[go: up one dir, main page]

Next Issue
Volume 6, September
Previous Issue
Volume 6, March
 
 

Quantum Rep., Volume 6, Issue 2 (June 2024) – 11 articles

Cover Story (view full-size image): We consider a toy model for studying monitored dynamics in many-body quantum systems. We study the stochastic Schrödinger equation from continuous monitoring of a random operator drawn from the Gaussian Unitary Ensemble. Due to unitary invariance, the dynamics of the eigenvalues of the density matrix decouples from that of the eigenvectors. We consider two regimes: in the presence of an extra dephasing term, the density matrix has a stationary distribution, matching the inverse Marchenko–Pastur distribution. In the case of perfect measurements, instead, purification eventually occurs, and we find an exact solution for the joint distribution of the eigenvalues at each time and arbitrary Hilbert space size, transitioning from a Coulomb gas to a universal regime. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Select all
Export citation of selected articles as:
45 pages, 697 KiB  
Article
The Computational Universe: Quantum Quirks and Everyday Reality, Actual Time, Free Will, the Classical Limit Problem in Quantum Loop Gravity and Causal Dynamical Triangulation
by Piero Chiarelli and Simone Chiarelli
Quantum Rep. 2024, 6(2), 278-322; https://doi.org/10.3390/quantum6020020 - 20 Jun 2024
Viewed by 358
Abstract
The simulation analogy presented in this work enhances the accessibility of abstract quantum theories, specifically the stochastic hydrodynamic model (SQHM), by relating them to our daily experiences. The SQHM incorporates the influence of fluctuating gravitational background, a form of dark energy, into quantum [...] Read more.
The simulation analogy presented in this work enhances the accessibility of abstract quantum theories, specifically the stochastic hydrodynamic model (SQHM), by relating them to our daily experiences. The SQHM incorporates the influence of fluctuating gravitational background, a form of dark energy, into quantum equations. This model successfully addresses key aspects of objective-collapse theories, including resolving the ‘tails’ problem through the definition of quantum potential length of interaction in addition to the De Broglie length, beyond which coherent Schrödinger quantum behavior and wavefunction tails cannot be maintained. The SQHM emphasizes that an external environment is unnecessary, asserting that the quantum stochastic behavior leading to wavefunction collapse can be an inherent property of physics in a spacetime with fluctuating metrics. Embedded in relativistic quantum mechanics, the theory establishes a coherent link between the uncertainty principle and the constancy of light speed, aligning seamlessly with finite information transmission speed. Within quantum mechanics submitted to fluctuations, the SQHM derives the indeterminacy relation between energy and time, offering insights into measurement processes impossible within a finite time interval in a truly quantum global system. Experimental validation is found in confirming the Lindemann constant for solid lattice melting points and the 4He transition from fluid to superfluid states. The SQHM’s self-consistency lies in its ability to describe the dynamics of wavefunction decay (collapse) and the measure process. Additionally, the theory resolves the pre-existing reality problem by showing that large-scale systems naturally decay into decoherent states stable in time. Continuing, the paper demonstrates that the physical dynamics of SQHM can be analogized to a computer simulation employing optimization procedures for realization. This perspective elucidates the concept of time in contemporary reality and enriches our comprehension of free will. The overall framework introduces an irreversible process impacting the manifestation of macroscopic reality at the present time, asserting that the multiverse exists solely in future states, with the past comprising the formed universe after the current moment. Locally uncorrelated projective decays of wavefunction, at the present time, function as a reduction of the multiverse to a single universe. Macroscopic reality, characterized by a foam-like consistency where microscopic domains with quantum properties coexist, offers insights into how our consciousness perceives dynamic reality. It also sheds light on the spontaneous emergence of gravity in discrete quantum spacetime evolution, and the achievement of the classical general relativity limit in quantum loop gravity and causal dynamical triangulation. The simulation analogy highlights a strategy focused on minimizing information processing, facilitating the universal simulation in solving its predetermined problem. From within, reality becomes the manifestation of specific physical laws emerging from the inherent structure of the simulation devised to address its particular issue. In this context, the reality simulation appears to employ an optimization strategy, minimizing information loss and data management in line with the simulation’s intended purpose. Full article
Show Figures

