[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,210)

Search Parameters:
Keywords = quantum integrability

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 4800 KiB  
Article
Design and Analysis of Compact High–Performance Lithium–Niobate Electro–Optic Modulator Based on a Racetrack Resonator
by Zixin Chen, Jianping Li, Weiqin Zheng, Hongkang Liu, Quandong Huang, Ya Han and Yuwen Qin
Photonics 2025, 12(1), 85; https://doi.org/10.3390/photonics12010085 (registering DOI) - 17 Jan 2025
Viewed by 106
Abstract
With the ever-growing demand for high-speed optical communications, microwave photonics, and quantum key distribution systems, compact electro-optic (EO) modulators with high extinction ratios, large bandwidth, and high tuning efficiency are urgently pursued. However, most integrated lithium–niobate (LN) modulators cannot achieve these high performances [...] Read more.
With the ever-growing demand for high-speed optical communications, microwave photonics, and quantum key distribution systems, compact electro-optic (EO) modulators with high extinction ratios, large bandwidth, and high tuning efficiency are urgently pursued. However, most integrated lithium–niobate (LN) modulators cannot achieve these high performances simultaneously. In this paper, we propose an improved theoretical model of a chip-scale electro-optic (EO) microring modulator (EO-MRM) based on X-cut lithium–niobate-on-insulator (LNOI) with a hybrid architecture consisting of a 180-degree Euler bend in the coupling region, double-layer metal electrode structure, and ground–signal–signal–ground (G-S-S-G) electrode configuration, which can realize highly comprehensive performance and a compact footprint. After parameter optimization, the designed EO-MRM exhibited an extinction ratio of 38 dB. Compared to the structure without Euler bends, the increase was 35 dB. It also had a modulation bandwidth of 29 GHz and a tunability of 8.24 pm/V when the straight waveguide length was 100 μm. At the same time, the proposed device footprint was 1.92 × 104 μm2. The proposed MRM model provides an efficient solution to high-speed optical communication systems and microwave photonics, which is helpful for the fabrication of high-performance and multifunctional photonic integrated devices. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A schematic diagram of the proposed racetrack resonator with a double-layer electrode. Inset: the cross-section of coupling area. (<b>b</b>) A top view of the racetrack microring resonator. (<b>c</b>) The optical mode field and intensity distribution of the Euler bend with a waveguide width of 0.8 µm, simulated by FDTD.</p>
Full article ">Figure 2
<p>(<b>a</b>) Lumerical MODE simulation of the fundamental TE<sub>0</sub> optical mode of the waveguide. (<b>b</b>) The calculated optical effective index of the waveguide.</p>
Full article ">Figure 3
<p>(<b>a</b>) The coupling coefficient <span class="html-italic">κ</span><sup>2</sup> and (<b>b</b>) the transmission coefficient <span class="html-italic">t</span><sup>2</sup> vary with w<sub>gap</sub> in the coupling region at the wavelength of 1550 nm.</p>
Full article ">Figure 4
<p>(<b>a</b>) The coupling coefficient <span class="html-italic">κ</span><sup>2</sup> and (<b>b</b>) the transmission coefficient <span class="html-italic">t</span><sup>2</sup> vary with w<sub>1</sub> in the coupling region at the wavelength of 1550 nm.</p>
Full article ">Figure 5
<p>The BW and <span class="html-italic">Q</span> factor performances with the variation in <span class="html-italic">Lc</span> of the resonator.</p>
Full article ">Figure 6
<p>(<b>a</b>) The coupling and transmission coefficients with a variation in wavelength, when w<sub>gap</sub> = 0.7 μm and w<sub>1</sub> = 0.6 μm. (<b>b</b>) Transmission spectrum of the resonator with different bends used in the coupling region at the wavelength of 1550 nm.</p>
Full article ">Figure 7
<p>(<b>a</b>) A top view of the proposed tunable racetrack resonator with double-layer electrodes. (<b>b</b>) The simulated TE optical mode field profile at 1550 nm and the electric field between the double-layer electrodes. Here, the TFLN waveguide was formed by a 300 nm × 0.8 µm LN loading ridge. (<b>c</b>) A schematic of a unit cell of the electrode structure. (<b>d</b>) The simulation result of the influence of h and d on metal loss.</p>
Full article ">Figure 8
<p>Metal loss analysis for different electrode designs. (<b>a</b>) Metal electrodes were placed directly on the waveguide. (<b>b</b>) A 2.8 μm-wide layer of SiO<sub>2</sub> was added between the double metal electrode and the waveguide.</p>
Full article ">Figure 9
<p>(<b>a</b>) The simulated transmission spectrum of the TE mode of the passive racetrack resonator. (<b>b</b>) The detailed spectrum at 1550.118 nm. (<b>c</b>) The spectrum under different voltages of the TE mode at 1550.118 nm. (<b>d</b>) Resonant wavelength shifts as a function of the applied voltage.</p>
Full article ">
26 pages, 2855 KiB  
Article
Photokinetics of Photothermal Reactions
by Mounir Maafi
Molecules 2025, 30(2), 330; https://doi.org/10.3390/molecules30020330 - 15 Jan 2025
Viewed by 220
Abstract
Photothermal reactions, involving both photochemical and thermal reaction steps, are the most abundant sequences in photochemistry. The derivation of their rate laws is standardized, but the integration of these rate laws has not yet been achieved. Indeed, the field still lacks integrated rate [...] Read more.
Photothermal reactions, involving both photochemical and thermal reaction steps, are the most abundant sequences in photochemistry. The derivation of their rate laws is standardized, but the integration of these rate laws has not yet been achieved. Indeed, the field still lacks integrated rate laws for the description of these reactions’ behavior and/or identification of their reaction order. This made difficult a comprehensive account of the photokinetics of photothermal reactions, which created a gap in knowledge. This gap is addressed in the present paper by introducing an unprecedented general model equation capable of mapping out the kinetic traces of such reactions when exposed to light or in the dark. The integrated rate law model equation also applies when the reactive medium is exposed to either monochromatic or polychromatic light irradiation. The validity of the model equation was established against simulated data obtained by a fourth-order Runge–Kutta method. It was then used to describe and quantify several situations of photothermal reactions, such as the effects of initial concentration, spectator molecules, and incident radiation intensity, and the impact of the latter on the photonic yield. The model equation facilitated a general elucidation method to determine the intrinsic reaction parameters (quantum yields and absorptivities of the reactive species) for any photothermal mechanism whose number of species is known. This paper contributes to rationalizing photokinetics along the same general guidelines adopted in chemical kinetics. Full article
Show Figures

