Enrico Prati’s research while affiliated with University of Milan and other places

What is this page?


This page lists works of an author who doesn't have a ResearchGate profile or hasn't added the works to their profile yet. It is automatically generated from public (personal) data to further our legitimate goal of comprehensive and accurate scientific recordkeeping. If you are this author and want this page removed, please let us know.

Publications (168)


Graphical representation of the approach followed in this paper. The circles represent variables, the rectangles represent operations. The variables are also represented by primary colors, while the methods are represented by the secondary colors obtained by combining the colors of their input and their output. In blue there are the neural network inputs. In purple there are the two neural networks, the main one that computes the eigenstates and the one that computes the energy. In red there are the outputs of the neural networks: the auxiliary outputs, the main outputs which is the eigenfunction and the eigenvalue E. In yellow there are the operations needed to calculate the losses. The losses are the integral loss, the boundary condition loss, the normalization loss, the inductive biases and the differential equation loss. SAE is the Sum of Absolute Errors, SSE is the Sum of Squared Errors. In green there is the final loss. It is calculated by summing the partial losses weighted by empirically adjusted hyperparameters wint, wbc, wnorm, wbias and weq. In cyan there are the criteria and the optimization method.
(a) Ground state for the harmonic oscillator. The red dots are the PINN’s predictions, while the blue line is the ground truth. (b) Fidelity throughout training for the ground state. (c) Loss behavior throughout training. The y axis is in logarithmic scale, therefore the oscillations on the y axis for low losses are overemphasized.
(a) Fifth excited state for the harmonic oscillator. The red dots are the PINN’s predictions, while the blue line is the ground truth. (b) Fidelity throughout training for the ground state. (c) Loss behavior throughout training. The y axis is in logarithmic scale, therefore the oscillations on the y axis for low losses are overemphasized.
Behavior of the seven losses throughout training for the (a) ground state and (b) fifth excited state of the harmonic oscillator. The orthogonality loss is displayed in blue. Note that this loss cannot be seen in (a) as it is always 0 for the ground state. The symmetry loss is displayed in magenta. The integral loss is displayed in olive green. The normalization loss is displayed in orange. The boundary conditions loss is displayed in brown. The differential equation loss is displayed in black.
Squared error relative to the position x for the (a) ground state and (c) fifth excited state.

+5

Addressing the non-perturbative regime of the quantum anharmonic oscillator by physics-informed neural networks
  • Article
  • Full-text available

October 2024

·

15 Reads

·

2 Citations

·

·

Enrico Prati

The use of deep learning in physical sciences has recently boosted the ability of researchers to tackle physical systems where little or no analytical insight is available. Recently, the Physics−Informed Neural Networks (PINNs) have been introduced as one of the most promising tools to solve systems of differential equations guided by some physically grounded constraints. In the quantum realm, such an approach paves the way to a novel approach to solve the Schrödinger equation for non-integrable systems. By following an unsupervised learning approach, we apply the PINNs to the anharmonic oscillator in which an interaction term proportional to the fourth power of the position coordinate is present. We compute the eigenenergies and the corresponding eigenfunctions while varying the weight of the quartic interaction. We bridge our solutions to the regime where both the perturbative and the strong coupling theory work, including the pure quartic oscillator. We investigate systems with real and imaginary frequency, laying the foundation for novel numerical methods to tackle problems emerging in quantum field theory.

Download

Figure 2. (a) Ground state for the well. The red dots are the PINN's predictions, while the blue line is the ground truth. The yellow vertical lines represent the walls of the well. (b) Fidelity throughout training for the ground state. (c) Loss behavior throughout training. The y-axis is in logarithmic scale; therefore, the oscillations on the y-axis for low losses are overemphasized.
Figure 3. (a) Fifth excited state for the well Equation (22). The red dots are the PINN's predictions, while the blue line is the ground truth. The yellow vertical lines represent the walls of the well. (b) Fidelity throughout training for the ground state. (c) Loss behavior throughout training. The y-axis is in logarithmic scale; therefore, the oscillations on the y-axis for low losses are overemphasized
Figure 4. (a) One of the two first excited states for the ring Equation (26). The red dots are the PINN's predictions, while the blue line is the ground truth. Note that in this case, the ground truth and the neural prediction are actually different states of the pair of degenerate states for |n| = 1. This discrepancy is due to the fact that the two minima given by the states for n = 1 and n = −1 are equally valid, and thus the neural network can converge freely to either one. However, in order to have fidelity be an informative metric, the real and imaginary parts of the neural network predictions are first multiplied by sign(⟨Re(ψ Ex )|Re(ψ PI NN )⟩) and sign(⟨Im(ψ Ex )|Im(ψ PI NN )⟩), respectively, resulting in the two plots superimposing in the figure. (b) Fidelity throughout training for the ground state. (c) Loss behavior throughout training. The y-axis is in logarithmic scale; therefore, the oscillations on the y-axis for low losses are overemphasized
Weights for the infinite potential well.
Metrics for the infinite potential well.
A Tutorial on the Use of Physics-Informed Neural Networks to Compute the Spectrum of Quantum Systems