Figure 1

Figure 1
<p>The Universal ‘Pasta-Maker’.</p>
Full article ">
15 pages, 2571 KiB  
Article
Nitrogen-Related High-Spin Vacancy Defects in Bulk (SiC) and 2D (hBN) Crystals: Comparative Magnetic Resonance (EPR and ENDOR) Study
by Larisa Latypova, Fadis Murzakhanov, George Mamin, Margarita Sadovnikova, Hans Jurgen von Bardeleben and Marat Gafurov
Quantum Rep. 2024, 6(2), 263-277; https://doi.org/10.3390/quantum6020019 - 14 Jun 2024
Viewed by 862
Abstract
The distinct spin, optical, and coherence characteristics of solid-state spin defects in semiconductors have positioned them as potential qubits for quantum technologies. Both bulk and two-dimensional materials, with varying structural properties, can serve as crystalline hosts for color centers. In this study, we [...] Read more.
The distinct spin, optical, and coherence characteristics of solid-state spin defects in semiconductors have positioned them as potential qubits for quantum technologies. Both bulk and two-dimensional materials, with varying structural properties, can serve as crystalline hosts for color centers. In this study, we conduct a comparative analysis of the spin–optical, electron–nuclear, and relaxation properties of nitrogen-bound vacancy defects using electron paramagnetic resonance (EPR) and electron–nuclear double resonance (ENDOR) techniques. We examine key parameters of the spin Hamiltonian for the nitrogen vacancy (NV) center in 4H-SiC: D = 1.3 GHz, Azz = 1.1 MHz, and CQ = 2.53 MHz, as well as for the boron vacancy (VB) in hBN: D = 3.6 GHz, Azz = 85 MHz, and CQ = 2.11 MHz, and their dependence on the material matrix. The spin–spin relaxation times T2 (NV center: 50 µs and VB: 15 µs) are influenced by the local nuclear environment and spin diffusion while Rabi oscillation damping times depend on crystal size and the spatial distribution of microwave excitation. The ENDOR absorption width varies significantly among color centers due to differences in crystal structures. These findings underscore the importance of selecting an appropriate material platform for developing quantum registers based on high-spin color centers in quantum information systems. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) hBN crystals mounted on an aluminum substrate before electron irradiation. The distance between the black horizontal lines on the right is 5 mm; (<b>b</b>) Samples under study prepared for high-frequency part of the spectrometer. The characteristic dimensions of the samples and capillaries correspond to the internal diameter of the resonator to achieve the highest filling factor; (<b>c</b>) Bulk crystal (0.42 × 0.67 × 1.22 mm<sup>3</sup>) of silicon carbide under an optical microscope during the preparation of samples for experiments; (<b>d</b>) Bruker Elexsys E680 spectrometer operating at 94 GHz (W-band) equipped with helium flow cryostat; (<b>e</b>) Measurement setup diagram including the main blocks of the spectrometer for the photoinduced EPR and ENDOR.</p>
Full article ">Figure 2
<p>(<b>a</b>) ESE-EPR spectra for an <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>V</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> </mrow> </semantics></math> center in 4H-SiC (top half, red line) and a <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>V</mi> </mrow> <mrow> <mi>B</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> in hBN (bottom half, green line—experiment; blue solid line—simulation). The two insets at top show the detailed recorded low- and high-field components (red solid lines at 532 nm and navy color—“dark” mode) for structurally nonequivalent centers along with the corresponding simulation (blue dashed line). Yellow arrows indicate splittings between the components of the “zero-field splitting”; an asterisk (hBN) and a dot (SiC) indicate optically neutral signals both with spin = 1/2 from ionic compensators and interstitial defects, respectively, and are outside the scope of our study. (<b>b</b>) Schematic of spin polarization of color centers under optical excitation, where GS is a ground state, ES is an excited state, and MS is a metastable state. <span class="html-italic">D</span> denotes zero-field splitting.</p>
Full article ">Figure 3
<p>Dynamic characteristics of color centers obtained at <span class="html-italic">Temp.</span> = 10 K and optical excitation with λ = 532 nm. The upper part shows the curves of Rabi oscillations (blue dots) and transverse relaxation time (red solid line in the inset) for <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>V</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> </mrow> </semantics></math> centers in SiC, and for <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>V</mi> </mrow> <mrow> <mi>B</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> in hBN (Rabi oscillations are shown as green dots, transverse relaxations are shown as a solid dark green line). Red dashed lines for each center show decay traces of Rabi oscillations with characteristic damping time <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi mathvariant="normal">R</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>(<b>a</b>) EPR spectra of color centers at <span class="html-italic">Temp.</span> = 297 K, where the green solid line is a <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>V</mi> </mrow> <mrow> <mi>B</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> in hBN, the red line in the inset is an <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>V</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> </mrow> </semantics></math> center in SiC. The middle peak marked by a violet asterisk on the inset refers to an interstitial defect with electron spin <span class="html-italic">S</span> = ½. This spin center is independent of optical excitation of any wavelength (260–980 nm) and is beyond the scope of our study. (<b>b</b>) Spin–spin (<span class="html-italic">T</span><sub>2</sub>) or transverse relaxation and spin–lattice (<span class="html-italic">T</span><sub>1</sub>) or longitudinal relaxation (inset) curves for both color centers, where green is the <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>V</mi> </mrow> <mrow> <mi>B</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> in hBN, red is the <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>V</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> </mrow> </semantics></math> center in SiC.</p>
Full article ">Figure 5
<p>ENDOR spectra for SiC and hBN irradiated crystals. Hyperfine and quadrupole splitting values of the spin Hamiltonian (1) are shown in <a href="#quantumrep-06-00019-t004" class="html-table">Table 4</a>. The top inset shows individual NMR absorption lines for <sup>14</sup>N nuclei in the hBN and SiC crystal with significantly different line widths Δν.</p>
Full article ">
19 pages, 1115 KiB  
Article
Diversifying Investments and Maximizing Sharpe Ratio: A Novel Quadratic Unconstrained Binary Optimization Formulation
by Mirko Mattesi, Luca Asproni, Christian Mattia, Simone Tufano, Giacomo Ranieri, Davide Caputo and Davide Corbelletto
Quantum Rep. 2024, 6(2), 244-262; https://doi.org/10.3390/quantum6020018 - 27 May 2024
Viewed by 644
Abstract
The optimization of investment portfolios represents a pivotal task within the field of financial economics. Its objective is to identify asset combinations that meet specified criteria for return and risk. Traditionally, the maximization of the Sharpe Ratio, often achieved through quadratic programming, has [...] Read more.
The optimization of investment portfolios represents a pivotal task within the field of financial economics. Its objective is to identify asset combinations that meet specified criteria for return and risk. Traditionally, the maximization of the Sharpe Ratio, often achieved through quadratic programming, has constituted a popular approach for this purpose. However, real-world scenarios frequently necessitate more complex considerations, particularly in relation to portfolio diversification with a view to mitigating sector-specific risks and enhancing stability. The incorporation of diversification alongside the Sharpe Ratio into the optimization model creates a joint optimization task, which can be formulated as Quadratic Unconstrained Binary Optimization (QUBO) and addressed using quantum annealing or hybrid computing techniques. These techniques offer promising solutions. We present a novel QUBO formulation for this optimization, detailing its mathematical formulation and demonstrating its advantages over classical methods, particularly in handling diversification objectives. By leveraging available QUBO solvers and hybrid approaches, we explore the feasibility of handling large-scale problems while highlighting the importance of diversification in achieving robust portfolio performance. We finally elaborate on the results showing the trade-off between the observed values of the portfolio’s Sharpe Ratio and diversification, as a natural consequence of solving a multi-objective optimization problem. Full article
Show Figures