Figure 1

Figure 1
<p>Electronic spectra (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ε</mi> </mrow> <mrow> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mi>j</mi> </mrow> </msub> </mrow> <mrow> <msub> <mrow> <mi>λ</mi> </mrow> <mrow> <mi>i</mi> <mi>r</mi> <mi>r</mi> </mrow> </msub> </mrow> </msubsup> </mrow> </semantics></math>), lamp profile (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>P</mi> </mrow> <mrow> <mn>0</mn> </mrow> <mrow> <msub> <mrow> <mi>λ</mi> </mrow> <mrow> <mi>i</mi> <mi>r</mi> <mi>r</mi> </mrow> </msub> </mrow> </msubsup> </mrow> </semantics></math>), (<b>a</b>) and quantum yield wavelength-dependent patterns (<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Φ</mi> </mrow> </semantics></math>), (<b>b</b>) of a tetramolecular reaction <math display="inline"><semantics> <mrow> <mi>X</mi> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> <mfenced separators="|"> <mrow> <mn>6</mn> <mi mathvariant="sans-serif">Φ</mi> <mo>,</mo> <mn>5</mn> <mi>k</mi> </mrow> </mfenced> </mrow> </semantics></math>, <a href="#molecules-30-00330-sch001" class="html-scheme">Scheme 1</a>, proposed for 3<span class="html-italic">H</span>-naphthopyrans [<a href="#B51-molecules-30-00330" class="html-bibr">51</a>].</p>
Full article ">Figure 2
<p>Excellent fittings of RK-calculated trace (circles) by the adequate (lines) Equation (6) for each species of the <math display="inline"><semantics> <mrow> <mi>X</mi> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> <mfenced separators="|"> <mrow> <mn>6</mn> <mi mathvariant="sans-serif">Φ</mi> <mo>,</mo> <mn>5</mn> <mi>k</mi> </mrow> </mfenced> </mrow> </semantics></math> reaction (<a href="#molecules-30-00330-sch001" class="html-scheme">Scheme 1</a>, <a href="#molecules-30-00330-f001" class="html-fig">Figure 1</a>).</p>
Full article ">Figure 3
<p>An excellent linear relationship is found between the values of <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>h</mi> <mi>e</mi> <mi>o</mi> <mi>:</mi> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>T</mi> <mi>h</mi> <mi>e</mi> <mi>o</mi> <mi>:</mi> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>X</mi> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>h</mi> <mi>e</mi> <mi>o</mi> <mi>:</mi> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>A</mi> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math>) against the respective <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>K</mi> <mi>:</mi> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>F</mi> <mi>i</mi> <mi>t</mi> <mi>:</mi> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>X</mi> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>A</mi> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math>). (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>A</mi> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math>, which have generally much larger values than <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>X</mi> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math>, where scaled down by an adequate multiplicative factor). The data reported here belong to different reactive systems and experimental conditions.</p>
Full article ">Figure 4
<p>Evolution towards more negative values of the initial rate (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>r</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>X</mi> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> </mrow> </semantics></math>) when the initial concentration (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>C</mi> </mrow> <mrow> <mi>X</mi> </mrow> <mrow> <mi>L</mi> <mi>p</mi> <mo>,</mo> <mo>∆</mo> <mi>λ</mi> <mo>,</mo> <mi>T</mi> </mrow> </msubsup> <mfenced separators="|"> <mrow> <mn>0</mn> </mrow> </mfenced> </mrow> </semantics></math>) increases for an <math display="inline"><semantics> <mrow> <mi>X</mi> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> <mo> </mo> <mo>(</mo> <mn>4</mn> <mi mathvariant="sans-serif">Φ</mi> <mo>,</mo> <mn>2</mn> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math> photothermal reaction.</p>
Full article ">Figure 5
<p>Electronic absorption of reactant (plain blue line), photoproduct (plain orange line), patterns of the quantum yields of <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="sans-serif">Φ</mi> </mrow> <mrow> <mi>X</mi> <mo>→</mo> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> <mrow> <msub> <mrow> <mi>λ</mi> </mrow> <mrow> <mi>i</mi> <mi>r</mi> <mi>r</mi> </mrow> </msub> </mrow> </msubsup> </mrow> </semantics></math> (×3 10<sup>5</sup>, long−dashed purple line), and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="sans-serif">Φ</mi> </mrow> <mrow> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> <mo>→</mo> <mi>X</mi> </mrow> <mrow> <msub> <mrow> <mi>λ</mi> </mrow> <mrow> <mi>i</mi> <mi>r</mi> <mi>r</mi> </mrow> </msub> </mrow> </msubsup> </mrow> </semantics></math> (×3 10<sup>5</sup>, long−dashed red line), the absorption spectrum of <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>P</mi> <mi>M</mi> </mrow> </semantics></math> (×4 10<sup>3</sup>, dotted−dashed black line), and lamp profile (×10<sup>11</sup>, dashed green−line), of a photochromic <math display="inline"><semantics> <mrow> <mi>X</mi> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> <mo> </mo> <mo>(</mo> <mn>2</mn> <mi mathvariant="sans-serif">Φ</mi> <mo>,</mo> <mn>1</mn> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math> reaction (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>k</mi> </mrow> <mrow> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> <mo>→</mo> <mi>X</mi> </mrow> </msub> <mo>=</mo> <mn>0.0028</mn> <mo> </mo> <msup> <mrow> <mi>s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>X</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mn>0</mn> </mrow> </mfenced> <mo>=</mo> <mn>4</mn> <mo>×</mo> <mn>1</mn> <msup> <mrow> <mn>0</mn> </mrow> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> <mo> </mo> <mi>M</mi> </mrow> </semantics></math>).</p>
Full article ">Figure 6
<p>Reduction of the initial reactant velocity with increasing <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>P</mi> <mi>M</mi> </mrow> </semantics></math> concentration represented here by the values of absorbance at 283 nm on the <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>P</mi> <mi>M</mi> </mrow> </semantics></math> spectrum.</p>
Full article ">Figure 7
<p>Examples of linear relationships of the initial velocity vs. incident radiation intensity for various photothermal reactions proposed in the literature.</p>
Full article ">Figure 8
<p>Examples of linear relationships of <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>Y</mi> </mrow> </semantics></math> vs. time of intensity for <math display="inline"><semantics> <mrow> <mi>X</mi> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> <mo> </mo> <mo>(</mo> <mn>4</mn> <mi mathvariant="sans-serif">Φ</mi> <mo>,</mo> <mn>2</mn> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math> reaction. The photonic yield is determined for a variation of either the incident light intensity (<b>a</b>) or the temperature of the medium (<b>b</b>).</p>
Full article ">Scheme 1
<p>Mechanism of <math display="inline"><semantics> <mrow> <mi>X</mi> <msub> <mrow> <mi>Y</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> <mfenced separators="|"> <mrow> <mn>6</mn> <mi mathvariant="sans-serif">Φ</mi> <mo>,</mo> <mn>5</mn> <mi>k</mi> </mrow> </mfenced> </mrow> </semantics></math> photothermal reaction, showing photochemical (blue) and thermal (red) reaction steps (a similar mechanism was proposed for naphthopyrans photochromic reactions [<a href="#B51-molecules-30-00330" class="html-bibr">51</a>]).</p>
Full article ">
16 pages, 313 KiB  
Article
On Superization of Nonlinear Integrable Dynamical Systems
by Anatolij K. Prykarpatski, Radosław A. Kycia and Volodymyr M. Dilnyi
Symmetry 2025, 17(1), 125; https://doi.org/10.3390/sym17010125 - 15 Jan 2025
Viewed by 212
Abstract
We study an interesting superization problem of integrable nonlinear dynamical systems on functional manifolds. As an example, we considered a quantum many-particle Schrödinger–Davydov model on the axis, whose quasi-classical reduction proved to be a completely integrable Hamiltonian system on a smooth functional manifold. [...] Read more.
We study an interesting superization problem of integrable nonlinear dynamical systems on functional manifolds. As an example, we considered a quantum many-particle Schrödinger–Davydov model on the axis, whose quasi-classical reduction proved to be a completely integrable Hamiltonian system on a smooth functional manifold. We checked that the so-called “naive” approach, based on the superization of the related phase space variables via extending the corresponding Poisson brackets upon the related functional supermanifold, fails to retain the dynamical system super-integrability. Moreover, we demonstrated that there exists a wide class of classical Lax-type integrable nonlinear dynamical systems on axes in relation to which a superization scheme consists in a reasonable superization of the related Lax-type representation by means of passing from the basic algebra of pseudo-differential operators on the axis to the corresponding superalgebra of super-pseudodifferential operators on the superaxis. Full article
(This article belongs to the Special Issue Symmetry in Nonlinear Dynamics and Chaos II)
20 pages, 8292 KiB  
Review
Monte Carlo Simulations in Nanomedicine: Advancing Cancer Imaging and Therapy
by James C. L. Chow
Nanomaterials 2025, 15(2), 117; https://doi.org/10.3390/nano15020117 - 15 Jan 2025
Viewed by 351
Abstract
Monte Carlo (MC) simulations have become important in advancing nanoparticle (NP)-based applications for cancer imaging and therapy. This review explores the critical role of MC simulations in modeling complex biological interactions, optimizing NP designs, and enhancing the precision of therapeutic and diagnostic strategies. [...] Read more.
Monte Carlo (MC) simulations have become important in advancing nanoparticle (NP)-based applications for cancer imaging and therapy. This review explores the critical role of MC simulations in modeling complex biological interactions, optimizing NP designs, and enhancing the precision of therapeutic and diagnostic strategies. Key findings highlight the ability of MC simulations to predict NP bio-distribution, radiation dosimetry, and treatment efficacy, providing a robust framework for addressing the stochastic nature of biological systems. Despite their contributions, MC simulations face challenges such as modeling biological complexity, computational demands, and the scarcity of reliable nanoscale data. However, emerging technologies, including hybrid modeling approaches, high-performance computing, and quantum simulation, are poised to overcome these limitations. Furthermore, novel advancements such as FLASH radiotherapy, multifunctional NPs, and patient-specific data integration are expanding the capabilities and clinical relevance of MC simulations. This topical review underscores the transformative potential of MC simulations in bridging fundamental research and clinical translation. By facilitating personalized nanomedicine and streamlining regulatory and clinical trial processes, MC simulations offer a pathway toward more effective, tailored, and accessible cancer treatments. The continued evolution of simulation techniques, driven by interdisciplinary collaboration and technological innovation, ensures that MC simulations will remain at the forefront of nanomedicine’s progress. Full article
Show Figures