September 2024

·

30 Reads

·

1 Citation

Technologies

Quantum many-body systems are of great interest for many research areas, including physics, biology, and chemistry. However, their simulation is extremely challenging, due to the exponential growth of the Hilbert space with system size, making it exceedingly difficult to parameterize the wave functions of large systems by using exact methods. Neural networks and machine learning, in general, are a way to face this challenge. For instance, methods like tensor networks and neural quantum states are being investigated as promising tools to obtain the wave function of a quantum mechanical system. In this tutorial, we focus on a particularly promising class of deep learning algorithms. We explain how to construct a Physics-Informed Neural Network (PINN) able to solve the Schrödinger equation for a given potential, by finding its eigenvalues and eigenfunctions. This technique is unsupervised, and utilizes a novel computational method in a manner that is barely explored. PINNs are a deep learning method that exploit automatic differentiation to solve integro-differential equations in a mesh-free way. We show how to find both the ground and the excited states. The method discovers the states progressively by starting from the ground state. We explain how to introduce inductive biases in the loss to exploit further knowledge of the physical system. Such additional constraints allow for a faster and more accurate convergence. This technique can then be enhanced by a smart choice of collocation points in order to take advantage of the mesh-free nature of the PINN. The methods are made explicit by applying them to the infinite potential well and the particle in a ring, a challenging problem to be learned by an artificial intelligence agent due to the presence of complex-valued eigenfunctions and degenerate states


Quantum data encoding as a distinct abstraction layer in the design of quantum circuits

September 2024

·

16 Reads

Complex quantum circuits are constituted by combinations of quantum subroutines. The computation is possible as long as the quantum data encoding is consistent throughout the circuit. Despite its fundamental importance, the formalization of quantum data encoding has never been addressed systematically so far. We formalize the concept of quantum data encoding, namely the format providing a representation of a data set through a quantum state, as a distinct abstract layer with respect to the associated data loading circuit. We survey existing encoding methods and their respective strategies for classical-to-quantum exact and approximate data loading, for the quantum-to-classical extraction of information from states, and for quantum-to-quantum encoding conversion. Next, we show how major quantum algorithms find a natural interpretation in terms of data loading. For instance, the Quantum Fourier Transform is described as a quantum encoding converter, while the Quantum Amplitude Estimation as an extraction routine. The new conceptual framework is exemplified by considering its application to quantum-based Monte Carlo simulations, thus showcasing the power of the proposed formalism for the description of complex quantum circuits. Indeed, the approach clarifies the structure of complex quantum circuits and enables their efficient design.


Leveraging non-unital noise for gate-based quantum reservoir computing

September 2024

·

21 Reads

·

1 Citation

We identify a noise model that ensures the functioning of an echo state network employing a gate-based quantum computer for reservoir computing applications. Energy dissipation induced by amplitude damping drastically improves the short-term memory capacity and expressivity of the network, by simultaneously providing fading memory and richer dynamics. There is an ideal dissipation rate that ensures the best operation of the echo state network around γ\gamma\sim 0.03. Nevertheless, these beneficial effects are stable as the intensity of the applied noise increases. The improvement of the learning is confirmed by emulating a realistic noise model applied to superconducting qubits, paving the way for the application of reservoir computing methods in current non-fault-tolerant quantum computers.


Fig. 1. The quantum register architecture
Enforcing fading memory of noisy quantum echo state networks

Reservoir computing is a versatile paradigm in computational neuroscience and machine learning that uses recurrent neural networks to process time-dependent inputs. We leverage noise in gate-based quantum computers to enforce the fading memory property of a quantum circuit acting as a reservoir. We test the learning architecture on standard bench-marking tasks, such as NARMA and short-term memory tasks, by providing numerical evidence supporting the existence of an optimal noise regime that enhances predictive performances. This critical behavior indicates the presence of a regime in which several network capabilities, such as short-term memory capacity and expressivity, are maximized. Our results pave the way for the efficient use of current noisy quantum devices in quantum machine learning applications.