Figure 1

Figure 1
<p>Quantile−Quantile plots for simple returns and log-returns.</p>
Full article ">Figure 2
<p>Optimization runs on D-Wave’s QBSolv (CPU) and D-Wave’s Hybrid (QPU) of multiple QUBO instances as <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>2</mn> </msub> </semantics></math> varies. The plots report the behavior of Sharpe Ratio (blue line) and Diversification Entropy (red line), respectively, as <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>2</mn> </msub> </semantics></math> increases. All runs are feasible, considering the constraint <math display="inline"><semantics> <mrow> <msup> <mi>μ</mi> <mi>T</mi> </msup> <mi>y</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> satisfied if <math display="inline"><semantics> <mrow> <msup> <mi>μ</mi> <mi>T</mi> </msup> <mi>y</mi> </mrow> </semantics></math> is in a neighborhood of 1 (ref. to <a href="#sec3dot2-quantumrep-06-00018" class="html-sec">Section 3.2</a>), up to a factor equal to <math display="inline"><semantics> <mrow> <mn>2.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math>, which is given by multiplying the minimum discretization coefficient by the minimum expected return.</p>
Full article ">Figure 3
<p>Results provided by each combination of QUBO formulation and solver. The statistics are drawn from 10 feasible solutions with fixed values for the <math display="inline"><semantics> <mi>λ</mi> </semantics></math> coefficients. All solutions are feasible: for the Sharpe Ratio proxy formulation, feasibility is given by the sum of asset weights equal to 1, while for the Proposed Sharpe Ratio formulation, we consider the constraint satisfied if <math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mi>T</mi> </msup> <mi mathvariant="bold-italic">y</mi> </mrow> </semantics></math> is in a neighborhood of 1, up to a factor equal to <math display="inline"><semantics> <mrow> <mn>2.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math>. The colour of each violin plot refers to the average Sharpe Ratio value of the runs of each configuration. The darker the colour, the higher the average Sharpe Ratio value.</p>
Full article ">Figure 4
<p>Minimum, maximum, and mean (identified by the dot) in the number of assets selected over the 10 feasible solutions found. Left plot reports the statistics solving the two QUBO formulations using D-Wave Hybrid. Right plot, likewise, shows these results derived from the QBSolv solver.</p>
Full article ">Figure A1
<p>Percentage of feasible solutions obtained from 20 runs for each lambda configuration, for each combination of formulation and solver. Bar colour refers to the percentage value of run feasible for each configuration. The darker the colour, the higher the value.</p>
Full article ">
13 pages, 330 KiB  
Article
Wave Function and Information
by Leonardo Chiatti
Quantum Rep. 2024, 6(2), 231-243; https://doi.org/10.3390/quantum6020017 - 23 May 2024
Viewed by 706
Abstract
Two distinct measures of information, connected respectively to the amplitude and phase of the wave function of a particle, are proposed. There are relations between the time derivatives of these two measures and their gradients on the configuration space, which are equivalent to [...] Read more.
Two distinct measures of information, connected respectively to the amplitude and phase of the wave function of a particle, are proposed. There are relations between the time derivatives of these two measures and their gradients on the configuration space, which are equivalent to the wave equation. The information related to the amplitude measures the strength of the potential coupling of the particle (which is itself aspatial) with each volume of its configuration space, i.e., its tendency to participate in an interaction localized in a region of ordinary physical space corresponding to that volume. The information connected to the phase is that required to obtain the time evolution of the particle as a persistent entity starting from a random succession of bits. It can be considered as the information provided by conservation principles. The meaning of the so-called “quantum potential” in this context is briefly discussed. Full article
31 pages, 2408 KiB  
Article
A Dyson Brownian Motion Model for Weak Measurements in Chaotic Quantum Systems
by Federico Gerbino, Pierre Le Doussal, Guido Giachetti and Andrea De Luca
Quantum Rep. 2024, 6(2), 200-230; https://doi.org/10.3390/quantum6020016 - 16 May 2024
Cited by 1 | Viewed by 726
Abstract
We consider a toy model for the study of monitored dynamics in many-body quantum systems. We study the stochastic Schrödinger equation resulting from continuous monitoring with a rate Γ of a random Hermitian operator, drawn from the Gaussian unitary ensemble (GUE) at every [...] Read more.
We consider a toy model for the study of monitored dynamics in many-body quantum systems. We study the stochastic Schrödinger equation resulting from continuous monitoring with a rate Γ of a random Hermitian operator, drawn from the Gaussian unitary ensemble (GUE) at every time t. Due to invariance by unitary transformations, the dynamics of the eigenvalues {λα}α=1n of the density matrix decouples from that of the eigenvectors, and is exactly described by stochastic equations that we derive. We consider two regimes: in the presence of an extra dephasing term, which can be generated by imperfect quantum measurements, the density matrix has a stationary distribution, and we show that in the limit of large size n it matches with the inverse-Marchenko–Pastur distribution. In the case of perfect measurements, instead, purification eventually occurs and we focus on finite-time dynamics. In this case, remarkably, we find an exact solution for the joint probability distribution of λ’s at each time t and for each size n. Two relevant regimes emerge: at short times tΓ=O(1), the spectrum is in a Coulomb gas regime, with a well-defined continuous spectral distribution in the n limit. In that case, all moments of the density matrix become self-averaging and it is possible to exactly characterize the entanglement spectrum. In the limit of large times tΓ=O(n), one enters instead a regime in which the eigenvalues are exponentially separated log(λα/λβ)=O(Γt/n), but fluctuations O(Γt/n) play an essential role. We are still able to characterize the asymptotic behaviors of the entanglement entropy in this regime. Full article
(This article belongs to the Special Issue Exclusive Feature Papers of Quantum Reports in 2024–2025)
Show Figures