Figure 1

Figure 1
<p>Contribution of NPs in nanomedicine for cancer imaging and therapy.</p>
Full article ">Figure 2
<p>MC models of (<b>a</b>) cone-beam CT detector and Shepp–Logan phantom and (<b>b</b>) PET detection system and ACR-type phantom. Reproduced from reference [<a href="#B9-nanomaterials-15-00117" class="html-bibr">9</a>] under the Creative Commons Attribution 4.0 International License (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a> (accessed on 9 December 2024)).</p>
Full article ">Figure 3
<p>MC simulation models of irradiated single AuNPs of varying radii interacting with a DNA molecule. (<b>a</b>) shows a NP with a 5 nm radius, (<b>b</b>) shows a NP with a 3.97 nm radius, and (<b>c</b>) shows a NP with a 3.15 nm radius. The red tracks in each subfigure represent the paths of secondary electrons generated during the simulation. Reproduced from reference [<a href="#B39-nanomaterials-15-00117" class="html-bibr">39</a>] under the Creative Commons Attribution 4.0 International License (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a> (accessed on 10 December 2024)).</p>
Full article ">Figure 4
<p>Relationship between total strand breaks and electron energy (keV) is influenced by the number of AuNPs present. Reproduced from reference [<a href="#B39-nanomaterials-15-00117" class="html-bibr">39</a>] under the Creative Commons Attribution 4.0 International License (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a> (accessed on 9 December 2024)).</p>
Full article ">Figure 5
<p>Role of MC Simulations in NP Research.</p>
Full article ">Figure 6
<p>Three common radiolabeling methods involving metallic radionuclides and NPs. Reproduced from reference [<a href="#B49-nanomaterials-15-00117" class="html-bibr">49</a>] under the Creative Commons Attribution 4.0 International License (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a> (accessed on 10 December 2024)).</p>
Full article ">Figure 7
<p>(<b>A</b>) Visualized samples of uncertainty maps generated by the MC dropout and MC-ASM, along with the corresponding error maps. (<b>B</b>) Quantitative comparison between (1) Arbitrary-Masked Mamba-based model (MambaMIR) without MC-ASM or MC dropout (control group), (2) MambaMIR with MC-ASM, and (3) MambaMIR with MC dropout across three datasets. Reproduced from reference [<a href="#B58-nanomaterials-15-00117" class="html-bibr">58</a>] under the Creative Commons Attribution 4.0 International License (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a> (accessed on 10 December 2024)).</p>
Full article ">Figure 8
<p>Relationship between the dose enhancement ratio (DER) and skin target thickness with varying concentrations of AuNPs using 105 and 220 kVp photon beams. Reproduced from reference [<a href="#B74-nanomaterials-15-00117" class="html-bibr">74</a>] under the Creative Commons Attribution 4.0 International License (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a> (accessed on 10 December 2024)).</p>
Full article ">Figure 9
<p>Temperature distribution of the medium (φ<sub>drr</sub> = 1, 3 AuNP injections, Pl = 50 mW). Reproduced from reference [<a href="#B81-nanomaterials-15-00117" class="html-bibr">81</a>] under the Creative Commons Attribution 4.0 International License (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a> (accessed on 10 December 2024)).</p>
Full article ">Figure 10
<p>Depiction of an AuNP for theranostic applications. Reproduced from reference [<a href="#B5-nanomaterials-15-00117" class="html-bibr">5</a>] under the Creative Commons Attribution 4.0 International License (<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a> (accessed on 10 December 2024)).</p>
Full article ">
31 pages, 4117 KiB  
Article
A Decentralized Storage and Security Engine (DeSSE) Using Information Fusion Based on Stochastic Processes and Quantum Mechanics
by Gerardo Iovane and Riccardo Amatore
Appl. Sci. 2025, 15(2), 759; https://doi.org/10.3390/app15020759 - 14 Jan 2025
Viewed by 312
Abstract
In the context of data security, this work aims to present a novel solution that, rather than addressing the topic of endpoint security—which has already garnered significant attention within the international scientific community—offers a different perspective on the subject. In other words, the [...] Read more.
In the context of data security, this work aims to present a novel solution that, rather than addressing the topic of endpoint security—which has already garnered significant attention within the international scientific community—offers a different perspective on the subject. In other words, the focus is not on device security but rather on the protection and security of the information contained within those devices. As we will see, the result is a next-generation decentralized infrastructure that simultaneously integrates two cognitive areas: data storage and its protection and security. In this context, an innovative Multiscale Relativistic Quantum (MuReQua) chain is considered to realize a novel decentralized and security solution for storing data. This engine is based on the principles of Quantum Mechanics, stochastic processes, and a new approach of decentralization for data storage focused on information security. The solution is broken down into four main components, considered four levels of security against attackers: (i) defocusing, (ii) fogging, (iii) puzzling, and (iv) crypto agility. The defocusing is realized thanks to a fragmentation of the contents and their distributions on different allocations, while the fogging is a component consisting of a solution of hybrid cyphering. Then, the puzzling is a unit of Information Fusion and Inverse Information Fusion, while the crypto agility component is a frontier component based on Quantum Computing, which gives a stochastic dynamic to the information and, in particular, to its data fragments. The data analytics show a very effective and robust solution, with executions time comparable with cloud technologies, but with a level of security that is a post quantum one. In the end, thanks to a specific application example, going beyond purely technical and technological aspects, this work introduces a new cognitive perspective regarding (i) the distinction between data and information, and (ii) the differentiation between the owner and the custodian of data. Full article
(This article belongs to the Special Issue New Advances in Computer Security and Cybersecurity)
Show Figures

Figure 1

Figure 1
<p>Conceptual upload architecture. This figure illustrates the system flow during the upload process, highlighting key components and operations involved.</p>
Full article ">Figure 2
<p>Description of the logical architecture of the upload process.</p>
Full article ">Figure 3
<p>Diagram illustrating the interaction between fragments FrA and FrB, which, through entanglement and mixing processes, transform into new fragments, FrC and FrD.</p>
Full article ">Figure 4
<p>Conceptual download architecture. This figure illustrates the system flow during the download process, highlighting key components and operations involved.</p>
Full article ">Figure 5
<p>Description of the logical architecture of the download process.</p>
Full article ">Figure 6
<p>This metric represents the average number of fragments into which a file is divided before being saved to storage per system.</p>
Full article ">Figure 7
<p>This figure measures how long a system takes to completely process and save a file of a certain size, including the division into fragments, the saving of each fragment, and any other process overhead.</p>
Full article ">Figure 8
<p>This indicator measures the average time required to create and save a single fragment of a file.</p>
Full article ">Figure 9
<p>These charts represent the distribution of average creation and saving times of file fragments in relation to their size.</p>
Full article ">Figure 10
<p>The figure indicates how large the data blocks are into which a file is divided during the saving process.</p>
Full article ">Figure 11
<p>Efficiency ratio of different solutions relative to file size.</p>
Full article ">Figure 12
<p>Here, we display the proportion of the total time attributed to overhead activities.</p>
Full article ">
40 pages, 7115 KiB  
Article
Emotion Recognition from EEG Signals Using Advanced Transformations and Deep Learning
by Jonathan Axel Cruz-Vazquez, Jesús Yaljá Montiel-Pérez, Rodolfo Romero-Herrera and Elsa Rubio-Espino
Mathematics 2025, 13(2), 254; https://doi.org/10.3390/math13020254 - 14 Jan 2025
Viewed by 486
Abstract
Affective computing aims to develop systems capable of effectively interacting with people through emotion recognition. Neuroscience and psychology have established models that classify universal human emotions, providing a foundational framework for developing emotion recognition systems. Brain activity related to emotional states can be [...] Read more.
Affective computing aims to develop systems capable of effectively interacting with people through emotion recognition. Neuroscience and psychology have established models that classify universal human emotions, providing a foundational framework for developing emotion recognition systems. Brain activity related to emotional states can be captured through electroencephalography (EEG), enabling the creation of models that classify emotions even in uncontrolled environments. In this study, we propose an emotion recognition model based on EEG signals using deep learning techniques on a proprietary database. To improve the separability of emotions, we explored various data transformation techniques, including Fourier Neural Networks and quantum rotations. The convolutional neural network model, combined with quantum rotations, achieved a 95% accuracy in emotion classification, particularly in distinguishing sad emotions. The integration of these transformations can further enhance overall emotion recognition performance. Full article
(This article belongs to the Special Issue Deep Neural Networks: Theory, Algorithms and Applications)
Show Figures

