated by our RBM with the three different annealing strategies (RBM, RBMeffeff{}_{\textrm{eff}}start_FLOATSUBSCRIPT eff end_FLOATSUBSCRIPT, and RBMpar2par2{}_{\textrm{par2}}start_FLOATSUBSCRIPT par2 end_FLOATSUBSCRIPT), and the corresponding movements from the predator and prey. At each step, we show the deviation of the RBM’s move from a no-lookahead exhaustive search.

Refer to caption
(a) Multiple anneals (RBM).
Refer to caption
(b) Sample-efficient anneal (RBMeffeff{}_{\textrm{eff}}start_FLOATSUBSCRIPT eff end_FLOATSUBSCRIPT).
Refer to caption
(c) Two-sample parallel anneal (RBMpar2par2{}_{\textrm{par2}}start_FLOATSUBSCRIPT par2 end_FLOATSUBSCRIPT).
Figure 16: Predator-Prey screen with different RBM strategies.

Figure 17a shows the execution time of our RBM implementations. Compared to RBM, RBMeffeff{}_{\textrm{eff}}start_FLOATSUBSCRIPT eff end_FLOATSUBSCRIPT takes nearly 100×\times× longer. This is because RBMeffeff{}_{\textrm{eff}}start_FLOATSUBSCRIPT eff end_FLOATSUBSCRIPT packs 2 passes into one anneal resulting in significantly complex interactions, and thus, noise. The number of physical qubits to realize this network is 100×\times× higher. On the other hand, RBMpar2par2{}_{\textrm{par2}}start_FLOATSUBSCRIPT par2 end_FLOATSUBSCRIPT is \approx2×\times× faster. Even though it packs the processing of 2 samples into one anneal, the two samples are processed independently and do not complicate qubit interactions. Thus, it only needs 2×\times× more qubits.

Refer to caption
(a) Timing.
Refer to caption
(b) Accuracy.
Figure 17: Predator-Prey performance.

Figure 17b shows the accuracy of the RBMs with simulated (SA) and quantum annealing (QA). RBMeffeff{}_{\textrm{eff}}start_FLOATSUBSCRIPT eff end_FLOATSUBSCRIPT has poor accuracy due to its complex interactions, which is more pronounced with noise-sensitive QA. On the other hand, RBMpar2par2{}_{\textrm{par2}}start_FLOATSUBSCRIPT par2 end_FLOATSUBSCRIPT has comparable accuracy to RBM. Note that in training, RBMpar2par2{}_{\textrm{par2}}start_FLOATSUBSCRIPT par2 end_FLOATSUBSCRIPT updates the weights after 2 data samples, deviating from the standard algorithm that updates weights after each sample. However, this doesn’t affect the RBM’s accuracy, similar to what has been observed in some classical ML training algorithms [95]. We, thus, find a new opportunity to accelerate quantum RBMs. But, existing quantum clouds don’t support such parallel execution, and we had to emulate it with 0-strength qubit couplings.

6.4 LCA

Figure 18 shows the two outputs of LCA (f1subscript𝑓1f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and f2subscript𝑓2f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) for 10 timesteps, computed with different methods: SA, QA-P, and QA-NP (QA without pausing). We find 5 timesteps in one anneal, requiring 60 logical qubits and 252 physical qubits.

LCA with SA is close to the exact values, with a mean local error (Section 5) of (0.58%, 0.49%) for the two LCA outputs. Among the QA methods, QA-NP has larger deviations due to noise, with an error of (5.75%, 7.54%). QA-P has a better fit with an error of (5.18%, 5.82%). Note that our error metric does not consider the conditioning of the problem i.e., for ill-conditioned calculations, errors are amplified over time.

Refer to caption
Figure 18: Quantum annealing for the LCA.

For execution time, SA took 84.26 ms for one anneal with 100 reads and 1000 sweeps. For the best fidelity, QA required 175 ms QPU (quantum processing unit) time for 3500 reads with 50 μ𝜇\muitalic_μs anneals each. With simple noise mitigation (i.e., pausing) QA for the LCA model achieved comparable accuracy to SA. However, QA also needed 3.5×\times× more reads, and consequently, 2×\times× longer time. Noise mitigation and device reliability could bridge this gap in the near term.

Exiting annealers are noisy to run LCA, but mitigation can offer superior performance in the near-term.

7 Case Study: Cloud Co-design

We demonstrate research that QUATRO inspires through the design of efficient quantum clouds that support parallelism.

Existing quantum clouds do not leverage the parallelism found in QUATRO, e.g., in eigensolvers, or RBMs. Furthermore, we persistently encountered long scheduling delays in our evaluation. These occur due to an inefficient cloud. Figure 19a shows the scheduling on quantum clouds today [104, 3]. Users place each job into device-specific queues, where they stay without migration until execution. This inflexibility combined with the ad hoc job execution heuristics [105] results in load imbalance across devices, and long wait times for users. Iterative jobs like SSVQE or QITE additionally suffer a long network round trip between quantum and classical devices for every iteration. The present design also fails to benefit from parallelism because when jobs like SSVQE spawn multiple parallel instances, they are all run serially on the same device.

Refer to caption
(a) Existing.
Refer to caption
(b) Proposed.
Figure 19: Reorganizing quantum cloud scheduling.

We envision a new cloud that (i) separates user and device scheduling, (ii) runs hybrid, and not just quantum jobs, and (iii) allows autonomous hybrid job execution with cloudlets [106]. Figure 19b shows our design. We distinguish user-facing front-end scheduling from back-end device-specific job selection. Users submit jobs to the front-end without selecting a specific device. Instead, they are mapped to any available device (considering preferences and priorities), reducing imbalance and wait times. Next, to reduce long-latency network round trips, we propose adding cloudlets [106] with classical compute (CPUs/GPUs). Cloudlets are small, geographically close servers, that have gained prominence in mobile network systems. Users submit both quantum and hybrid tasks onto this system, where the hybrid jobs run autonomously on respective devices.

Existing systems lack the features (i)-(iii) above. Amazon [1] allows hybrid job scheduling, but users must select a single device. The hybrid jobs cannot also create additional instances for other devices, which is needed for dynamic parallelism.