Figure 1

Figure 1
<p>Inverse-Marchenko–Pastur distribution. The spectral density <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>(</mo> <mover accent="true"> <mi>λ</mi> <mo stretchy="false">˜</mo> </mover> <mo>)</mo> </mrow> </semantics></math> for the rescaled eigenvalues <math display="inline"><semantics> <mover accent="true"> <mi>λ</mi> <mo stretchy="false">˜</mo> </mover> </semantics></math> defined in Equation (<a href="#FD28-quantumrep-06-00016" class="html-disp-formula">28</a>) is shown. In the large <span class="html-italic">n</span> limit, it takes the inverse-Marchenko–Pastur form given in Equation (<a href="#FD35-quantumrep-06-00016" class="html-disp-formula">35</a>). In both plots, the orange line displays the theoretical curve, whereas the blue histogram bars are computed after a numerical simulation (with <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>) of the weak measurement protocol. (<b>a</b>) The spectral density at <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math> in the range <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>4</mn> <mo>/</mo> <mi>x</mi> <mo>]</mo> </mrow> </semantics></math>. At small <span class="html-italic">x</span>, most eigenvalues are located in vicinity of <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>λ</mi> <mo stretchy="false">˜</mo> </mover> <mo>−</mo> </msub> <mo stretchy="false">→</mo> <mn>1</mn> <mo>/</mo> <mn>4</mn> </mrow> </semantics></math>, as an effect of purification. (<b>b</b>) The spectral density at <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>5.0</mn> </mrow> </semantics></math> in its domain <math display="inline"><semantics> <mrow> <mo>[</mo> <msub> <mover accent="true"> <mi>λ</mi> <mo stretchy="false">˜</mo> </mover> <mo>−</mo> </msub> <mo>,</mo> <msub> <mover accent="true"> <mi>λ</mi> <mo stretchy="false">˜</mo> </mover> <mo>+</mo> </msub> <mo>]</mo> </mrow> </semantics></math>. For larger values of <span class="html-italic">x</span>, the rescaled eigenvalues take finite values around their average <math display="inline"><semantics> <mrow> <mrow> <mo>〈</mo> <mover accent="true"> <mi>λ</mi> <mo stretchy="false">˜</mo> </mover> <mo>〉</mo> </mrow> <mo>=</mo> <mn>2</mn> <mo>/</mo> <mi>x</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Short time behavior. The density <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mi>τ</mi> </msub> <mrow> <mo>(</mo> <mi>w</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> in the short time <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mi>τ</mi> <mo>/</mo> <mn>4</mn> <mo>Γ</mo> <mo>∼</mo> <mn>1</mn> <mo>/</mo> <mo>Γ</mo> </mrow> </semantics></math> regime, with <math display="inline"><semantics> <mi>τ</mi> </semantics></math> finite. <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mi>τ</mi> </msub> <mrow> <mo>(</mo> <mi>w</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> features a crossover between a semi-circle at small <math display="inline"><semantics> <mi>τ</mi> </semantics></math> and a square distribution at larger <math display="inline"><semantics> <mi>τ</mi> </semantics></math>. (<b>a</b>) The small-<math display="inline"><semantics> <mi>τ</mi> </semantics></math> semi-circle distribution is displayed for <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> and increasing <span class="html-italic">n</span> from <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math>. The red solid line shows the theoretical <math display="inline"><semantics> <mrow> <mi>n</mi> <mo stretchy="false">→</mo> <mo>∞</mo> </mrow> </semantics></math> curve. (<b>b</b>) The large-<math display="inline"><semantics> <mi>τ</mi> </semantics></math> square distribution is shown for <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> and increasing <span class="html-italic">n</span>, with the red solid line showing the <math display="inline"><semantics> <mrow> <mi>n</mi> <mo stretchy="false">→</mo> <mo>∞</mo> </mrow> </semantics></math> curve. (<b>c</b>) The crossover from semi-circle towards square distribution is shown for increasing <math display="inline"><semantics> <mi>τ</mi> </semantics></math> from <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>. All solid lines represent the theoretical density <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mi>e</mi> </msub> <mi>w</mi> <mo>·</mo> <msub> <mi>f</mi> <mi>τ</mi> </msub> <mrow> <mo>(</mo> <mi>w</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> on the rescaled <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>/</mo> <msub> <mi>w</mi> <mi>e</mi> </msub> </mrow> </semantics></math> axis, where <math display="inline"><semantics> <msub> <mi>w</mi> <mi>e</mi> </msub> </semantics></math> is the edge coordinate for <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mi>τ</mi> </msub> <mrow> <mo>(</mo> <mi>w</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> in the <math display="inline"><semantics> <mrow> <mi>n</mi> <mo stretchy="false">→</mo> <mo>∞</mo> </mrow> </semantics></math> limit.</p>
Full article ">Figure 3
<p>Long-time ranked diffusion. Long-time dynamics of the averaged eigenvalues <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>, whereas the maximum eigenvalue tends to one as <math display="inline"><semantics> <mrow> <mrow> <mo>〈</mo> <mrow> <msub> <mi>λ</mi> <mn>10</mn> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> <mo>〉</mo> </mrow> <mo>∼</mo> <mn>1</mn> <mo>−</mo> <msup> <mi>e</mi> <mrow> <mo>−</mo> <mn>4</mn> <mi>t</mi> </mrow> </msup> </mrow> </semantics></math> as time increases, the first <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> eigenvalues are equispaced in logarithmic scale, i.e., <math display="inline"><semantics> <mrow> <mrow> <mo>〈</mo> <mrow> <msub> <mi>λ</mi> <mi>α</mi> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> <mo>〉</mo> </mrow> <mo>/</mo> <mrow> <mo>〈</mo> <mrow> <msub> <mi>λ</mi> <mi>β</mi> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> <mo>〉</mo> </mrow> <mo>=</mo> <msup> <mi>e</mi> <mrow> <mn>4</mn> <mo>(</mo> <mi>α</mi> <mo>−</mo> <mi>β</mi> <mo>)</mo> <mi>t</mi> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">
16 pages, 1173 KiB  
Article
Fisher Information for a System Composed of a Combination of Similar Potential Models
by Clement Atachegbe Onate, Ituen B. Okon, Edwin Samson Eyube, Ekwevugbe Omugbe, Kizito O. Emeje, Michael C. Onyeaju, Olumide O. Ajani and Jacob A. Akinpelu
Quantum Rep. 2024, 6(2), 184-199; https://doi.org/10.3390/quantum6020015 - 13 May 2024
Viewed by 575
Abstract
The solutions to the radial Schrödinger equation for a pseudoharmonic potential and Kratzer potential have been studied separately in the past. Despite different reports on the Kratzer potential, the fundamental theoretical quantities such as Fisher information have not been reported. In this study, [...] Read more.
The solutions to the radial Schrödinger equation for a pseudoharmonic potential and Kratzer potential have been studied separately in the past. Despite different reports on the Kratzer potential, the fundamental theoretical quantities such as Fisher information have not been reported. In this study, we obtain the solution to the radial Schrödinger equation for the combination of the pseudoharmonic and Kratzer potentials in the presence of a constant-dependent potential, utilizing the concepts and formalism of the supersymmetric and shape invariance approach. The position expectation value and momentum expectation value are calculated employing the Hellmann–Feynman Theory. These expectation values are then used to calculate the Fisher information for both position and momentum spaces in both the absence and presence of the constant-dependent potential. The results obtained revealed that the presence of the constant-dependent potential leads to an increase in the energy eigenvalue, as well as in the position and momentum expectation values. Additionally, the constant-dependent potential increases the Fisher information for both position and momentum spaces. Furthermore, the product of the position expectation value and the momentum expectation value, along with the product of the Fisher information, satisfies both Fisher’s inequality and Cramer–Rao’s inequality. Full article
Show Figures