Figure 1

Figure 1
<p>Emotion classification processing workflow with EEG signals.</p>
Full article ">Figure 2
<p>Recording protocol and stimulus exposure at different times of the day.</p>
Full article ">Figure 3
<p>Recording room for the experiment.</p>
Full article ">Figure 4
<p>(<b>a</b>) Diagram of electrode positions on the scalp according to the 10–20 system; (<b>b</b>) representation of brain regions corresponding to the electrodes.</p>
Full article ">Figure 5
<p>Graph of raw EEG signals from 14 channels over time.</p>
Full article ">Figure 6
<p>Notch filter application: (<b>a</b>) before applying the notch filter, (<b>b</b>) after applying the notch filter.</p>
Full article ">Figure 7
<p>Spatial distribution of independent components (ICA).</p>
Full article ">Figure 8
<p>Detailed analysis of ICA004 component: (<b>a</b>) topographic map, (<b>b</b>) segment image and ERP/ERF, (<b>c</b>) frequency spectrum, (<b>d</b>) dropped segments.</p>
Full article ">Figure 9
<p>Compares EEG signals before and after cleaning artifacts using ICA.</p>
Full article ">Figure 10
<p>Recording of EEG signals from 14 channels during ERP segmentation.</p>
Full article ">Figure 11
<p>Average ERP signals across the 14 EEG channels.</p>
Full article ">Figure 12
<p>Topographic maps of temporal evolution of an ERP.</p>
Full article ">Figure 13
<p>Representation of the 14 EEG channels and topographic maps of an ERP.</p>
Full article ">Figure 14
<p>Scatter plots of extracted features.</p>
Full article ">Figure 15
<p>Correlation matrix of EEG time and frequency domain features.</p>
Full article ">Figure 16
<p>Distribution of features transformed with the Fourier Neural Network.</p>
Full article ">Figure 17
<p>Quantum rotations: (<b>a</b>) emotions before applying quantum rotations, (<b>b</b>) emotions after applying quantum rotations.</p>
Full article ">Figure 18
<p>Quantum rotated features for different emotional states: (<b>a</b>) quantum-rotated features for happy, (<b>b</b>) quantum-rotated features for sad, (<b>c</b>) quantum-rotated features for neutral.</p>
Full article ">Figure 19
<p>Performance of dense network with Fourier features: (<b>a</b>) confusion matrix, (<b>b</b>) precision curves for training and validation, (<b>c</b>) loss curves for training and validation.</p>
Full article ">Figure 20
<p>Performance of dense network with quantum-rotated features: (<b>a</b>) confusion matrix, (<b>b</b>) precision curves for training and validation, (<b>c</b>) loss curves for training and validation.</p>
Full article ">Figure 21
<p>Performance of convolutional neural network (CNN) with Fourier features: (<b>a</b>) confusion matrix, (<b>b</b>) precision curves for training and validation, (<b>c</b>) loss curves for training and validation.</p>
Full article ">Figure 22
<p>Performance of convolutional neural network (CNN) with quantum-rotated features: (<b>a</b>) confusion matrix, (<b>b</b>) precision curves for training and validation, (<b>c</b>) loss curves for training and validation.</p>
Full article ">
10 pages, 216 KiB  
Article
Issues in the Expansion and Contraction of Operators
by Mukhtarbay Otelbaev, Abdukhali Shynybekov and Karlygash Dosmagulova
Symmetry 2025, 17(1), 117; https://doi.org/10.3390/sym17010117 - 14 Jan 2025
Viewed by 258
Abstract
This paper explores fundamental issues in the correct contraction and expansion of operators, with a primary focus on the concept of symmetry within operator theory. Special attention is given to how symmetry influences the behavior of operators, particularly regarding their approximation and convergence [...] Read more.
This paper explores fundamental issues in the correct contraction and expansion of operators, with a primary focus on the concept of symmetry within operator theory. Special attention is given to how symmetry influences the behavior of operators, particularly regarding their approximation and convergence properties. In the domains of quantum mechanics and condensed matter physics, such operators are essential for modeling phenomena like superconductivity, excitons, and surface states. The symmetric properties of operators have a profound impact on the physical interpretations and predictions these models generate. A rigorous analysis is provided regarding the existence of correct contractions and expansions for a specific class of nonlinear operators, demonstrating how symmetry affects the structural integrity of operators under natural conditions. The study presents a comprehensive description of the set of all correct contractions, expansions, and regular expansions, with an application to a third-order nonlinear differential expression. Additionally, a condition for the unique solvability of a Bitsadze–Samarskii-type problem is derived, showcasing how symmetry plays a crucial role in guiding the solution of complex physical models. Furthermore, the paper emphasizes the importance of preserving symmetry in the construction of operators, ensuring the consistency and accuracy of mathematical models. This has significant implications for both theoretical research and practical applications in various fields, including nuclear physics and quantum theory. Full article
(This article belongs to the Section Mathematics)
20 pages, 12787 KiB  
Article
Exploring the Properties of Quantum Scars in a Toy Model
by Sudip Sinha and Subhasis Sinha
Condens. Matter 2025, 10(1), 5; https://doi.org/10.3390/condmat10010005 - 12 Jan 2025
Viewed by 229
Abstract
We introduce the concept of ergodicity and explore its deviation caused by quantum scars in an isolated quantum system, employing a pedagogical approach based on a toy model. Quantum scars, originally identified as traces of classically unstable orbits in certain wavefunctions of chaotic [...] Read more.
We introduce the concept of ergodicity and explore its deviation caused by quantum scars in an isolated quantum system, employing a pedagogical approach based on a toy model. Quantum scars, originally identified as traces of classically unstable orbits in certain wavefunctions of chaotic systems, have recently regained interest for their role in non-ergodic dynamics, as they retain memory of their initial states. We elucidate these features of quantum scars within the same framework of this toy model. The integrable part of the model consists of two large spins, with a classical counterpart, which we combine with a random matrix to induce ergodic behavior. Scarred states can be selectively generated from the integrable spin Hamiltonian by protecting them from the ergodic states using a projector method. Deformed projectors mimic the ‘quantum leakage’ of scarred states, enabling tunable mixing with ergodic states and thereby controlling the degree of scarring. In this simple model, we investigate various properties of quantum scarring and shed light on different aspects of many-body quantum scars observed in more complex quantum systems. Notably, the underlying classicality can be revealed through the entanglement spectrum and the dynamics of ‘out-of-time-ordered correlators’. Full article
(This article belongs to the Special Issue Non-equilibrium Dynamics in Ultra-Cold Quantum Gases)
Show Figures