Quantum Machine Learning Algorithms for Anomaly Detection: a Survey

August 2024

·

44 Reads

·

Lorenzo Moro

·

·

[...]

·

Enrico Prati

The advent of quantum computers has justified the development of quantum machine learning algorithms , based on the adaptation of the principles of machine learning to the formalism of qubits. Among such quantum algorithms, anomaly detection represents an important problem crossing several disciplines from cybersecurity, to fraud detection to particle physics. We summarize the key concepts involved in quantum computing, introducing the formal concept of quantum speed up. The survey provides a structured map of anomaly detection based on quantum machine learning. We have grouped existing algorithms according to the different learning methods, namely quantum supervised, quantum unsupervised and quantum reinforcement learning, respectively. We provide an estimate of the hardware resources to provide sufficient computational power in the future. The survey provides a systematic and compact understanding of the techniques belonging to each category. We eventually provide a discussion on the computational complexity of the learning methods in real application domains.


FIG. 2. The minimization problem solved by the variational quantum deflation (VQD) algorithm has a penalty term that accounts for the orthogonality of the state with a previously-found ground state. The coefficient λ is the weight of such constraint in the minimization problem. The choice of λ is self-correcting, meaning that by choosing an incorrect value λ = γ − E 0 ≤ ∆E, will cause the algorithm to find a minimum proportional to γ. From the upper plot, it is evident the linear behaviour of the cost function for λ < ∆E (red shaded area), and then a constant optimal value for λ ≥ ∆E (green shaded area). The lower plot keeps track of the overlap of the candidate excited state with the ground state. For the results of this article, λ is set to be equal to 4.
FIG. 5. Performance of the minimization of the triton energy min θ E(θ) by using three different variational algorithms: vanilla VQE (in blue), VQD (in red), and VQE/AC (in orange). The VQE performs the minimization of energy in search of the ground state of the triton. The VQD and VQE/AC, on the other hand, minimize the energy with an orthogonality constraint with respect to the ground state previously found. The lowest energy levels of the triton Hamiltonian obtained numerically are shown in gray. The ground state energy is shown in black.
FIG. 6. a) Estimated expectation values of the Hamiltonian, calculated on the first excited state (top) and on the ground state (bottom). The violin plot refers to the distribution obtained via Monte Carlo simulation sampling the parameters of the ansatz from a normal distribution centered in the best estimate and with standard deviation equal to the error. b) Numerical (and classical) energy levels of the triton Hamiltonian together with the quantum estimation.
FIG. 7. a) Success probability of the LCU method applied to the excitation operator responsible for the transition of the triton from the first excited (VQD) state into the ground state. b) the transition probability computed by the LCU. Both are studied for variable polarization angle ϑ ∈ [0, π]. With a solid radial line, we denote the angle corresponding to the maximum transition probability, measured to be ϑ ≃ 2.79 rad. c) Polar view of the transition probability associated with the excitation of the triton to the first excited state (VQD) extended for angles ϑ ∈ [0 • , 360 • ]. With a solid radial line, we denote the angle corresponding to the maximum transition probability, measured to be ϑ ≃ −20 • . d) The 3-dimensional plot of the same transition probability, obtained by adding an angle on the xy-plane and repeating the simulation. One can retrieve the twodimensional plot by considering the intersection with the plane in y = 0.
Simulation of a nuclear process on a quantum computer

August 2024

·

17 Reads

Quantum computers have proven to be effective in simulating many quantum systems. Simulating nuclear processes and state preparation poses significant challenges, even for traditional supercomputers. This study demonstrates the feasibility of a complete simulation of a nuclear transition, including the preparation of both ground and first excited states. To tackle the complexity of strong interactions between two and three nucleons, the states are modeled on the tritium nucleus. Both the initial and final states are represented using quantum circuits with variational quantum algorithms and inductive biases. Describing the spin-isospin states requires four qubits, and a parameterized quantum circuit that exploits a total of 16 parameters is initialized. The estimated energy has a relative error of approximately 2% for the ground state and about 10% for the first excited state of the system. The quantum computer simulation estimates the transition probability between the two states as a function of the dipole polarization angle. This work marks a first step towards leveraging digital quantum computers to simulate nuclear physics.