Figure 1

Figure 1
<p>Variation in energy against the quantum number with <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mo>ℏ</mo> <mo>=</mo> <mn>1</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mi>e</mi> </msub> <mo>=</mo> <mn>0.25</mn> <mo>,</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mi>e</mi> </msub> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> for three values of the angular momentum number.</p>
Full article ">Figure 2
<p>Fisher information in momentum <math display="inline"><semantics> <mrow> <mi>I</mi> <mo stretchy="false">(</mo> <mi>γ</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> (<b>a</b>) and Fisher information in position <math display="inline"><semantics> <mrow> <mi>I</mi> <mo stretchy="false">(</mo> <mi>ρ</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> (<b>b</b>) against the dissociation energy with <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mrow> <mstyle mathvariant="italic"> <mi mathvariant="script">l</mi> </mstyle> </mrow> <mo>=</mo> <mo>ℏ</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mi>e</mi> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Fisher information in momentum <math display="inline"><semantics> <mrow> <mi>I</mi> <mo stretchy="false">(</mo> <mi>γ</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> (<b>a</b>) and Fisher information in position <math display="inline"><semantics> <mrow> <mi>I</mi> <mo stretchy="false">(</mo> <mi>ρ</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> (<b>b</b>) against the angular momentum with <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mrow> <mstyle mathvariant="italic"> <mi mathvariant="script">l</mi> </mstyle> </mrow> <mo>=</mo> <mo>ℏ</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mi>e</mi> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>.</p>
Full article ">
12 pages, 4665 KiB  
Article
Spectral Analysis of Proton Eigenfunctions in Crystalline Environments
by Luca Gamberale and Giovanni Modanese
Quantum Rep. 2024, 6(2), 172-183; https://doi.org/10.3390/quantum6020014 - 6 May 2024
Viewed by 618
Abstract
The Schrödinger equation and Bloch theorem are applied to examine a system of protons confined within a periodic potential, accounting for deviations from ideal harmonic behavior due to real-world conditions like truncated and non-quadratic potentials, in both one-dimensional and three-dimensional scenarios. Numerical computation [...] Read more.
The Schrödinger equation and Bloch theorem are applied to examine a system of protons confined within a periodic potential, accounting for deviations from ideal harmonic behavior due to real-world conditions like truncated and non-quadratic potentials, in both one-dimensional and three-dimensional scenarios. Numerical computation of the energy spectrum of bound eigenfunctions in both cases reveals intriguing structures, including bound states with degeneracy matching the site number Nw, reminiscent of a finite harmonic oscillator spectrum. In contrast to electronic energy bands, the proton system displays a greater number of possible bound states due to the significant mass of protons. Extending previous research, this study rigorously determines the constraints on the energy gap and oscillation amplitude of the previously identified coherent states. The deviations in energy level spacing identified in the computed spectrum, leading to the minor splitting of electromagnetic modes, are analyzed and found not to hinder the onset of coherence. Finally, a more precise value of the energy gap is determined for the proton coherent states, ensuring their stability against thermal decoherence up to the melting temperature of the hosting metal. Full article
(This article belongs to the Special Issue Exclusive Feature Papers of Quantum Reports in 2024–2025)
Show Figures