Figure 1

Figure 1
<p>Ergodic properties for <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> are as follows: (<b>a</b>) Variation in Entanglement entropy (EE), <math display="inline"><semantics> <msub> <mi>S</mi> <mrow> <mi>e</mi> <mi>n</mi> </mrow> </msub> </semantics></math>, of the energy eigenstates with energy density, <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>≡</mo> <msub> <mi mathvariant="script">E</mi> <mi>n</mi> </msub> <mo>/</mo> <mi>S</mi> </mrow> </semantics></math>. The pink dashed line corresponds to the page value of EE [see Equation (<a href="#FD14-condensedmatter-10-00005" class="html-disp-formula">14</a>)]. The insets (<b>a1</b>,<b>a2</b>) display the ergodic and projected states zoomed in a small window around <span class="html-italic">E</span> ∼ 0, respectively. (<b>b</b>) Distribution of the consecutive level spacing ratio <math display="inline"><semantics> <mrow> <mi>P</mi> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </semantics></math> for 100 different realizations (ensembles) of random matrix <math display="inline"><semantics> <msub> <mover accent="true"> <mi mathvariant="script">H</mi> <mo stretchy="false">^</mo> </mover> <mi>GOE</mi> </msub> </semantics></math>, which agrees well with the GOE class as <math display="inline"><semantics> <mrow> <mo stretchy="false">〈</mo> <mi>r</mi> <mo stretchy="false">〉</mo> </mrow> </semantics></math> ∼ 0.523. We set <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math> in this and the rest of the figures unless otherwise specified.</p>
Full article ">Figure 2
<p>The properties of the scarred states for a single realization of random matrix: (<b>a1</b>,<b>b1</b>) EE of the energy eigenstates for different values of <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math>. The black squares denote EE of the eigenstates for <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. Entanglement spectrum (ES) and Husimi distribution corresponding to the first spin sector of the energy eigenstates (blue circles) marked in (<b>a1</b>,<b>b1</b>) for (<b>a2</b>,<b>a3</b>) <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> ∼ 0.01 and (<b>b2</b>,<b>b3</b>) <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> ∼ 0.05, respectively. The double headed arrows mark the gap, <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>λ</mi> </mrow> </semantics></math>, separating the largest eigenvalue and the extended tail in the ES. The red diamonds correspond to the ES of a random eigenstate of a GOE matrix of the same dimension, reflecting the ergodic behavior.</p>
Full article ">Figure 3
<p>Degree of scarring: (<b>a</b>) Variation in the average deviation of the EE <math display="inline"><semantics> <msub> <mover> <mrow> <mo>Δ</mo> <mi>S</mi> </mrow> <mo>¯</mo> </mover> <mrow> <mi>e</mi> <mi>n</mi> </mrow> </msub> </semantics></math> scaled by the maximum limit (page value), and the average gap <math display="inline"><semantics> <mover> <mrow> <mo>Δ</mo> <mi>λ</mi> </mrow> <mo>¯</mo> </mover> </semantics></math> in the ES (inset) at energy density <span class="html-italic">E</span> ∼ 0 with increasing deformation strength <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math>, for <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>. (<b>b</b>) Variation in <math display="inline"><semantics> <mrow> <mover> <mrow> <mo>Δ</mo> <msub> <mi>S</mi> <mrow> <mi>e</mi> <mi>n</mi> </mrow> </msub> </mrow> <mo>¯</mo> </mover> <mo>/</mo> <msub> <mi>S</mi> <mrow> <mi>p</mi> <mi>a</mi> <mi>g</mi> <mi>e</mi> </mrow> </msub> </mrow> </semantics></math> with increasing large spin magnitude <span class="html-italic">S</span> for different values of <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math>. The averaging (indicated by <math display="inline"><semantics> <mover accent="true"> <mo>·</mo> <mo>¯</mo> </mover> </semantics></math> ) is performed in two steps: first, over the states with minimum EE in a small energy density window <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>E</mi> <mo>=</mo> <mn>0.09</mn> <mo>/</mo> <mi>S</mi> <mo>&lt;</mo> <mn>1</mn> <mo>/</mo> <mi>S</mi> </mrow> </semantics></math> around <span class="html-italic">E</span> ∼ 0, and then over an ensemble of (<b>a</b>) 100 and (<b>b</b>) 1000 random matrices <math display="inline"><semantics> <msub> <mover accent="true"> <mi mathvariant="script">H</mi> <mo stretchy="false">^</mo> </mover> <mi>GOE</mi> </msub> </semantics></math>.</p>
Full article ">Figure 4
<p>Quantum dynamics starting from the initial state <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mi>θ</mi> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>⟩</mo> </mrow> <mo>⊗</mo> <mrow> <mo>|</mo> <msub> <mi>m</mi> <mrow> <mn>2</mn> <mi>z</mi> </mrow> </msub> <mo>=</mo> <mn>0</mn> <mo>⟩</mo> </mrow> </mrow> </semantics></math> at <span class="html-italic">E</span> ∼ 0: Time evolution of survival probability <math display="inline"><semantics> <mrow> <mi>F</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mrow> <mo>|</mo> <mrow> <mo stretchy="false">〈</mo> <mi>ψ</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>|</mo> <mi>ψ</mi> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo stretchy="false">〉</mo> </mrow> <mo>|</mo> </mrow> <mn>2</mn> </msup> </mrow> </semantics></math> for (<b>a1</b>) <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and (<b>b1</b>) <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math> (red lines), compared with the dynamics for <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0.0</mn> </mrow> </semantics></math> (blue line). Snapshots of Husimi distribution corresponding to the first spin sector at different times for (<b>a2</b>,<b>a3</b>) <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and (<b>b2</b>,<b>b3</b>) <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>. The green line in (<b>a1</b>) exhibits a rapid relaxation of <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> starting from the initial state <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mi>θ</mi> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>⟩</mo> </mrow> <mo>⊗</mo> <mrow> <mo>|</mo> <msub> <mi>m</mi> <mrow> <mn>2</mn> <mi>z</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo>⟩</mo> </mrow> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, indicating a complete loss of memory of the initial state (ergodic dynamics), as it is not a linear combination of the perfectly projected states.</p>
Full article ">Figure 5
<p>Quantum dynamics starting from the initial state <math display="inline"><semantics> <mrow> <mrow> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mi>θ</mi> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> </mrow> <mo>⊗</mo> <mrow> <msub> <mi>m</mi> <mrow> <mn>2</mn> <mi>z</mi> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </mrow> </semantics></math> at <span class="html-italic">E</span> ∼ 0: (<b>a</b>–<b>d</b>) Time evolution of <math display="inline"><semantics> <mrow> <mover> <mrow> <mo stretchy="false">〈</mo> <msub> <mover accent="true"> <mi>S</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mn>1</mn> <mi>x</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>/</mo> <mi>S</mi> <mo stretchy="false">〉</mo> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> for different values of <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math>. Fluctuation dynamics of the spin operators (<b>e</b>) <math display="inline"><semantics> <mrow> <msubsup> <mover accent="true"> <mi>S</mi> <mo>¯</mo> </mover> <mrow> <mn>1</mn> <mi>z</mi> </mrow> <mrow> <mi>f</mi> <mi>l</mi> <mi>u</mi> <mi>c</mi> </mrow> </msubsup> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> and (<b>f</b>) <math display="inline"><semantics> <mrow> <msubsup> <mover accent="true"> <mi>S</mi> <mo>¯</mo> </mover> <mrow> <mn>1</mn> <mi>x</mi> </mrow> <mrow> <mi>f</mi> <mi>l</mi> <mi>u</mi> <mi>c</mi> </mrow> </msubsup> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>. The dotted line denotes the microcanonical value given by Equation (<a href="#FD28-condensedmatter-10-00005" class="html-disp-formula">28</a>). Note that, <math display="inline"><semantics> <mover accent="true"> <mrow> <mo stretchy="false">〈</mo> <mo>·</mo> <mo stretchy="false">〉</mo> </mrow> <mo>¯</mo> </mover> </semantics></math> and <math display="inline"><semantics> <mover accent="true"> <mo>·</mo> <mo>¯</mo> </mover> </semantics></math> represent averaging over an ensemble of 100 random matrices.</p>
Full article ">Figure 6
<p>OTOC dynamics: Time evolution of <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>C</mi> <mo>¯</mo> </mover> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for eigenstate with minimum entanglement in a small energy density window <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>E</mi> <mo>=</mo> <mn>0.045</mn> <mo>&lt;</mo> <mn>1</mn> <mo>/</mo> <mi>S</mi> </mrow> </semantics></math> around <span class="html-italic">E</span> ∼ 0. The inset shows the OTOC dynamics zoomed at long time and the dotted line denotes the microcanonical value given by Equation (<a href="#FD34-condensedmatter-10-00005" class="html-disp-formula">34</a>). Note that, the OTOC dynamics is averaged over 100 ensembles of random matrices.</p>
Full article ">
60 pages, 821 KiB  
Review
Introduction to Thermal Field Theory: From First Principles to Applications
by Alberto Salvio
Universe 2025, 11(1), 16; https://doi.org/10.3390/universe11010016 - 11 Jan 2025
Viewed by 253
Abstract
This review article provides the basics and discusses some important applications of thermal field theory, namely, the combination of statistical mechanics and relativistic quantum field theory. In the first part, the fundamentals are covered: the density matrix, the corresponding averages, and the treatment [...] Read more.
This review article provides the basics and discusses some important applications of thermal field theory, namely, the combination of statistical mechanics and relativistic quantum field theory. In the first part, the fundamentals are covered: the density matrix, the corresponding averages, and the treatment of fields of various spin in a medium. The second part is dedicated to the computation of thermal Green’s function for scalars, vectors, and fermions with path-integral methods. These functions play a crucial role in thermal field theory as explained here. A more applicative part of the review is dedicated to the production of particles in a medium and to phase transitions in field theory, including the process of vacuum decay in a general theory featuring a first-order phase transition. To understand this review, the reader should have good knowledge of non-statistical quantum field theory. Full article
(This article belongs to the Section Field Theory)
Show Figures