A Tutorial on the Use of Physics-Informed Neural Networks to Compute the Spectrum of Quantum Systems

July 2024

·

25 Reads

Quantum many-body systems are of great interest for many research areas, including physics, biology and chemistry. However, their simulation is extremely challenging, due to the exponential growth of the Hilbert space with the system size, making it exceedingly difficult to parameterize the wave functions of large systems by using exact methods. Neural networks and machine learning in general are a way to face this challenge. For instance, methods like Tensor networks and Neural Quantum States are being investigated as promising tools to obtain the wave function of a quantum mechanical system. In this tutorial, we focus on a particularly promising class of deep learning algorithms. We explain how to construct a Physics-Informed Neural Network (PINN) able to solve the Schr\"odinger equation for a given potential, by finding its eigenvalues and eigenfunctions. This technique is unsupervised, and utilizes a novel computational method in a manner that is barely explored. PINNs are a deep learning method that exploits Automatic Differentiation to solve Integro-Differential Equations in a mesh-free way. We show how to find both the ground and the excited states. The method discovers the states progressively by starting from the ground state. We explain how to introduce inductive biases in the loss to exploit further knowledge of the physical system. Such additional constraints allow for a faster and more accurate convergence. This technique can then be enhanced by a smart choice of collocation points in order to take advantage of the mesh-free nature of the PINN. The methods are made explicit by applying them to the infinite potential well and the particle in a ring, a challenging problem to be learned by an AI agent due to the presence of complex-valued eigenfunctions and degenerate states.


FIG. 1: Final wave function for the ground state. The blue line is the analytic wave function, while the red dots are the neural network prediction. In the left inset (in blue): the behavior of the loss throughout training, in logarithmic scale. In the right inset (in purple): fidelity throughout training
FIG. 3: (a) Ground state and (b) second excited state of the anharmonic oscillator for ω = 1; λ = 0.005, 0.16, 0.64, 1.28, and 10.24. The color of the wave functions goes from blue to red as λ increases. (c) Even and (d) Odd stationary states of the anharmonic oscillator for ω = 1; λ = 1.28. The number of dashes between each dot in the plot corresponds to the value of n of that eigenstate
FIG. 4: Wave function for the ground state (blue sparse dashed line) and first excited state (blue-green dashedotted line) of the double-well. The potential (black dense dashed line) has been centered and scaled for better visualization, and should only be taken qualitatively
FIG. 5: (a) Ground state and (b) second excited state of the double-well for r = −14; λ = {1, 1.1, 1.2, 1.4, 1.8, 2.6, 3.2}. The color of the wave functions goes from blue to red as λ increases. Furthermore, the spacing between each dash becomes tighter the higher λ.
FIG. 7: Energy levels of the anharmonic oscillator (blue or red circles) in comparison with the pure quartic oscillator (green squares), with exponential fits for the anharmonic eigenvalues at low λ (blue dashed line), anharmonic eigenvalues at high λ (red dashed line), pure eigenvalues at high λ (green dashed line). Both axes are in logarithmic scale. In the inset: A plot showing the values of λ corresponding to the intersections of the two exponentials corresponding to the low-λ and high-λ regimes for each energy level (λ c ). On the horizontal axis is the quantum number n, whereas and on the vertical axis is λ c . These values have also been fitted and show a roughly exponential behavior. The intersection points have also been signaled as black triangles in the main plot.
Addressing the Non-perturbative Regime of the Quantum Anharmonic Oscillator by Physics-Informed Neural Networks

May 2024

·

17 Reads

The use of deep learning in physical sciences has recently boosted the ability of researchers to tackle physical systems where little or no analytical insight is available. Recently, the Physics-Informed Neural Networks (PINNs) have been introduced as one of the most promising tools to solve systems of differential equations guided by some physically grounded constraints. In the quantum realm, such approach paves the way to a novel approach to solve the Schroedinger equation for non-integrable systems. By following an unsupervised learning approach, we apply the PINNs to the anharmonic oscillator in which an interaction term proportional to the fourth power of the position coordinate is present. We compute the eigenenergies and the corresponding eigenfunctions while varying the weight of the quartic interaction. We bridge our solutions to the regime where both the perturbative and the strong coupling theory work, including the pure quartic oscillator. We investigate systems with real and imaginary frequency, laying the foundation for novel numerical methods to tackle problems emerging in quantum field theory.