Our envisioned cloud is more aligned with the needs of the QUATRO applications. It enables exploiting the embarrassing parallelism that we find in eigensolvers and RBMs, without any hardware change or synchronization, complementing other forms of parallelism explored in prior work [89, 90, 91, 107].

Figure 20a shows the normalized speedup for MPMW that can be obtained by executing the 3 instances of SSVQE B in parallel, or one SSVQE C instance (algorithmic multitasking/parallelism) vs serially running SSVQE B (current option), even without cloudlets. Parallel execution yields \approx2.4×\times× speedup with no design effort. When other forms of parallelism are combined, the improvements would be multiplicative.

Refer to caption
(a) MPMW SSVQE.
Refer to caption
(b) Parallel Predator-Prey (time).
Refer to caption
(c) Parallel Predator-Prey (accuracy).
Figure 20: Impact of a parallelism-aware quantum cloud.

For the RBM in the Predator-Prey model, Figure 20b shows the training and testing times, and Figure 20c shows the accuracy when processing varying number of samples in parallel on a system that permits such execution. We also show a design that is trained in serial, but tested in parallel (Tmixedmixed{}_{\textrm{mixed}}start_FLOATSUBSCRIPT mixed end_FLOATSUBSCRIPT). The RBM’s accuracy remains relatively unchanged up to 10-parallel evaluations during training, but is 10×\times× faster than serial execution. Furthermore, testing can always run in parallel without loss of accuracy. These results show the significant benefits that a new cloud inspired from QUATRO can provide with no change in hardware or toolchains.

8 Lessons Learnt

The design and execution of QUATRO applications required innovation, and has also revealed new research. Table 3 organizes these findings. In addition to the discussion on organization in section 7, we briefly discuss the rest here.

Table 3: Innovation and limitations exposed.
Algorithms Compilers & Tools Organization Hardware
Exposed limitations

Few algorithms for analog quantum computers, nonlinear dynamics

Limited support for projections, analog quantum computers, non-unitary algorithms, ansatz search

Poor scheduling, lack of support for parallelism

Limited capability, noise

Innovation & discovery

State-detector, Data-parallel RBMs, parallelism in eigensolvers

Hamiltonian walks on annealers

Cloud co-design for parallelism

Software (Algorithms, Compilers): Present quantum software frameworks are heavily specialized (e.g., [108, 109, 110, 111]). We explored them for QUATRO but the abstractions mismatch. The toolchains are also heavily focused on specific computation-styles (unitary or Ising Hamiltonians), devices (gate-based or annealers) and algorithms (variational). There is little tool support for competitive alternatives like Hamiltonian-based quantum systems [85, 84], or non-unitary algorithms. Even popular constructs like ansatz search, or partial projectors lack generalizable tools. This reduces programmability and restricts the applications that can be run on quantum systems today.

We also identify new research into the asynchronous training of RBMs. Understanding the theoretical mechanisms of this approach, and its applicability for other quantum ML networks would be valuable, as it did in classical computing [95].

Lastly, the new cloud design we envision also calls for new runtime managers and quantum operating systems to provide performance and protection.

Hardware-Software interfaces: Existing hardware exposes device control parameters such as the pulse information required to implement gates [112], or the annealing schedules for annealers. While this enables cross-layer co-design [113], it also creates additional programming burden to identify the correct settings—e.g., anneal schedules for pausing in the LCA. This is especially hard with the limited tools available today.

Striking a balance, we propose that device control mechanisms be abstracted with compiler heuristics. For example, pausing can be automated by using recent work that can generate a quick estimate [114].

Hardware: While current NISQ (noisy intermediate-scale quantum) devices may not outperform the best classical systems, we show promising empirical results that the QUATRO models could do so in the near future. We highlight (in Section 6.1) that, with improved noise rate (at 104superscript10410^{-4}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT), we can already start to explore quantum walk models accurately, representing meaningful NISQ or early fault tolerant applications.

We also argue for research increasing device functionality, complementing the present focus on noise resilience and qubit counts. Lack of functionality, e.g., the absence of primitives to express projections in Quantum Walk, requires users to design custom circuits and pre-/post-selection methods. This is tedious, error-prone since quantum programming is not intuitive [115], and inefficient since it wastes qubits and execution time. While capabilities towards these features exist [116], hardware development inevitably takes long, and such features could at least be software emulated in the interim.

9 Related Work

Quantum applications: Numerous advances [117, 118, 119, 120, 121, 122] have helped make quantum computing a reality. However, even though there is great interest in adopting quantum computing for new domains (e.g., [20, 123]), realizing new applications has been a challenge [42, 6]. Applications are focused in only a few areas [124, 37, 39, 40, 41]. We identify cognitive modeling as a new domain that can benefit from quantum computing, and present a suite of real-world applications. These are full models unlike benchmark suites, which use kernels for performance comparisons [55, 125, 126, 20]. Our work also spans analog and digital quantum computers, which is rare.

Quantum computer architecture: Significant strides have been made in quantum programming and compilation [32], circuit synthesis [127, 128], noise mitigation and reliability [26, 27, 28, 29, 30, 31], and microarchitecture design [34]. Recent work [129] on exploring improving fidelity of partial measurements, and improving variational algorithms for near-term machines [130, 131] can help the models we study.

One limitation of prior research is that it is heavily device-centric [26, 27, 28, 29, 30, 31, 32, 33, 34], with only a few analyses at the system level [132, 133, 134]. Our work on higher level organization complements existing research. Furthermore, existing systems research and software frameworks are almost entirely focused on gate-based systems, which are only one type of quantum computers. Many applications, like some of ours, are more naturally suited to Hamiltonian-based computers (e.g., QuEra [84, 3]). Unfortunately, the stack for these types of systems is limited.

Ravi et al. [132, 133] raised the issue of growing application demand in the cloud, and proposed fidelity- and queuing-aware scheduling. We consider other aspects like parallelism.

Parallelism and Scheduling: Prior work explored microarchitecture and circuit-level parallelism with co-location on gate-based computers [135, 136, 137, 138, 139] and annealers [140, 134] while mitigating reliability issues that arise. We study a complementary form of embarassing parallelism exposed by QUATRO, that can benefit from, but does not need co-location. Capturing all these forms of parallelism could enable near-term machines to be competitive for real applications.