Figure 1

Figure 1
<p>A contour <italic>C</italic> which allows us to compute the thermal Green’s function directly at real times. Here, <inline-formula><mml:math id="mm1567"><mml:semantics><mml:mrow><mml:msub><mml:mi>t</mml:mi><mml:mn>0</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:msub><mml:mi>t</mml:mi><mml:mi>i</mml:mi></mml:msub><mml:mo>→</mml:mo><mml:mo>−</mml:mo><mml:mo>∞</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 2
<p>The diagram corresponding to the non-time-ordered self-energy of the weakly coupled particle. The external vertex on the left is circled, the external vertex on the right is uncircled, and one sums over all possible ways of circling the internal vertices.</p>
Full article ">Figure 3
<p><bold>Left</bold>: The temperature-dependent effective potential built around the Coleman–Weinberg potential (azure solid line). This effective potential features a first-order phase transition. <bold>Right</bold>: The temperature-dependent effective potential of Equation (<xref ref-type="disp-formula" rid="FD362-universe-11-00016">362</xref>), which features a second-order phase transition. Here, <inline-formula><mml:math id="mm1568"><mml:semantics><mml:mrow><mml:mi>B</mml:mi><mml:mo>=</mml:mo><mml:mi>λ</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 4
<p>Bounces at various temperatures. The vertical and horizontal axis represent the Euclidean time direction and the spatial radius, respectively. At <inline-formula><mml:math id="mm1569"><mml:semantics><mml:mrow><mml:mi>T</mml:mi><mml:mo>=</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (left-most panel), the bounce solution represents a bubble. At finite temperature, the bounce solution becomes a series of bubbles placed at distance <inline-formula><mml:math id="mm1570"><mml:semantics><mml:mi>β</mml:mi></mml:semantics></mml:math></inline-formula> in the time direction. At large temperature (right-most panel), the bounce no longer depends on time. Figure reproduced from [<xref ref-type="bibr" rid="B63-universe-11-00016">63</xref>].</p>
Full article ">
54 pages, 671 KiB  
Article
Quantum-Ordering Ambiguities in Weak Chern—Simons 4D Gravity and Metastability of the Condensate-Induced Inflation
by Panagiotis Dorlis, Nick E. Mavromatos and Sotirios-Neilos Vlachos
Universe 2025, 11(1), 15; https://doi.org/10.3390/universe11010015 - 11 Jan 2025
Viewed by 260
Abstract
In this work, we elaborate further on a (3+1)-dimensional cosmological Running-Vacuum-type-Model (RVM) of inflation based on string-inspired Chern-Simons(CS) gravity, involving axions coupled to gravitational-CS(gCS) anomalous terms. Inflation in such models is caused by primordial-gravitational-waves(GW)-induced condensation of the gCS terms, which leads to a [...] Read more.
In this work, we elaborate further on a (3+1)-dimensional cosmological Running-Vacuum-type-Model (RVM) of inflation based on string-inspired Chern-Simons(CS) gravity, involving axions coupled to gravitational-CS(gCS) anomalous terms. Inflation in such models is caused by primordial-gravitational-waves(GW)-induced condensation of the gCS terms, which leads to a linear-axion potential. We demonstrate that this inflationary phase may be metastable, due to the existence of imaginary parts of the gCS condensate. These are quantum effects, proportional to commutators of GW perturbations, hence vanishing in the classical theory. Their existence is quantum-ordering-scheme dependent. We argue in favor of a physical importance of such imaginary parts, which we compute to second order in the GW (tensor) perturbations in the framework of a gauge-fixed effective Lagrangian, within a (mean field) weak-quantum-gravity-path-integral approach. We thus provide estimates of the inflation lifetime. On matching our results with the inflationary phenomenology, we fix the quantum-ordering ambiguities, and obtain an order-of-magnitude constraint on the String-Mass-Scale-to-Planck-Mass ratio, consistent with previous estimates by the authors in the framework of a dynamical-system approach to linear-axion RVM inflation. Finally, we examine the role of periodic modulations in the axion potential induced by non-perturbative effects on the slow-roll inflationary parameters, and find compatibility with the cosmological data. Full article
28 pages, 2948 KiB  
Review
Integration of Functional Materials in Photonic and Optoelectronic Technologies for Advanced Medical Diagnostics
by Naveen Thanjavur, Laxmi Bugude and Young-Joon Kim
Biosensors 2025, 15(1), 38; https://doi.org/10.3390/bios15010038 - 10 Jan 2025
Viewed by 730
Abstract
Integrating functional materials with photonic and optoelectronic technologies has revolutionized medical diagnostics, enhancing imaging and sensing capabilities. This review provides a comprehensive overview of recent innovations in functional materials, such as quantum dots, perovskites, plasmonic nanomaterials, and organic semiconductors, which have been instrumental [...] Read more.
Integrating functional materials with photonic and optoelectronic technologies has revolutionized medical diagnostics, enhancing imaging and sensing capabilities. This review provides a comprehensive overview of recent innovations in functional materials, such as quantum dots, perovskites, plasmonic nanomaterials, and organic semiconductors, which have been instrumental in the development of diagnostic devices characterized by high sensitivity, specificity, and resolution. Their unique optical properties enable real-time monitoring of biological processes, advancing early disease detection and personalized treatment. However, challenges such as material stability, reproducibility, scalability, and environmental sustainability remain critical barriers to their clinical translation. Breakthroughs such as green synthesis, continuous flow production, and advanced surface engineering are addressing these limitations, paving the way for next-generation diagnostic tools. This article highlights the transformative potential of interdisciplinary research in overcoming these challenges and emphasizes the importance of sustainable and scalable strategies for harnessing functional materials in medical diagnostics. The ultimate goal is to inspire further innovation in the field, enabling the creation of practical, cost-effective, and environmentally friendly diagnostic solutions. Full article
Show Figures