Comparing positive and negative statistics and their embeddings on the QPU. The left side of the Figure illustrates the multi‐embedding of positive statistics. Each numbered square represents a different configuration of hidden nodes, adjusted according to Equation (5). The white and black visible nodes have been fixed to 0 and 1, respectively. The numbers highlight the differences between the graphs. The multi‐embedding of negative statistics is illustrated on the right side of the Figure. In this scenario, each embedding maps the same graph in a different location on the QPU.
Likelihood test (a) and mean square error of the reconstruction (c) for the maximally parallelized classic (blue) and the full quantum (red) training algorithms, respectively. Corresponding results in terms of wall time are reported in (b) and (d). There, the green dashed curve represents the forecast performance that the quantum algorithm could achieve if fully parallelized. Note that the dashed green curve is a forecast obtained by appropriately rescaling (along the wall time axis) the experimental results presented in the red curve. Blue and red curves show the average of 10 distinct runs initialized with different random weights. The shaded regions represent the standard deviation. All plots are in log scale on the x‐axis and linear scale on the y‐axis.
a) The scatter plot shows the total number of qubits (right axis) and the average length of the qubits chains (left axis) required for embedding a single BM model of different sizes on the AQC. The x‐axis represents the total number of nodes in the model. b) time taken to generate 2000 samples for both positive and negative statistics relative to the model's size. In the classical case we kept k=2800$k=2800$ for all sizes.Note that the parallel classic scatter plot refers to the maximally parallelized approach where each sample is computed simultaneously on a NVIDIA A5000 GPU. With ”full quantum m.e.” we intend the samples extracted employing the multi embedding technique (parallel quantum annealing).
Sampling time as the problem size increases. The number of Gibbs cycles is set at a constant value of k=100$k=100$, and each algorithm extracts 100 samples. Quantum sampling exhibits a slow, nearly linear scaling pattern, whereas the classical algorithm begins at a higher time value and follows a power scaling trend of N1.66$N^{1.66}$ (see Table 2).
a) Moving from the left to the right, four images taken from the 4×4$4\times 4$ BAS datasetand b) their reconstruction: the same image with the four central pixels blurred; c) the images where the central pixels have been reconstructed by the full quantum BM at epoch 60; d) the the same blurred images reconstructed by the BM at epoch 140. Each picture contains 4 elements of the dataset.
Quantum Parallel Training of a Boltzmann Machine on an Adiabatic Quantum Computer

Despite the anticipated speed‐up of quantum computing, the achievement of a measurable advantage remains subject to ongoing debate. Adiabatic Quantum Computers (AQCs) are quantum devices designed to solve quadratic uncostrained binary optimization (QUBO) problems, but their intrinsic thermal noise can be leveraged to train computationally demanding machine learning algorithms such as the Boltzmann Machine (BM). Despite an asymptotic advantage is expected only for large networks, a limited quantum speed up can be already achieved on a small 16×1616×1616\times 16 BM is shown, by exploiting parallel adiabatic computation. This approach exhibits a 8.6‐fold improvement in wall time on the 4×44×44\times 4 Bars and Stripes dataset when compared to a parallelized classical Gibbs sampling method, which has never been outperformed before by quantum approaches.


Citations (57)


... Physics-informed Neural Networks [3] are a promising tool to discover and address the parametrization of a system governed by Partial Differential Equations (PDEs) or Integro-Differential Equations [4]. This rather new family of models has already shown success in a plethora of different fields [5,6], thanks to their ability to learn the implicit solution of a PDE even when given little to no data [7]. It allows to build models able to more efficiently simulate those systems and even to discover new physics [8]. ...

Reference:

A Tutorial on the Use of Physics-Informed Neural Networks to Compute the Spectrum of Quantum Systems
Addressing the non-perturbative regime of the quantum anharmonic oscillator by physics-informed neural networks

... Some of us have already developed both supervised [27][28][29] and unsupervised [30] machine learning methods, including supervised quantum machine learning to address ground state classification [31][32][33][34] of a physical system. Here, we aim to address the unsupervised solution of the Schrödinger [35,36] equation, to explore the potential of Physics-Informed Neural Networks (PINNs) for solving Partial Differential Equations (PDEs) as a paramount representative application to nontrivial physical phenomena [37]. ...

A Tutorial on the Use of Physics-Informed Neural Networks to Compute the Spectrum of Quantum Systems

Technologies

... Interestingly, in various contexts, controlling the degree of decoherence introduced into the reservoir has been shown to improve its memory capacity [41,47,54,63]. In this work, we build on this insight by applying it to a weak measurement-based protocol [43]. ...