10 Conclusion

Quantum computing can benefit many domains, but it has been a challenge to identify new applications, and stimulate architecture-application co-design. This work presented cognitive modeling as a new application area for quantum computing. We developed QUATRO, a suite of real-world cognitive models that can be run on existing quantum hardware. Developing QUATRO required innovation, and running them helped us identify new research in the quantum stack, some of which we evaluate with real data. Our work simultaneously advances the cognitive sciences, and quantum computer architecture.

References

  • [1] “Amazon Braket.” [Online]. Available: https://aws.amazon.com/braket/
  • [2] “Azure Quantum cloud service.” [Online]. Available: https://azure.microsoft.com/en-us/products/quantum
  • [3] “D-Wave Leap.” [Online]. Available: https://www.dwavesys.com/solutions-and-products/cloud-platform/
  • [4] “IBM Quantum.” [Online]. Available: https://quantum-computing.ibm.com/
  • [5] “Xanadu Cloud.” [Online]. Available: https://platform.xanadu.ai/
  • [6] Y. Alexeev, D. Bacon, K. R. Brown, R. Calderbank, L. D. Carr, F. T. Chong, B. DeMarco, D. Englund, E. Farhi, B. Fefferman, A. V. Gorshkov, A. Houck, J. Kim, S. Kimmel, M. Lange, S. Lloyd, M. D. Lukin, D. Maslov, P. Maunz, C. Monroe, J. Preskill, M. Roetteler, M. J. Savage, and J. Thompson, “Quantum computer systems for scientific discovery,” PRX Quantum, vol. 2, p. 017001, Feb 2021. [Online]. Available: https://link.aps.org/doi/10.1103/PRXQuantum.2.017001
  • [7] D. Kumaran, D. Hassabis, and J. L. McClelland, “What learning systems do intelligent agents need? complementary learning systems theory updated,” Trends in cognitive sciences, vol. 20, no. 7, pp. 512–534, 2016.
  • [8] E. M. Pothos and J. R. Busemeyer, “Quantum cognition,” Annual review of psychology, vol. 73, pp. 749–778, 2022.
  • [9] D. Kahneman, “Maps of bounded rationality: Psychology for behavioral economics,” American economic review, vol. 93, no. 5, pp. 1449–1475, 2003.
  • [10] J. Veselỳ, R. P. Pothukuchi, K. Joshi, S. Gupta, J. D. Cohen, and A. Bhattacharjee, “Distill: Domain-specific compilation for cognitive models,” in 2022 IEEE/ACM International Symposium on Code Generation and Optimization (CGO).   IEEE, 2022, pp. 301–312.
  • [11] K. J. Holyoak, “Analogy and relational reasoning,” in The Oxford handbook of thinking and reasoning, K. J. Holyoak and R. G. Morrison, Eds.   Oxford University Press, 2012, pp. 234–259.
  • [12] M. K. Ho, D. Abel, J. D. Cohen, M. L. Littman, and T. L. Griffiths, “The efficiency of human cognition reflects planned information processing,” in AAAI Conference on Artificial Intelligence, Apr. 2020.
  • [13] L. Rosendahl, A. S. Bizyaeva, and J. Cohen, “A novel quantum approach to the dynamics of decision making.” in CogSci, 2020.
  • [14] M. L. Rosendahl, “The multi-particle multi-well (mpmw) framework: A quantum framework incorporating attentional capture, representational generality, and arousal to perceptual choice problems,” Ph.D. dissertation, Princeton University, 2022.
  • [15] J. R. Busemeyer, Z. Wang, and R. M. Shiffrin, “Bayesian model comparison favors quantum over standard decision theory account of dynamic inconsistency.” Decision, vol. 2, no. 1, p. 1, 2015.
  • [16] J. R. Busemeyer and P. D. Bruza, Quantum models of cognition and decision.   Cambridge University Press, 2012.
  • [17] E. M. Pothos and J. R. Busemeyer, “Quantum principles in psychology: The debate, the evidence, and the future,” Behavioral and Brain Sciences, vol. 36, no. 3, pp. 310–327, 2013.
  • [18] P. D. Bruza, Z. Wang, and J. R. Busemeyer, “Quantum cognition: a new theoretical approach to psychology,” Trends in cognitive sciences, vol. 19, no. 7, pp. 383–393, 2015.
  • [19] G. P. Epping and J. R. Busemeyer, “Using diverging predictions from classical and quantum models to dissociate between categorization systems,” Journal of Mathematical Psychology, vol. 112, p. 102738, 2023.
  • [20] D. Widdows, J. Rani, and E. M. Pothos, “Quantum circuit components for cognitive decision-making,” Entropy, vol. 25, no. 4, 2023. [Online]. Available: https://www.mdpi.com/1099-4300/25/4/548
  • [21] J. Liu, G. T. Byrd, and H. Zhou, “Quantum circuits for dynamic runtime assertions in quantum computation,” in Proceedings of the Twenty-Fifth International Conference on Architectural Support for Programming Languages and Operating Systems, ser. ASPLOS ’20.   New York, NY, USA: Association for Computing Machinery, 2020, p. 1017–1030. [Online]. Available: https://doi.org/10.1145/3373376.3378488
  • [22] J. Liu and H. Zhou, “Systematic approaches for precise and approximate quantum state runtime assertion,” in 2021 IEEE International Symposium on High-Performance Computer Architecture (HPCA), 2021, pp. 179–193.
  • [23] K. Temme, S. Bravyi, and J. M. Gambetta, “Error Mitigation for Short-Depth Quantum Circuits,” Phys. Rev. Lett., vol. 119, no. 18, p. 180509, Nov. 2017.
  • [24] S. Endo, S. C. Benjamin, and Y. Li, “Practical Quantum Error Mitigation for Near-Future Applications,” Phys. Rev. X, vol. 8, no. 3, p. 031027, Jul. 2018.