Figure 1

Figure 1
<p>Illustration of key properties and biomedical imaging applications. (<b>a</b>) Stretchable organic optoelectronic devices. (<b>b</b>) A smart sensing system placed on the skin can detect artery and vein impulses. (<b>c</b>) Smart sensing devices are being used to read information from the human body. (<b>d</b>) Smart sensing devices can be seamlessly incorporated into the human body, facilitating real-time signal capture and data transmission through wireless networks. (<b>e</b>) The combination of wireless data transmission technologies and smart sensing systems is critical for developing human-friendly electronics and diagnostic medical applications that rely on timely feedback from healthcare facilities. These advancements make optoelectronic devices indispensable in applications like high-resolution imaging and real-time therapeutic monitoring. Reproduced from ref. [<a href="#B8-biosensors-15-00038" class="html-bibr">8</a>] with permission from Elsevier Copyright 2021.</p>
Full article ">Figure 2
<p>Depiction of non-functional QDs in targeted cancer therapies. Demonstrates how QDs selectively bind to tumor sites, enhancing imaging contrast and enabling precise localized therapy. These properties contribute to improved tumor identification and treatment efficacy, underscoring the potential of QDs in advanced diagnostic and therapeutic applications. Reproduced from ref. [<a href="#B22-biosensors-15-00038" class="html-bibr">22</a>].</p>
Full article ">Figure 3
<p>Application of plasmonic biosensors for biomarker detection in human samples. Illustrates their high sensitivity and rapid diagnostic capabilities, enabling early disease detection and monitoring of molecular changes in real time. Such advancements are critical for precision medicine and improving clinical outcomes. Reproduced from ref. [<a href="#B76-biosensors-15-00038" class="html-bibr">76</a>] with permission from ACS Copyright 2021.</p>
Full article ">Figure 4
<p>Overview of current challenges and emerging solutions in nanoplasmonic biosensors. It details innovations that address material stability, biocompatibility, and enhanced sensitivity, which are critical for effective integration into clinical diagnostics and for improving real-time disease detection. Reproduced from ref. [<a href="#B82-biosensors-15-00038" class="html-bibr">82</a>].</p>
Full article ">Figure 5
<p>Utilization of organic semiconductors in biomedical imaging and therapeutics. A schematic representation of their biocompatibility and functional flexibility for non-invasive diagnostics and targeted treatment options. Reproduced from ref. [<a href="#B89-biosensors-15-00038" class="html-bibr">89</a>].</p>
Full article ">Figure 6
<p>Dual-modality imaging of pancreatic tumors using targeted nanoplates. (<b>a</b>) Ultrasound Imaging of Pancreatic Tumor: Displays an ultrasound image of a pancreatic tumor in a mouse model, demonstrating non-invasive tumor localization and monitoring potential. (<b>b</b>) Photoacoustic Imaging with Targeted Silver Nanoplates: Shows a photoacoustic image of the tumor using silver nanoplates conjugated shows (yellow). This modality differentiates oxygenated (red) and deoxygenated (blue) blood within the tumor, providing detailed vascular mapping and enhanced tumor visualization for precise diagnostic assessments. Reproduced from ref. [<a href="#B132-biosensors-15-00038" class="html-bibr">132</a>] with permission from ACS Copyright 2012.</p>
Full article ">
13 pages, 4728 KiB  
Article
Graphene/TiO2 Heterostructure Integrated with a Micro-Lightplate for Low-Power NO2 Gas Detection
by Paniz Vafaei, Margus Kodu, Harry Alles, Valter Kiisk, Olga Casals, Joan Daniel Prades and Raivo Jaaniso
Sensors 2025, 25(2), 382; https://doi.org/10.3390/s25020382 - 10 Jan 2025
Viewed by 401
Abstract
Low-power gas sensors that can be used in IoT (Internet of Things) systems, consumer devices, and point-of-care devices will enable new applications in environmental monitoring and health protection. We fabricated a monolithic chemiresistive gas sensor by integrating a micro-lightplate with a 2D sensing [...] Read more.
Low-power gas sensors that can be used in IoT (Internet of Things) systems, consumer devices, and point-of-care devices will enable new applications in environmental monitoring and health protection. We fabricated a monolithic chemiresistive gas sensor by integrating a micro-lightplate with a 2D sensing material composed of single-layer graphene and monolayer-thick TiO2. Applying ultraviolet (380 nm) light with quantum energy above the TiO2 bandgap effectively enhanced the sensor responses. Low (<1 μW optical) power operation of the device was demonstrated by measuring NO2 gas at low concentrations, which is typical in air quality monitoring, with an estimated limit of detection < 0.1 ppb. The gas response amplitudes remained nearly constant over the studied light intensity range (1–150 mW/cm2) owing to the balance between the photoinduced adsorption and desorption processes of the gas molecules. The rates of both processes followed an approximately square-root dependence on light intensity, plausibly because the electron–hole recombination of photoinduced charge carriers is the primary rate-limiting factor. These results pave the way for integrating 2D materials with micro-LED arrays as a feasible path to advanced electronic noses. Full article
(This article belongs to the Special Issue Recent Advances in Sensors for Chemical Detection Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic cross-section of the device, (<b>b</b>) photograph of the μLP with a magnified area of interdigitated electrodes, and (<b>c</b>) sequence of sensor layer fabrication on the μLP.</p>
Full article ">Figure 2
<p>Schematic illustration of the gas sensing setup.</p>
Full article ">Figure 3
<p>(<b>a</b>) Image of the μLP with a working LED, its above-threshold (<b>b</b>) volt–ampere characteristic, (<b>c</b>) electroluminescence (EL) spectrum, and (<b>d</b>) dependence of the μLP optical power and surface intensity on the applied electrical power.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>b</b>) SEM images of CVD graphene on μLP. (<b>c</b>) Raman spectra of graphene before and after the PLD of TiO<sub>2</sub>. (<b>d</b>) SEM image of the sensor material after coating the graphene with a TiO<sub>2</sub> nanolayer.</p>
Full article ">Figure 5
<p>(<b>a</b>) Dynamic responses of pristine graphene to NO<sub>2</sub> gas at concentrations of 20, 50, and 150 ppb at different irradiation intensities on μLP. (<b>b</b>) The same for the Gr/TiO<sub>2</sub> μLP sensor, recorded in the dark and under incremental UV illumination with the μLED optical power of 0.46, 1.9, and 5.5 μW (corresponding to 0.8, 3.3, and 9.7 mW/cm<sup>2</sup>). Synthetic air was used as the background gas. (<b>c</b>) Sensor conductance during the exposures to 150 ppb of NO<sub>2</sub> gas at different levels of μLP optical power. The power levels in μW units are shown in the gray area at the bottom. Synthetic air with a relative humidity (RH) of 20% was used as the background gas.</p>
Full article ">Figure 6
<p>The dependence of average response and recovery rates on light intensity. Approximations with power functions and power exponents of the intensity (I) dependence are shown in red. The inset shows the response curves approximated with Equation (5).</p>
Full article ">Figure 7
<p>Relative responses to (<b>a</b>) 150 ppb of NO<sub>2</sub> at different levels of relative humidity and (<b>b</b>) different toxic gases at concentrations as indicated.</p>
Full article ">
17 pages, 4522 KiB  
Article
The Temperature-Dependent Tight Binding Theory Modelling of Strain and Composition Effects on the Electronic Structure of CdSe- and ZnSe-Based Core/Shell Quantum Dots
by Derya Malkoç and Hilmi Ünlü
Materials 2025, 18(2), 283; https://doi.org/10.3390/ma18020283 - 10 Jan 2025
Viewed by 326
Abstract
We propose a temperature-dependent optimization procedure for the second-nearest neighbor (2NN) sp3s* tight-binding (TB) theory parameters to calculate the effects of strain, structure dimensions, and alloy composition on the band structure of heterostructure spherical core/shell quantum dots (QDs). We integrate [...] Read more.
We propose a temperature-dependent optimization procedure for the second-nearest neighbor (2NN) sp3s* tight-binding (TB) theory parameters to calculate the effects of strain, structure dimensions, and alloy composition on the band structure of heterostructure spherical core/shell quantum dots (QDs). We integrate the thermoelastic theory of solids with the 2NN sp3s* TB theory to calculate the strain, core and shell dimensions, and composition effects on the band structure of binary/ternary CdSe/Cd(Zn)S and ZnSe/Zn(Cd)S QDs at any temperature. We show that the 2NN sp3s* TB theory with optimized parameters greatly improves the prediction of the energy dispersion curve at and in the vicinity of L and X symmetry points. We further used the optimized 2NN sp3s* TB parameters to calculate the strain, core and shell dimensions, and composition effects on the nanocrystal bandgaps of binary/ternary CdSe/Cd(Zn)S and ZnSe/Zn(Cd)S core/shell QDs. We conclude that the 2NN sp3s* TB theory provides remarkable agreement with the measured nanocrystal bandgaps of CdSe/Cd(Zn)S and ZnSe/Zn(Cd)S QDs and accurately reproduces the energy dispersion curves of the electronic band structure at any temperature. We believe that the proposed optimization procedure makes the 2NN sp3s* TB theory reliable and accurate in the modeling of core/shell QDs for nanoscale devices. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Band diagrams of type I (<b>left</b>) and type II (<b>right</b>) heterostructures.</p>
Full article ">Figure 2
<p>A schematic representation of spherical core/shell QD. The outer part (<span class="html-italic">a</span> &lt; <span class="html-italic">r</span> &lt; <span class="html-italic">b</span>) is defined as shell and the inner part (0 &lt; <span class="html-italic">r</span> &lt; <span class="html-italic">a</span>) is defined as core.</p>
Full article ">Figure 3
<p>The strain effect on (<b>left</b>) the core and (<b>right</b>) the shell side of CdSe/Cd(Zn)S and ZnSe/Zn(Cd)S QDs due to core diameter for each quantum dot.</p>
Full article ">Figure 4
<p>The strain effects in CdSe- (<b>left</b>) and ZnSe-(<b>right</b>) based QDs with di = 3.0 nm at 300 K.</p>
Full article ">Figure 5
<p>The strain and lattice constant variation with composition in the cores of (<b>left</b>) ZnSe/CdZnS and CdSe/CdZnS; and (<b>right</b>) ZnSe/CdZnSe and CdSe/CdZnSe QDs for core diameter di = 3.0 nm at 300 K.</p>
Full article ">Figure 6
<p>A comparison of the band structure of bulk CdSe and ZnSe compounds at T = 0 K, calculated by using the 2NN <span class="html-italic">sp</span><sup>3</sup><span class="html-italic">s</span>* TBM and 2NN <span class="html-italic">sp</span><sup>3</sup> tight-binding theories.</p>
Full article ">Figure 7
<p>A comparison of the band structure of the bulk CdSe and ZnSe compounds at T = 0 K, calculated by using 2NN <span class="html-italic">sp</span><sup>3</sup><span class="html-italic">s</span>* TB theory with optimized parameters and <span class="html-italic">k.p</span> effective mass approximation.</p>
Full article ">Figure 8
<p>The band structure of bulk CdSe at T = 0 K, 300 K, 600 K, calculated by using the optimized tight-binding parameters in the 2NN <span class="html-italic">sp</span><sup>3</sup><span class="html-italic">s</span>* TB theory (<b>a</b>). A magnified view of the lowest conduction-band structure in (<b>b</b>) indicates a larger shift in the bandgap at high symmetry points.</p>
Full article ">Figure 9
<p>The band structure of bulk ZnSe at T = 0 K, 300 K, 600 K, calculated by using the optimized tight-binding parameters in the frame of 2NN <span class="html-italic">sp</span><sup>3</sup><span class="html-italic">s</span>* TB theory (<b>a</b>). A magnified view of the lowest conduction-band structure in (<b>b</b>) indicates a larger shift in the bandgap at high symmetry points.</p>
Full article ">Figure 10
<p>The core (<b>a</b>) and shell (<b>b</b>) diameter variations in the nanocrystal bandgap energies of four QDs, determined from Equations (3) and (4) at 300 K.</p>
Full article ">Figure 11
<p>The composition effect on the bandgap energies at the Γ, L, and X symmetry points for (<b>a</b>) CdSe/CdZnS and (<b>b</b>) ZnSe/CdZnSe, QDs with di = 3.0 nm at T = 300 K.</p>
Full article ">
10 pages, 2038 KiB  
Article
Room-Temperature Fiber-Coupled Single-Photon Source from CdTeSeS Core Quantum Dots
by Surasak Chiangga
Photonics 2025, 12(1), 52; https://doi.org/10.3390/photonics12010052 - 9 Jan 2025
Viewed by 393
Abstract
Single-photon sources with photon antibunching characteristics are essential for quantum information technologies. This paper investigates the potential of quaternary-alloy CdTeSeS colloidal core quantum dots (cQDs) as compact, room-temperature, and fiber-integrated single-photon sources. Single-photon emission from CdTeSeS cQDs was verified by measuring the second-order [...] Read more.
Single-photon sources with photon antibunching characteristics are essential for quantum information technologies. This paper investigates the potential of quaternary-alloy CdTeSeS colloidal core quantum dots (cQDs) as compact, room-temperature, and fiber-integrated single-photon sources. Single-photon emission from CdTeSeS cQDs was verified by measuring the second-order correlation function, g2τ, using a Hanbury-Brown and Twiss setup. A novel method to determine zero-time delay through afterpulsing analysis is presented. The results demonstrate strong photon antibunching with g20=0.13, confirming that the photoemission from the CdTeSeS cQDs function as a single-photon source. This work highlights the potential of CdTeSeS cQDs as reliable and efficient single-photon sources for practical use in fiber-based quantum information technologies. Full article
(This article belongs to the Special Issue Recent Progress in Single-Photon Generation and Detection)
Show Figures