Leveraging non-unital noise for gate-based quantum reservoir computing

... Eventually, it would be beneficial to assess the readiness of the QVAR subroutine for NISQ (Preskill 2018) devices, considering its logarithmic properties in both width and depth. Moreover, since QODA and HQFS depend on the amplitude estimation algorithm, which is not optimized for NISQ devices, we plan to study the efficacy of our algorithms that employ amplitude estimation implementations that are better suited for the NISQ era (Maronese et al. 2024). ...

The Quantum Amplitude Estimation Algorithms on Near-Term Devices: A Practical Guide

Quantum Reports

... A key motivation for our work is the establishment of a growing literature of complex quantum circuits resulting from the combination of underlying blocks (Fig. 1). For example, quantumbased Monte Carlo simulations in finance [9,10,11,12] or physics [13] combine the loading of a random distribution with some form of function processing, to then extract output via Quantum Amplitude Estimation [14]. More advanced examples [15] rely on quantum arithmetic [16], namely the ability to perform arithmetic calculations operating on basis states, as we shall discuss later. ...

The maximum-likelihood quantum amplitude estimation algorithm provides the best tradeoff between accuracy and circuit depth among quantum solutions for integral estimation
  • Citing Conference Paper
  • September 2023

... Furthermore, such a system has been consistently used in several contexts, most notably in solid-state physics to account for the hardening of the phonon dispersion relation [18][19][20], to investigate the formation of molecules with hydrogen bonds and their vibrational modes [21,22], and also for foundational studies such as nonlinear quantum mechanics [23,24], quantum chaos and control [25,26]. Some of us have already developed both supervised [27][28][29] and unsupervised [30] machine learning methods, including supervised quantum machine learning to address ground state classification [31][32][33][34] of a physical system. Here, we aim to address the unsupervised solution of the Schrödinger [35,36] equation, to explore the potential of Physics-Informed Neural Networks (PINNs) for solving Partial Differential Equations (PDEs) as a paramount representative application to nontrivial physical phenomena [37]. ...

A Max K-Cut Implementation for QAOA in the Measurement Based Quantum Computing Formalism
  • Citing Conference Paper
  • September 2023

... The quantum RBM has been used to address anomaly detection of IP traffic data, performing 64x faster than classic hardware in the inference [134]. More general machines called Boltzmann machines, based on a complete (not bipartite) graph, have also been addressed on an adiabatic quantum computer [135]. ...

Fast training of fully-connected Boltzmann Machines on an Adiabatic Quantum Computer
  • Citing Conference Paper
  • September 2023

... Furthermore, such a system has been consistently used in several contexts, most notably in solid-state physics to account for the hardening of the phonon dispersion relation [18][19][20], to investigate the formation of molecules with hydrogen bonds and their vibrational modes [21,22], and also for foundational studies such as nonlinear quantum mechanics [23,24], quantum chaos and control [25,26]. Some of us have already developed both supervised [27][28][29] and unsupervised [30] machine learning methods, including supervised quantum machine learning to address ground state classification [31][32][33][34] of a physical system. Here, we aim to address the unsupervised solution of the Schrödinger [35,36] equation, to explore the potential of Physics-Informed Neural Networks (PINNs) for solving Partial Differential Equations (PDEs) as a paramount representative application to nontrivial physical phenomena [37]. ...

Anomaly detection speed-up by quantum restricted Boltzmann machines

... However, previous implementations [6,7] do not meet the bandwidth requirements for fast and multiplexed readout, limiting their use to an individual device readout. In [8,9], integrator-based architectures have been introduced as an alternative achieving fast readout. However, these solutions do not enable the parallel readout of multiple quantum devices and only allow a sequential time multiplexing technique. ...

Fully Integrated Cryo-CMOS Spin-to-Digital Readout for Semiconductor Qubits
  • Citing Article
  • January 2023

IEEE Solid-State Circuits Letters

... Previous quantum computing research has linked the Rubik's cube to quantum walks for encryption [2], scrambling in many-body dynamics [3], Clifford gate synthesis [4], energy-level diagrams [5], and nonlocal entangled states for quantum secret sharing [6]. Additionally, Corli et al. [7] used reinforcement learning to solve the classical cube by embedding its group into a boson-fermion model. Similarly, the 15 puzzle has received attention due to it's relationship to itinerant ferromagnetism [8]. ...

Casting Rubik’s Group into a Unitary Representation for Reinforcement Learning

Journal of Physics Conference Series