  • [25] Z. Cai, R. Babbush, S. C. Benjamin, S. Endo, W. J. Huggins, Y. Li, J. R. McClean, and T. E. O’Brien, “Quantum Error Mitigation,” arXiv, Oct. 2022.
  • [26] T. Patel and D. Tiwari, “Veritas: Accurately estimating the correct output on noisy intermediate-scale quantum computers,” in SC20: International Conference for High Performance Computing, Networking, Storage and Analysis, 2020, pp. 1–16.
  • [27] ——, “Qraft: Reverse your quantum circuit and know the correct program output,” in Proceedings of the 26th ACM International Conference on Architectural Support for Programming Languages and Operating Systems, ser. ASPLOS ’21.   New York, NY, USA: Association for Computing Machinery, 2021, p. 443–455. [Online]. Available: https://doi.org/10.1145/3445814.3446743
  • [28] S. S. Tannu and M. K. Qureshi, “Mitigating measurement errors in quantum computers by exploiting state-dependent bias,” in Proceedings of the 52nd Annual IEEE/ACM International Symposium on Microarchitecture, ser. MICRO ’52.   New York, NY, USA: Association for Computing Machinery, 2019, p. 279–290. [Online]. Available: https://doi.org/10.1145/3352460.3358265
  • [29] P. Murali, D. C. Mckay, M. Martonosi, and A. Javadi-Abhari, “Software mitigation of crosstalk on noisy intermediate-scale quantum computers,” in Proceedings of the Twenty-Fifth International Conference on Architectural Support for Programming Languages and Operating Systems, ser. ASPLOS ’20.   New York, NY, USA: Association for Computing Machinery, 2020, p. 1001–1016. [Online]. Available: https://doi.org/10.1145/3373376.3378477
  • [30] P. Das, C. A. Pattison, S. Manne, D. M. Carmean, K. M. Svore, M. Qureshi, and N. Delfosse, “Afs: Accurate, fast, and scalable error-decoding for fault-tolerant quantum computers,” in 2022 IEEE International Symposium on High-Performance Computer Architecture (HPCA), 2022, pp. 259–273.
  • [31] Y. Suzuki, T. Sugiyama, T. Arai, W. Liao, K. Inoue, and T. Tanimoto, “Q3de: A fault-tolerant quantum computer architecture for multi-bit burst errors by cosmic rays,” in 2022 55th IEEE/ACM International Symposium on Microarchitecture (MICRO), 2022, pp. 1110–1125.
  • [32] A. JavadiAbhari, S. Patil, D. Kudrow, J. Heckey, A. Lvov, F. T. Chong, and M. Martonosi, “Scaffcc: A framework for compilation and analysis of quantum computing programs,” in Proceedings of the 11th ACM Conference on Computing Frontiers, ser. CF ’14.   New York, NY, USA: Association for Computing Machinery, 2014. [Online]. Available: https://doi.org/10.1145/2597917.2597939
  • [33] W. Tang, T. Tomesh, M. Suchara, J. Larson, and M. Martonosi, “Cutqc: using small quantum computers for large quantum circuit evaluations,” in Proceedings of the 26th ACM International conference on architectural support for programming languages and operating systems, 2021, pp. 473–486.
  • [34] P. Murali, N. M. Linke, M. Martonosi, A. J. Abhari, N. H. Nguyen, and C. H. Alderete, “Full-stack, real-system quantum computer studies: Architectural comparisons and design insights,” in Proceedings of the 46th International Symposium on Computer Architecture, ser. ISCA ’19.   New York, NY, USA: Association for Computing Machinery, 2019, p. 527–540. [Online]. Available: https://doi.org/10.1145/3307650.3322273
  • [35] S. Mittal and S. Vaishay, “A survey of techniques for optimizing deep learning on gpus,” Journal of Systems Architecture, vol. 99, p. 101635, 2019.
  • [36] I. M. Georgescu, S. Ashhab, and F. Nori, “Quantum simulation,” Rev. Mod. Phys., vol. 86, no. 1, pp. 153–185, Mar. 2014.
  • [37] A. Aspuru-Guzik, A. D. Dutoi, P. J. Love, and M. Head-Gordon, “Simulated Quantum Computation of Molecular Energies,” Science, vol. 309, no. 5741, pp. 1704–1707, Sep. 2005.
  • [38] R. Orús, S. Mugel, and E. Lizaso, “Quantum computing for finance: Overview and prospects,” Rev. Phys., vol. 4, p. 100028, Nov. 2019.
  • [39] P. Rebentrost, B. Gupt, and T. R. Bromley, “Quantum computational finance: Monte carlo pricing of financial derivatives,” Physical Review A, vol. 98, no. 2, p. 022321, 2018.
  • [40] J. Biamonte, P. Wittek, N. Pancotti, P. Rebentrost, N. Wiebe, and S. Lloyd, “Quantum machine learning,” Nature, vol. 549, no. 7671, pp. 195–202, 2017.
  • [41] 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, vol. 3, no. 3, p. 030503, 2018.
  • [42] “Quantum computing: Progress and prospects,” National Academies of Sciences, Engineering, and Medicine, 2019, National Academies Press.
  • [43] R. Ratcliff, “A theory of memory retrieval,” Psychological Review, vol. 85, no. 2, pp. 59–108, 1978.
  • [44] M. Usher and J. L. McClelland, “The time course of perceptual choice: the leaky, competing accumulator model.” Psychological review, vol. 108, no. 3, p. 550, 2001.
  • [45] S. Musslick and J. D. Cohen, “Rationalizing constraints on the capacity for cognitive control,” Trends in Cognitive Sciences, vol. 25, no. 9, pp. 757–775, 2021.
  • [46] A. Graves, G. Wayne, and I. Danihelka, “Neural turing machines,” arXiv preprint arXiv:1410.5401, 2014.
  • [47] D. Rolnick, A. Ahuja, J. Schwarz, T. Lillicrap, and G. Wayne, “Experience replay for continual learning,” Advances in Neural Information Processing Systems, vol. 32, 2019.
  • [48] T. W. Webb, I. Sinha, and J. Cohen, “Emergent symbols through binding in external memory,” in International Conference on Learning Representations, 2020.
  • [49] B. M. Lake, T. D. Ullman, J. B. Tenenbaum, and S. J. Gershman, “Building machines that learn and think like people,” Behavioral and brain sciences, vol. 40, p. e253, 2017.
  • [50] A. Tversky and E. Shafir, “The disjunction effect in choice under uncertainty,” Psychological science, vol. 3, no. 5, pp. 305–310, 1992.
  • [51] E. M. Pothos and J. R. Busemeyer, “Can quantum probability provide a new direction for cognitive modeling?” Behavioral and brain sciences, vol. 36, no. 3, pp. 255–274, 2013.
  • [52] P. D. Kvam, T. J. Pleskac, S. Yu, and J. R. Busemeyer, “Interference effects of choice on confidence: Quantum characteristics of evidence accumulation,” Proceedings of the National Academy of Sciences, vol. 112, no. 34, pp. 10 645–10 650, 2015.
  • [53] J. R. Busemeyer, Z. Wang, and J. T. Townsend, “Quantum dynamics of human decision-making,” Journal of Mathematical Psychology, vol. 50, no. 3, pp. 220–241, 2006.
  • [54] J. R. Busemeyer, P. D. Kvam, and T. J. Pleskac, “Comparison of markov versus quantum dynamical models of human decision making,” Wiley Interdisciplinary Reviews: Cognitive Science, vol. 11, no. 4, p. e1526, 2020.
  • [55] T. Tomesh, P. Gokhale, V. Omole, G. S. Ravi, K. N. Smith, J. Viszlai, X.-C. Wu, N. Hardavellas, M. R. Martonosi, and F. T. Chong, “Supermarq: A scalable quantum benchmark suite,” in 2022 IEEE International Symposium on High-Performance Computer Architecture (HPCA).   IEEE, 2022, pp. 587–603.
  • [56] A. Mallick, S. Mandal, A. Karan, and C. M. Chandrashekar, “Simulating dirac hamiltonian in curved space-time by split-step quantum walk,” Journal of Physics Communications, vol. 3, no. 1, p. 015012, 2019.
  • [57] T. Oka, N. Konno, R. Arita, and H. Aoki, “Breakdown of an electric-field driven system: a mapping to a quantum walk,” Physical review letters, vol. 94, no. 10, p. 100602, 2005.
  • [58] M. Ferdman, A. Adileh, O. Kocberber, S. Volos, M. Alisafaee, D. Jevdjic, C. Kaynak, A. D. Popescu, A. Ailamaki, and B. Falsafi, “Clearing the clouds: A study of emerging scale-out workloads on modern hardware,” SIGPLAN Not., vol. 47, no. 4, p. 37–48, mar 2012. [Online]. Available: https://doi.org/10.1145/2248487.2150982
  • [59] R. Ratcliff, P. L. Smith, S. D. Brown, and G. McKoon, “Diffusion decision model: Current issues and history,” Trends in cognitive sciences, vol. 20, no. 4, pp. 260–281, 2016.
  • [60] J. R. Busemeyer, S. Gluth, J. Rieskamp, and B. M. Turner, “Cognitive and neural bases of multi-attribute, multi-alternative, value-based decisions,” Trends in cognitive sciences, vol. 23, no. 3, pp. 251–263, 2019.
  • [61] T. L. Willke, S. B. M. Yoo, M. Capotă, S. Musslick, B. Y. Hayden, and J. D. Cohen, “A comparison of non-human primate and deep reinforcement learning agent performance in a virtual pursuit-avoidance task,” in Reinforcement Learning and Decision Making Conference, 2019.
  • [62] R. Bogacz, M. Usher, J. Zhang, and J. L. McClelland, “Extending a biologically inspired model of choice: multi-alternatives, nonlinearity and value-based multidimensional choice,” Philosophical Transactions of the Royal Society B: Biological Sciences, vol. 362, no. 1485, pp. 1655–1670, 2007.
  • [63] Y. Ding and F. T. Chong, “Circuit synthesis and compilation,” in Quantum Computer Systems: Research for Noisy Intermediate-Scale Quantum Computers.   Springer, 2020, pp. 91–125.
  • [64] A. Matuschak and M. A. Nielsen, “Quantum Computing for the Very Curious,” 2019. [Online]. Available: https://quantum.country/qcvc
  • [65] E. Farhi and S. Gutmann, “Quantum computation and decision trees,” Physical Review A, vol. 58, no. 2, p. 915, 1998.
  • [66] S. E. Venegas-Andraca, “Quantum walks: a comprehensive review,” Quantum Information Processing, vol. 11, no. 5, pp. 1015–1106, 2012.
  • [67] R. Wille, R. Van Meter, and Y. Naveh, “Ibm’s qiskit tool chain: Working with and developing for real quantum computers,” in 2019 Design, Automation & Test in Europe Conference & Exhibition (DATE).   IEEE, 2019, pp. 1234–1240.
  • [68] A. Peruzzo, J. McClean, P. Shadbolt, M.-H. Yung, X.-Q. Zhou, P. J. Love, A. Aspuru-Guzik, and J. L. O’Brien, “A variational eigenvalue solver on a photonic quantum processor,” Nat. Commun., vol. 5, no. 4213, pp. 1–7, Jul. 2014.
  • [69] J. R. McClean, J. Romero, R. Babbush, and A. Aspuru-Guzik, “The theory of variational hybrid quantum-classical algorithms,” New Journal of Physics, vol. 18, no. 2, p. 023023, 2016.
  • [70] E. Farhi, J. Goldstone, and S. Gutmann, “A quantum approximate optimization algorithm,” arXiv preprint arXiv:1411.4028, 2014.
  • [71] K. M. Nakanishi, K. Mitarai, and K. Fujii, “Subspace-search variational quantum eigensolver for excited states,” Physical Review Research, vol. 1, no. 3, p. 033062, 2019.
  • [72] S. Sim, P. D. Johnson, and A. Aspuru-Guzik, “Expressibility and entangling capability of parameterized quantum circuits for hybrid quantum-classical algorithms,” Advanced Quantum Technologies, vol. 2, no. 12, p. 1900070, 2019.
  • [73] M. Motta, C. Sun, A. T. Tan, M. J. O’Rourke, E. Ye, A. J. Minnich, F. G. Brandao, and G. K.-L. Chan, “Determining eigenstates and thermal states on a quantum computer using quantum imaginary time evolution,” Nature Physics, vol. 16, no. 2, pp. 205–210, 2020.
  • [74] S. McArdle, T. Jones, S. Endo, Y. Li, S. C. Benjamin, and X. Yuan, “Variational ansatz-based quantum simulation of imaginary time evolution,” npj Quantum Inf., vol. 5, no. 75, pp. 1–6, Sep. 2019.
  • [75] T. H. Kyaw, M. B. Soley, B. Allen, P. Bergold, C. Sun, V. S. Batista, and A. Aspuru-Guzik, “Variational quantum iterative power algorithms for global optimization,” arXiv, Aug. 2022.
  • [76] G. Saxena, A. Shalabi, and T. H. Kyaw, “Practical limitations of quantum data propagation on noisy quantum processors,” arXiv, Jun. 2023.
  • [77] K. Kurowski, M. Slysz, M. Subocz, and R. Różycki, “Applying a quantum annealing based restricted boltzmann machine for mnist handwritten digit classification.” Computational Methods in Science & Technology, vol. 27, no. 3, 2021.
  • [78] P. Hauke, H. G. Katzgraber, W. Lechner, H. Nishimori, and W. D. Oliver, “Perspectives of quantum annealing: Methods and implementations,” Reports on Progress in Physics, vol. 83, no. 5, p. 054401, 2020.
  • [79] E. Farhi, J. Goldstone, S. Gutmann, and M. Sipser, “Quantum computation by adiabatic evolution,” arXiv preprint quant-ph/0001106, 2000.
  • [80] T. H. Kyaw, Y. Li, and L.-C. Kwek, “Measurement-Based Quantum Computation on Two-Body Interacting Qubits with Adiabatic Evolution,” Phys. Rev. Lett., vol. 113, no. 18, p. 180501, Oct. 2014.
  • [81] T. H. Kyaw and L.-C. Kwek, “Cluster state generation in one-dimensional Kitaev honeycomb model via shortcut to adiabaticity,” New J. Phys., vol. 20, no. 4, p. 045007, Apr. 2018.
  • [82] T. H. Kyaw, S. Allende, L.-C. Kwek, and G. Romero, “Parity-preserving light-matter system mediates effective two-body interactions,” Quantum Sci. Technol., vol. 2, no. 2, p. 025007, May 2017.
  • [83] W. van Dam, M. Mosca, and U. Vazirani, “How powerful is adiabatic quantum computation?” in Proceedings 42nd IEEE Symposium on Foundations of Computer Science.   IEEE, Oct. 2001, pp. 279–287.
  • [84] “QuEra.” [Online]. Available: https://www.quera.com/
  • [85] Y. Peng, J. Young, P. Liu, and X. Wu, “Simuq: A domain-specific language for quantum simulation with analog compilation,” arXiv preprint arXiv:2303.02775, 2023.
  • [86] R. Xia, T. Bian, and S. Kais, “Electronic structure calculations and the ising hamiltonian,” The Journal of Physical Chemistry B, vol. 122, no. 13, pp. 3384–3395, 2017.
  • [87] G. Li, L. Zhou, N. Yu, Y. Ding, M. Ying, and Y. Xie, “Projection-based runtime assertions for testing and debugging quantum programs,” Proceedings of the ACM on Programming Languages, vol. 4, no. OOPSLA, pp. 1–29, 2020.
  • [88] E. Boros and P. L. Hammer, “Pseudo-boolean optimization,” Discrete applied mathematics, vol. 123, no. 1-3, pp. 155–225, 2002.
  • [89] P. Gokhale, O. Angiuli, Y. Ding, K. Gui, T. Tomesh, M. Suchara, M. Martonosi, and F. T. Chong, “Optimization of simultaneous measurement for variational quantum eigensolver applications,” in 2020 IEEE International Conference on Quantum Computing and Engineering (QCE), 2020, pp. 379–390.
  • [90] J. M. Clary, E. B. Jones, D. Vigil-Fowler, C. Chang, and P. Graf, “Exploring the scaling limitations of the variational quantum eigensolver with the bond dissociation of hydride diatomic molecules,” International Journal of Quantum Chemistry, p. e27097, 2023.
  • [91] L. Mineh and A. Montanaro, “Accelerating the variational quantum eigensolver using parallelism,” Quantum Science and Technology, vol. 8, no. 3, p. 035012, 2023.
  • [92] J. R. McClean, M. E. Kimchi-Schwartz, J. Carter, and W. A. de Jong, “Hybrid quantum-classical hierarchy for mitigation of decoherence and determination of excited states,” Phys. Rev. A, vol. 95, no. 4, p. 042308, Apr. 2017.
  • [93] P. Smolensky, “Information processing in dynamical systems: Foundations of harmony theory,” 1986, department of Computer Science, University of Colorado, Boulder.
  • [94] G. E. Hinton, “Learning multiple layers of representation,” Trends in cognitive sciences, vol. 11, no. 10, pp. 428–434, 2007.
  • [95] B. Recht, C. Re, S. Wright, and F. Niu, “Hogwild!: A lock-free approach to parallelizing stochastic gradient descent,” Advances in neural information processing systems, vol. 24, 2011.
  • [96] S. Jiang, K. A. Britt, A. J. McCaskey, T. S. Humble, and S. Kais, “Quantum annealing for prime factorization,” Scientific reports, vol. 8, no. 1, p. 17667, 2018.
  • [97] S. A. Rahman, R. Lewis, E. Mendicelli, and S. Powell, “Su (2) lattice gauge theory on a quantum annealer,” Physical Review D, vol. 104, no. 3, p. 034501, 2021.
  • [98] “QPU-Specific Characteristics.” [Online]. Available: https://docs.dwavesys.com/docs/latest/_downloads/114d18bf1a9d2fd6fea1a16c89c30799/09-1262A-C_QPU_Properties_Advantage_system4_1.pdf
  • [99] J. Marshall, D. Venturelli, I. Hen, and E. G. Rieffel, “Power of pausing: Advancing understanding of thermalization in experimental quantum annealers,” Phys. Rev. Appl., vol. 11, p. 044083, Apr 2019. [Online]. Available: https://link.aps.org/doi/10.1103/PhysRevApplied.11.044083
  • [100] M. A. Carreira-Perpinan and G. Hinton, “On contrastive divergence learning,” in International workshop on artificial intelligence and statistics.   PMLR, 2005, pp. 33–40.
  • [101] E. T. Campbell, B. M. Terhal, and C. Vuillot, “Roads towards fault-tolerant universal quantum computation,” Nature, vol. 549, no. 7671, pp. 172–179, 2017.
  • [102] A. Kandala, A. Mezzacapo, K. Temme, M. Takita, M. Brink, J. M. Chow, and J. M. Gambetta, “Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets,” nature, vol. 549, no. 7671, pp. 242–246, 2017.
  • [103] H. Wang, Y. Ding, J. Gu, Y. Lin, D. Z. Pan, F. T. Chong, and S. Han, “Quantumnas: Noise-adaptive search for robust quantum circuits,” in 2022 IEEE International Symposium on High-Performance Computer Architecture (HPCA).   IEEE, 2022, pp. 692–708.
  • [104] “IBM Quantum.” [Online]. Available: https://quantum-computing.ibm.com/lab/docs/iql/manage/systems/queue
  • [105] “IBM Quantum.” [Online]. Available: https://qiskit.org/ecosystem/ibm-runtime/sessions.html
  • [106] M. Satyanarayanan, P. Bahl, R. Caceres, and N. Davies, “The case for vm-based cloudlets in mobile computing,” IEEE pervasive Computing, vol. 8, no. 4, pp. 14–23, 2009.
  • [107] S. Stein, N. Wiebe, Y. Ding, P. Bo, K. Kowalski, N. Baker, J. Ang, and A. Li, “Eqc: ensembled quantum computing for variational quantum algorithms,” in Proceedings of the 49th Annual International Symposium on Computer Architecture, 2022, pp. 59–71.
  • [108] N. H. Stair and F. A. Evangelista, “Qforte: an efficient state simulator and quantum algorithms library for molecular electronic structure,” arXiv preprint arXiv:2108.04413, 2021.
  • [109] J. R. McClean, N. C. Rubin, K. J. Sung, I. D. Kivlichan, X. Bonet-Monroig, Y. Cao, C. Dai, E. S. Fried, C. Gidney, B. Gimby, P. Gokhale, T. Häner, T. Hardikar, V. Havlíček, O. Higgott, C. Huang, J. Izaac, Z. Jiang, X. Liu, S. McArdle, M. Neeley, T. O’Brien, B. O’Gorman, I. Ozfidan, M. D. Radin, J. Romero, N. P. D. Sawaya, B. Senjean, K. Setia, S. Sim, D. S. Steiger, M. Steudtner, Q. Sun, W. Sun, D. Wang, F. Zhang, and R. Babbush, “Openfermion: the electronic structure package for quantum computers,” Quantum Science and Technology, vol. 5, no. 3, p. 034014, 2020.
  • [110] J. S. Kottmann, S. Alperin-Lea, T. Tamayo-Mendoza, A. Cervera-Lierta, C. Lavigne, T.-C. Yen, V. Verteletskyi, P. Schleich, A. Anand, M. Degroote, S. Chaney, M. Kesibi, N. G. Curnow, B. Solo, G. Tsilimigkounakis, C. Zendejas-Morales, A. F. Izmaylov, and A. Aspuru-Guzik, “TEQUILA: a platform for rapid development of quantum algorithms,” Quantum Sci. Technol., vol. 6, no. 2, p. 024009, Mar. 2021.
  • [111] X.-Z. Luo, J.-G. Liu, P. Zhang, and L. Wang, “Yao.jl: Extensible, Efficient Framework for Quantum Algorithm Design,” Quantum, vol. 4, p. 341, Oct. 2020.
  • [112] T. Alexander, N. Kanazawa, D. J. Egger, L. Capelluto, C. J. Wood, A. Javadi-Abhari, and D. C. McKay, “Qiskit pulse: programming quantum computers through the cloud with pulses,” Quantum Science and Technology, vol. 5, no. 4, p. 044006, 2020.
  • [113] P. Gokhale, A. Javadi-Abhari, N. Earnest, Y. Shi, and F. T. Chong, “Optimized quantum compilation for near-term algorithms with openpulse,” in 2020 53rd Annual IEEE/ACM International Symposium on Microarchitecture (MICRO).   IEEE, 2020, pp. 186–200.
  • [114] T. Imoto, Y. Seki, Y. Matsuzaki, and S. Kawabata, “Guaranteed-accuracy quantum annealing,” Physical Review A, vol. 106, no. 4, p. 042615, 2022.
  • [115] T. Kopyciuk, M. Lewandowski, and P. Kurzyński, “Pre-and post-selection paradoxes in quantum walks,” New Journal of Physics, vol. 21, no. 10, p. 103054, 2019.
  • [116] J. T. Monroe, “Partial measurements of quantum systems,” Ph.D. dissertation, Washington University in St. Louis, 2021.
  • [117] 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, vol. 574, no. 7779, pp. 505–510, Oct 2019. [Online]. Available: https://doi.org/10.1038/s41586-019-1666-5
  • [118] 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, vol. 370, no. 6523, pp. 1460–1463, Dec. 2020.
  • [119] Y. Wu, W.-S. Bao, S. Cao, F. Chen, M.-C. Chen, X. Chen, T.-H. Chung, H. Deng, Y. Du, D. Fan, M. Gong, C. Guo, C. Guo, S. Guo, L. Han, L. Hong, H.-L. Huang, Y.-H. Huo, L. Li, N. Li, S. Li, Y. Li, F. Liang, C. Lin, J. Lin, H. Qian, D. Qiao, H. Rong, H. Su, L. Sun, L. Wang, S. Wang, D. Wu, Y. Xu, K. Yan, W. Yang, Y. Yang, Y. Ye, J. Yin, C. Ying, J. Yu, C. Zha, C. Zhang, H. Zhang, K. Zhang, Y. Zhang, H. Zhao, Y. Zhao, L. Zhou, Q. Zhu, C.-Y. Lu, C.-Z. Peng, X. Zhu, and J.-W. Pan, “Strong Quantum Computational Advantage Using a Superconducting Quantum Processor,” Phys. Rev. Lett., vol. 127, no. 18, p. 180501, Oct. 2021.
  • [120] L. S. Madsen, F. Laudenbach, M. F. Askarani, F. Rortais, T. Vincent, J. F. F. Bulmer, F. M. Miatto, L. Neuhaus, L. G. Helt, M. J. Collins, A. E. Lita, T. Gerrits, S. W. Nam, V. D. Vaidya, M. Menotti, I. Dhand, Z. Vernon, N. Quesada, and J. Lavoie, “Quantum computational advantage with a programmable photonic processor,” Nature, vol. 606, no. 7912, pp. 75–81, Jun 2022. [Online]. Available: https://doi.org/10.1038/s41586-022-04725-x
  • [121] K. Bharti, A. Cervera-Lierta, T. H. Kyaw, T. Haug, S. Alperin-Lea, A. Anand, M. Degroote, H. Heimonen, J. S. Kottmann, T. Menke, W.-K. Mok, S. Sim, L.-C. Kwek, and A. Aspuru-Guzik, “Noisy intermediate-scale quantum algorithms,” Rev. Mod. Phys., vol. 94, no. 1, p. 015004, Feb. 2022.
  • [122] J. Zhang, T. H. Kyaw, S. Filipp, L.-C. Kwek, E. Sjöqvist, and D. Tong, “Geometric and holonomic quantum computation,” Phys. Rep., vol. 1027, pp. 1–53, Jul. 2023.
  • [123] A. Fedorov and M. Gelfand, “Towards practical applications in quantum computational biology,” Nature Computational Science, vol. 1, no. 2, pp. 114–119, 2021.
  • [124] R. P. Feynman, “Simulating physics with computers,” International Journal of Theoretical Physics, vol. 21, no. 6/7, 1982.
  • [125] T. Lubinski, S. Johri, P. Varosy, J. Coleman, L. Zhao, J. Necaise, C. H. Baldwin, K. Mayer, and T. Proctor, “Application-oriented performance benchmarks for quantum computing,” IEEE Transactions on Quantum Engineering, 2023.
  • [126] J. R. Finžgar, P. Ross, L. Hölscher, J. Klepsch, and A. Luckow, “Quark: A framework for quantum computing application benchmarking,” in 2022 IEEE International Conference on Quantum Computing and Engineering (QCE).   IEEE, 2022, pp. 226–237.
  • [127] V. Shende, S. Bullock, and I. Markov, “Synthesis of quantum-logic circuits,” IEEE Transactions on Computer-Aided Design of Integrated Circuits and Systems, vol. 25, no. 6, pp. 1000–1010, 2006.
  • [128] M. Amy, D. Maslov, M. Mosca, and M. Roetteler, “A meet-in-the-middle algorithm for fast synthesis of depth-optimal quantum circuits,” IEEE Transactions on Computer-Aided Design of Integrated Circuits and Systems, vol. 32, no. 6, pp. 818–830, 2013.
  • [129] P. Das, S. Tannu, and M. Qureshi, “Jigsaw: Boosting fidelity of nisq programs via measurement subsetting,” in MICRO-54: 54th Annual IEEE/ACM International Symposium on Microarchitecture, 2021, pp. 937–949.
  • [130] G. S. Ravi, P. Gokhale, Y. Ding, W. Kirby, K. Smith, J. M. Baker, P. J. Love, H. Hoffmann, K. R. Brown, and F. T. Chong, “Cafqa: A classical simulation bootstrap for variational quantum algorithms,” in Proceedings of the 28th ACM International Conference on Architectural Support for Programming Languages and Operating Systems, Volume 1, 2022, pp. 15–29.
  • [131] G. S. Ravi, K. Smith, J. M. Baker, T. Kannan, N. Earnest, A. Javadi-Abhari, H. Hoffmann, and F. T. Chong, “Navigating the dynamic noise landscape of variational quantum algorithms with qismet,” in Proceedings of the 28th ACM International Conference on Architectural Support for Programming Languages and Operating Systems, Volume 2, 2023, pp. 515–529.
  • [132] G. S. Ravi, K. N. Smith, P. Murali, and F. T. Chong, “Adaptive job and resource management for the growing quantum cloud,” in 2021 IEEE International Conference on Quantum Computing and Engineering (QCE).   IEEE, 2021, pp. 301–312.
  • [133] G. S. Ravi, K. N. Smith, P. Gokhale, and F. T. Chong, “Quantum computing in the cloud: Analyzing job and machine characteristics,” in 2021 IEEE International Symposium on Workload Characterization (IISWC).   IEEE, 2021, pp. 39–50.
  • [134] J. Yao, J. Wang, F. Yue, J. Xu, and Z. Shan, “Mtmc: A scheduling framework of multi-tasking mapping on multi-chips,” 2022. [Online]. Available: https://www.researchsquare.com/article/rs-1857463/latest.pdf
  • [135] M. Zhang, L. Xie, Z. Zhang, Q. Yu, G. Xi, H. Zhang, F. Liu, Y. Zheng, Y. Zheng, and S. Zhang, “Exploiting different levels of parallelism in the quantum control microarchitecture for superconducting qubits,” in MICRO-54: 54th Annual IEEE/ACM International Symposium on Microarchitecture, 2021, pp. 898–911.
  • [136] S. Niu and A. Todri-Sanial, “How parallel circuit execution can be useful for nisq computing?” in 2022 Design, Automation & Test in Europe Conference & Exhibition (DATE).   IEEE, 2022, pp. 1065–1070.
  • [137] Y. Ohkura, T. Satoh, and R. Van Meter, “Simultaneous execution of quantum circuits on current and near-future nisq systems,” IEEE Transactions on Quantum Engineering, vol. 3, pp. 1–10, 2022.
  • [138] P. Das, S. S. Tannu, P. J. Nair, and M. Qureshi, “A case for multi-programming quantum computers,” in Proceedings of the 52nd Annual IEEE/ACM International Symposium on Microarchitecture, 2019, pp. 291–303.
  • [139] L. Liu and X. Dou, “Qucloud: A new qubit mapping mechanism for multi-programming quantum computing in cloud environment,” in 2021 IEEE International symposium on high-performance computer architecture (HPCA).   IEEE, 2021, pp. 167–178.
  • [140] T. Huang, Y. Zhu, R. S. M. Goh, and T. Luo, “When quantum annealing meets multitasking: Potentials, challenges and opportunities,” Array, p. 100282, 2023.
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" alt="[LOGO]">