Figure 1

Figure 1
<p>Schematic of the experimental setup for photoluminescence measurement.</p>
Full article ">Figure 2
<p>Schematic diagram of the setup. A stream of photons is incident on a beam splitter, directing them to single-photon counting detectors, D1 (start) and D2 (stop). The output pulses from D1 and D2 are fed into the start and stop inputs of a time-to-digital converter (TDC), which records the time interval between photon detection events.</p>
Full article ">Figure 3
<p>Schematic diagram of the setup. LD: laser diode, QDs: CdTeSeS colloidal core quantum dots, FC: fiber collimator, F: filter, APD: avalanche photodetector, TDC: time-to-digital converter.</p>
Full article ">Figure 4
<p>Color map of (<b>a</b>) the 405 nm laser diode spectra and (<b>b</b>) the photoluminescence spectra of CdTeSeS core quantum dots in a methanol solution.</p>
Full article ">Figure 5
<p>(<b>a</b>) Histogram of time delays between photon pairs with a time bin width of 12 ns. (<b>b</b>) Zero delay time as a function of bin width in nanoseconds.</p>
Full article ">Figure 6
<p>Measured second-order correlation function, <math display="inline"><semantics> <mrow> <msup> <mi>g</mi> <mrow> <mfenced> <mn>2</mn> </mfenced> </mrow> </msup> <mfenced> <mi>τ</mi> </mfenced> </mrow> </semantics></math>, showing single-photon antibunching with <math display="inline"><semantics> <mrow> <msup> <mi>g</mi> <mrow> <mfenced> <mn>2</mn> </mfenced> </mrow> </msup> <mfenced> <mn>0</mn> </mfenced> <mo>=</mo> <mn>0.13</mn> </mrow> </semantics></math>. The solid line represents a fit based on Equation (3).</p>
Full article ">
21 pages, 4169 KiB  
Article
Enhancing Deepfake Detection Through Quantum Transfer Learning and Class-Attention Vision Transformer Architecture
by Bekir Eray Katı, Ecir Uğur Küçüksille and Güncel Sarıman
Appl. Sci. 2025, 15(2), 525; https://doi.org/10.3390/app15020525 - 8 Jan 2025
Viewed by 382
Abstract
The widespread use of the internet, coupled with the increasing production of digital content, has caused significant challenges in information security and manipulation. Deepfake detection has become a critical research topic in both academic and practical domains, as it involves identifying forged elements [...] Read more.
The widespread use of the internet, coupled with the increasing production of digital content, has caused significant challenges in information security and manipulation. Deepfake detection has become a critical research topic in both academic and practical domains, as it involves identifying forged elements in artificially generated videos using various deep learning and artificial intelligence techniques. In this dissertation, an innovative model was developed for detecting deepfake videos by combining the Quantum Transfer Learning (QTL) and Class-Attention Vision Transformer (CaiT) architectures. The Deepfake Detection Challenge (DFDC) dataset was used for training, and a system capable of detecting spatiotemporal inconsistencies was constructed by integrating QTL and CaiT technologies. In addition to existing preprocessing methods in the literature, a novel preprocessing function tailored to the requirements of deep learning models was developed for the dataset. The advantages of quantum computing offered by QTL were merged with the global feature extraction capabilities of the CaiT. The results demonstrated that the proposed method achieved a remarkable performance in detecting deepfake videos, with an accuracy of 90% and ROC AUC score of 0.94 achieved. The model’s performance was compared with other methods evaluated on the DFDC dataset, highlighting its efficiency in resource utilization and overall effectiveness. The findings reveal that the proposed QTL-CaiT-based system provides a strong foundation for deepfake detection and contributes significantly to the academic literature. Future research should focus on testing the model on real quantum devices and applying it to larger datasets to further enhance its applicability. Full article
(This article belongs to the Section Quantum Science and Technology)
Show Figures

Figure 1

Figure 1
<p>Model flowchart for deepfake detection based on CaiT and QTL.</p>
Full article ">Figure 2
<p>Diagram of the process for extracting faces from videos and data transformation.</p>
Full article ">Figure 3
<p>Illustrates the stages of superposition, entanglement, and measurement in a quantum circuit. The numbers 0, 1, 2, and 3 represent the qubits in the circuit. The “H” denotes the Hadamard Gate, “RY” represents the Rotation-Y Gate, and the circuit also includes CNOT Gates and measurement symbols.</p>
Full article ">Figure 4
<p>Model construction and integration process.</p>
Full article ">Figure 5
<p>Data processing flowchart for the model training process.</p>
Full article ">Figure 6
<p>Confusion matrix for test data.</p>
Full article ">Figure 7
<p>ROC curve and AUC of the model. The dashed blue line represents the performance of a random classifier, where predictions are no better than chance, with equal true and false positive rates.</p>
Full article ">Figure 8
<p>(<b>a</b>) Confusion matrix showing classification performance on the cross-dataset evaluation; (<b>b</b>) ROC curve representing the model’s performance for the cross-dataset evaluation. The dashed blue line represents the performance of a random classifier, where predictions are no better than chance, with equal true and false positive rates.</p>
Full article ">
Back to TopTop