Figure 1

Figure 1
<p>Periodic 1D potential with 40 sites with lattice spacing <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mfrac> <mi>a</mi> <msqrt> <mn>2</mn> </msqrt> </mfrac> <mo>=</mo> <mn>2.5</mn> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>8</mn> </mrow> </msup> <mspace width="4.pt"/> <mi>cm</mi> <mo>=</mo> <mn>1.27</mn> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> <mspace width="4.pt"/> <msup> <mi>eV</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math> along the base vector <math display="inline"><semantics> <msub> <mover accent="true"> <mi>a</mi> <mo stretchy="false">→</mo> </mover> <mn>1</mn> </msub> </semantics></math>.</p>
Full article ">Figure 2
<p>One particle energy spectrum calculated using the parameters of <a href="#quantumrep-06-00014-t001" class="html-table">Table 1</a>. Note the harmonic oscillator-like energy spacing of the bound states and their degeneracy, equal to the number of lattice sites <math display="inline"><semantics> <msub> <mi>N</mi> <mi>w</mi> </msub> </semantics></math>. The free eigenvalues have a Brillouin-like structure. Different bans correspond to different colors. The finite number of wells causes intermediate eigenvalues to appear between adjacent bands (see sub-figure).</p>
Full article ">Figure 3
<p>Eigenfunctions of the bound states computed by direct diagonalization of Equation (<a href="#FD8-quantumrep-06-00014" class="html-disp-formula">8</a>) with the parameters of <a href="#quantumrep-06-00014-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Linear fit of the photon energies associated with the dipolar transitions between adjacent bound energy levels (the last two points to the right are excluded from the fit).</p>
Full article ">Figure 5
<p>Eigenvalues of the radial equation for all quantum numbers of the bound states. The confirmation of independence from angular momentum, as displayed by the eigenstates of the full harmonic oscillator, persists even for eigenvalues very close to the free spectrum.</p>
Full article ">
16 pages, 497 KiB  
Article
Measuring the Density Matrix of Quantum-Modeled Cognitive States
by Wendy Xiomara Chavarría-Garza, Osvaldo Aquines-Gutiérrez, Ayax Santos-Guevara, Humberto Martínez-Huerta, Jose Ruben Morones-Ibarra and Jonathan Rincon Saucedo
Quantum Rep. 2024, 6(2), 156-171; https://doi.org/10.3390/quantum6020013 - 27 Apr 2024
Viewed by 806
Abstract
Inspired by the principles of quantum mechanics, we constructed a model of students’ misconceptions about heat and temperature, conceptualized as a quantum system represented by a density matrix. Within this framework, the presence or absence of misconceptions is delineated as pure states, while [...] Read more.
Inspired by the principles of quantum mechanics, we constructed a model of students’ misconceptions about heat and temperature, conceptualized as a quantum system represented by a density matrix. Within this framework, the presence or absence of misconceptions is delineated as pure states, while the probability of mixed states is also considered, providing valuable insights into students’ cognition based on the mental models they employ when holding misconceptions. Using the analysis model previously employed by Lei Bao and Edward Redish, we represented these results in a density matrix. In our research, we utilized the Zeo and Zadnik Thermal Concept Evaluation among 282 students from a private university in Northeast Mexico. Our objective was to extract information from the analysis of multiple-choice questions designed to explore preconceptions, offering valuable educational insights beyond the typical Correct–Incorrect binary analysis of classical systems. Our findings reveal a probability of 0.72 for the appearance of misconceptions, 0.28 for their absence, and 0.43 for mixed states, while no significant disparities were observed based on gender or scholarship status, a notable difference was observed among programs (p < 0.05). These results are consistent with the previous literature, confirming a prevalence of misconceptions within the student population. Full article
Show Figures

Figure 1

Figure 1
<p>Results of similar tests (non-appearance of misconception in TCE, CRT and FCI) [<a href="#B19-quantumrep-06-00013" class="html-bibr">19</a>,<a href="#B20-quantumrep-06-00013" class="html-bibr">20</a>,<a href="#B26-quantumrep-06-00013" class="html-bibr">26</a>,<a href="#B29-quantumrep-06-00013" class="html-bibr">29</a>,<a href="#B30-quantumrep-06-00013" class="html-bibr">30</a>,<a href="#B31-quantumrep-06-00013" class="html-bibr">31</a>,<a href="#B32-quantumrep-06-00013" class="html-bibr">32</a>,<a href="#B33-quantumrep-06-00013" class="html-bibr">33</a>]. Means and confidence intervals (CIs) for the measurements outlined in <a href="#quantumrep-06-00013-t003" class="html-table">Table 3</a> at a 95% confidence level (CL). Solid colored lines depict the average for each test (TCE, CRT, FCT). The gray shaded area represents this measurement, accounting for both stochastic and systematic errors as described in Equation (<a href="#FD4-quantumrep-06-00013" class="html-disp-formula">4</a>).</p>
Full article ">
9 pages, 259 KiB  
Article
A Normalization Condition for the Probability Current in Some Remarkable Cases
by Antonio Feoli, Elmo Benedetto and Antonella Lucia Iannella
Quantum Rep. 2024, 6(2), 147-155; https://doi.org/10.3390/quantum6020012 - 23 Apr 2024
Viewed by 715
Abstract
Starting from the dynamics of a bouncing ball in classical and quantum regime, we have suggested in a previous paper to add an arbitrary function of time to the standard expression of the probability current in quantum mechanics. In this paper, we suggest [...] Read more.
Starting from the dynamics of a bouncing ball in classical and quantum regime, we have suggested in a previous paper to add an arbitrary function of time to the standard expression of the probability current in quantum mechanics. In this paper, we suggest a way to determine this function: imposing a suitable normalization condition. The application of our proposal to the case of the harmonic oscillator is discussed. Full article
5 pages, 199 KiB  
Editorial
The Many-Worlds Interpretation of Quantum Mechanics: Current Status and Relation to Other Interpretations
by Lev Vaidman
Quantum Rep. 2024, 6(2), 142-146; https://doi.org/10.3390/quantum6020011 - 18 Apr 2024
Viewed by 1218
Abstract
This is a preface to a Special Issue of Quantum Reports devoted to the results of the workshop “The Many-Worlds Interpretation of Quantum Mechanics: Current Status and Relation to Other Interpretations” [...] Full article
(This article belongs to the Special Issue The Many-Worlds Interpretation of Quantum Mechanics)
8 pages, 320 KiB  
Communication
Continuum Limit of the Green Function in Scaled Affine φ44 Quantum Euclidean Covariant Relativistic Field Theory
by Riccardo Fantoni
Quantum Rep. 2024, 6(2), 134-141; https://doi.org/10.3390/quantum6020010 - 14 Apr 2024
Viewed by 1285
Abstract
Through path integral Monte Carlo computer experiments, we prove that the affine quantization of the φ44-scaled Euclidean covariant relativistic scalar field theory is a valid quantum field theory with a well-defined continuum limit of the one- and two-point functions. Affine [...] Read more.
Through path integral Monte Carlo computer experiments, we prove that the affine quantization of the φ44-scaled Euclidean covariant relativistic scalar field theory is a valid quantum field theory with a well-defined continuum limit of the one- and two-point functions. Affine quantization leads to a completely satisfactory quantization of field theories in situations involving scaled behavior, leading to an unexpected term, 2/φ2, which arises only in the quantum aspects. Full article
Show Figures

Figure 1

Figure 1
<p>Two-point function <math display="inline"><semantics> <mrow> <mi>D</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mo>∑</mo> <mi>x</mi> </msub> <mrow> <mo>〈</mo> <mi>φ</mi> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mi>φ</mi> <mrow> <mo>(</mo> <mi>x</mi> <mo>+</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>〉</mo> </mrow> <mo>/</mo> <msup> <mi>N</mi> <mi>n</mi> </msup> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>3</mn> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mi>g</mi> <mo>=</mo> <mi>L</mi> <mo>=</mo> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mi>L</mi> <mo>/</mo> <mi>a</mi> <mo>=</mo> <mn>4</mn> <mo>,</mo> <mtext> </mtext> <mn>7</mn> <mo>,</mo> <mtext> </mtext> <mn>13</mn> <mo>,</mo> <mtext> </mtext> <mn>25</mn> </mrow> </semantics></math>. In our PIMC simulations, we use Equations (<a href="#FD4-quantumrep-06-00010" class="html-disp-formula">4</a>) and (<a href="#FD5-quantumrep-06-00010" class="html-disp-formula">5</a>).</p>
Full article ">Figure 2
<p>Two-point connected function <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mi>c</mi> </msub> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>=</mo> <mi>D</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>−</mo> <msup> <mi>V</mi> <mn>2</mn> </msup> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>3</mn> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mi>g</mi> <mo>=</mo> <mi>L</mi> <mo>=</mo> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mi>L</mi> <mo>/</mo> <mi>a</mi> <mo>=</mo> <mn>4</mn> <mo>,</mo> <mtext> </mtext> <mn>7</mn> <mo>,</mo> <mtext> </mtext> <mn>13</mn> <mo>,</mo> <mtext> </mtext> <mn>25</mn> </mrow> </semantics></math>. In our PIMC simulations, we use Equations (<a href="#FD4-quantumrep-06-00010" class="html-disp-formula">4</a>) and (<a href="#FD5-quantumrep-06-00010" class="html-disp-formula">5</a>).</p>
Full article ">Figure 3
<p>Two-point connected function <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mi>c</mi> </msub> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>=</mo> <mi>D</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>−</mo> <msup> <mi>V</mi> <mn>2</mn> </msup> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>3</mn> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mi>L</mi> <mo>=</mo> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <msqrt> <mi>N</mi> </msqrt> <mo>/</mo> <mi>L</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mi>L</mi> <mo>/</mo> <mi>a</mi> <mo>=</mo> <mn>4</mn> <mo>,</mo> <mtext> </mtext> <mn>7</mn> <mo>,</mo> <mtext> </mtext> <mn>13</mn> <mo>,</mo> <mtext> </mtext> <mn>25</mn> </mrow> </semantics></math>. In our PIMC simulations, we use Equations (<a href="#FD4-quantumrep-06-00010" class="html-disp-formula">4</a>) and (<a href="#FD5-quantumrep-06-00010" class="html-disp-formula">5</a>).</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop