Nothing Special   »   [go: up one dir, main page]

Previous Issue
Volume 27, January
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
entropy-logo

Journal Browser

Journal Browser

Entropy, Volume 27, Issue 2 (February 2025) – 119 articles

  • 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
Section
Select all
Export citation of selected articles as:
16 pages, 2475 KiB  
Article
Control of Inhibition-Stabilized Oscillations in Wilson-Cowan Networks with Homeostatic Plasticity
by Camille Godin, Matthew R. Krause, Pedro G. Vieira, Christopher C. Pack and Jean-Philippe Thivierge
Entropy 2025, 27(2), 215; https://doi.org/10.3390/e27020215 - 19 Feb 2025
Abstract
Interactions between excitatory and inhibitory neurons in the cerebral cortex give rise to different regimes of activity and modulate brain oscillations. A prominent regime in the cortex is the inhibition-stabilized network (ISN), defined by strong recurrent excitation balanced by inhibition. While theoretical models [...] Read more.
Interactions between excitatory and inhibitory neurons in the cerebral cortex give rise to different regimes of activity and modulate brain oscillations. A prominent regime in the cortex is the inhibition-stabilized network (ISN), defined by strong recurrent excitation balanced by inhibition. While theoretical models have captured the response of brain circuits in the ISN state, their connectivity is typically hard-wired, leaving unanswered how a network may self-organize to an ISN state and dynamically switch between ISN and non-ISN states to modulate oscillations. Here, we introduce a mean-rate model of coupled Wilson-Cowan equations, link ISN and non-ISN states to Kolmogorov-Sinai entropy, and demonstrate how homeostatic plasticity (HP) allows the network to express both states depending on its level of tonic activity. This mechanism enables the model to capture a broad range of experimental effects, including (i) a paradoxical decrease in inhibitory activity, (ii) a phase offset between excitation and inhibition, and (iii) damped gamma oscillations. Further, the model accounts for experimental work on asynchronous quenching, where an external input suppresses intrinsic oscillations. Together, findings show that oscillatory activity is modulated by the dynamical regime of the network under the control of HP, thus advancing a framework that bridges neural dynamics, entropy, oscillations, and synaptic plasticity. Full article
(This article belongs to the Section Entropy and Biology)
Show Figures

Figure 1

Figure 1
<p>Mean-rate model of excitatory and inhibitory neurons exhibiting different dynamical regimes. (<b>a</b>) Wilson-Cowan circuit where a population of excitatory (E) neurons is coupled with inhibitory (I) neurons. Tonic activation is evenly applied to both populations. (<b>b</b>) The emergence of different dynamical regimes depends on <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>E</mi> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>E</mi> <mi>I</mi> </mrow> </msub> <mo>.</mo> </mrow> </semantics></math> Weak self-excitation (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>E</mi> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> &lt; 1) results in a stable non-ISN regime, while stronger <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>E</mi> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> yields either an ISN or unstable state.</p>
Full article ">Figure 2
<p>Homeostatic plasticity admits stable solutions for ISN and non-ISN regimes. The set point of HP (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>s</mi> <mi>e</mi> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math>) and tonic activation (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>I</mi> </mrow> <mrow> <mi>e</mi> <mi>x</mi> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math>) admit solutions (delineated by black and white dashed lines) that are stable and respect Dale’s law. Black and white circles provide an instance of each regime for recurrent excitation (<b>a</b>) and feedforward inhibition (<b>b</b>). For the parameters corresponding to the white circle in panel “a”, synaptic strengths settle to an ISN regime (<b>c</b>).</p>
Full article ">Figure 3
<p>Paradoxical deactivation of inhibitory cells in the ISN regime. A model with no plasticity captures the well-known paradoxical response observed in ISNs (<b>a</b>). With HP, the strength of tonic activation determines the resulting coupling between E and I populations (<b>b</b>). While weak tonic activation results in an ISN regime exhibiting a paradoxical response, strong tonic activation yields a non-ISN regime with no such response (<b>c</b>). The change in firing rate (<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> </mrow> </semantics></math> rate) from baseline to stimulation shows combinations of excitatory and inhibitory couplings where the paradoxical response is strongest (<b>d</b>). Filled grey circle: instance of a non-ISN state; filled black circle: ISN state.</p>
Full article ">Figure 4
<p>Phase offset between E and I populations in response to an external input. With a forced external oscillator, a tonic activation applied to I cells results in a large phase offset in a non-ISN state and a small offset in an ISN state (<b>a</b>). The phase offset (<math display="inline"><semantics> <mrow> <mo>∆</mo> </mrow> </semantics></math> phase) between E and I populations depends on the strength of excitatory and inhibitory couplings, which collectively determine the state of the network (<b>b</b>). Filled grey circle: non-ISN; filled black circle: ISN.</p>
Full article ">Figure 5
<p>Damped oscillations in the ISN state. Damped oscillations are present in the ISN but not in the non-ISN regime (<b>a</b>), as shown by power spectra in both regimes (<b>b</b>). Mean gamma (30–50 Hz) power (<b>c</b>) and phase offset (<b>d</b>) increase with stronger excitatory coupling. In panels (<b>a</b>–<b>d</b>), damped oscillations were obtained by setting the decay rate of activity to <math display="inline"><semantics> <mrow> <mi>α</mi> </mrow> </semantics></math> = 0.25. Weights were set to <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>E</mi> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> = 0.5 (non-ISN) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>E</mi> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> = 1.5 (ISN), <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>E</mi> <mi>I</mi> </mrow> </msub> </mrow> </semantics></math> = −1.5, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>I</mi> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> = 1.5, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>J</mi> </mrow> <mrow> <mi>I</mi> <mi>I</mi> </mrow> </msub> </mrow> </semantics></math> = −1.1.</p>
Full article ">Figure 6
<p>Asynchronous quenching of damped gamma oscillations. An external oscillator (amplitude: 0.045) was injected into both E and I cells of an ISN that produced damped oscillations (<b>a</b>). When the input matched the frequency of the damped oscillation, sustained activation was generated (top). A mismatched frequency yielded damped oscillations that decayed rapidly (bottom) (<b>b</b>). Summary of the effect of input frequency on the mean activity of E cells taken over a 500 ms window (<b>c</b>).</p>
Full article ">
6 pages, 206 KiB  
Article
Lacunary Series and Strong Approximation
by István Berkes
Entropy 2025, 27(2), 214; https://doi.org/10.3390/e27020214 - 19 Feb 2025
Abstract
Strong approximation, introduced by Strassen (1964), is one of the most powerful methods to prove limit theorems in probability and statistics. In this paper we use strong approximation of lacunary series with conditionally independent sequences to prove uniform and permutation-invariant limit theorems for [...] Read more.
Strong approximation, introduced by Strassen (1964), is one of the most powerful methods to prove limit theorems in probability and statistics. In this paper we use strong approximation of lacunary series with conditionally independent sequences to prove uniform and permutation-invariant limit theorems for such series. Full article
(This article belongs to the Special Issue The Random Walk Path of Pál Révész in Probability)
25 pages, 5127 KiB  
Article
Fault Root Cause Analysis Based on Liang–Kleeman Information Flow and Graphical Lasso
by Xiangdong Liu, Jie Liu, Xiaohua Yang, Zhiqiang Wu, Ying Wei, Zhuoran Xu and Juan Wen
Entropy 2025, 27(2), 213; https://doi.org/10.3390/e27020213 - 19 Feb 2025
Viewed by 28
Abstract
Root cause analysis is used to find the specific fault location and cause of a fault during system fault diagnosis. It is an important step in fault diagnosis. The root cause analysis method based on causality starts from the origin of the causal [...] Read more.
Root cause analysis is used to find the specific fault location and cause of a fault during system fault diagnosis. It is an important step in fault diagnosis. The root cause analysis method based on causality starts from the origin of the causal connection between transactions and infers the location and cause of the mechanism failure by analyzing the causal impact of variables between systems, which has methodological advantages. Causal analysis methods based on transfer entropy are proven to have biases in calculation results, so there is a phenomenon of calculating false causal relationships, which leads to the problem of insufficient accuracy in root cause analysis. Liang–Kleeman information flow (LKIF) is a kind of information entropy that can effectively carry out causal inference, which can avoid obtaining wrong causal relationships. We propose a root cause analysis method that combines graphical lasso and information flow. In view of the large amount of redundant information in industrial data due to the coupling effect of industrial systems, graphical lasso (Glasso) is a high-precision dimensionality reduction method suitable for large-scale and high-dimensional datasets. To ensure the timeliness of root cause analysis, graphical lasso uses dimensionality reduction of the data. Then, LKIF is used to calculate the information flow intensity of each relevant variable, infer the causal relationship between the variable pairs, and trace the root cause of the fault. On the Tennessee Eastman simulation platform, root cause analysis was performed on all faults, and two root cause analysis solutions, transfer entropy and information flow, were compared. Experimental results show that the LKIF–Glasso method can effectively detect the root cause of faults and display the propagation of faults throughout the process. It further shows that information flow has a better effect in root cause analysis than transfer entropy. And through the root cause analysis of the step failure of the stripper, the reason why information flow is superior to transfer entropy is explained in detail. Full article
(This article belongs to the Special Issue Entropy-Based Fault Diagnosis: From Theory to Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Implicit variables lead to spurious causality; (<b>b</b>) autoregression leads to spurious causality.</p>
Full article ">Figure 2
<p>(<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> takes 1 and the sequence on the right side of the dotted line to complete the calculation of transfer entropy and obtain the time series graph; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> takes 2 and the sequence on the right side of the dotted line to complete the calculation of transfer entropy and obtain the time series graph.</p>
Full article ">Figure 3
<p>Flowchart of root cause analysis using the proposed method.</p>
Full article ">Figure 4
<p>Iterative Glasso flowchart.</p>
Full article ">Figure 5
<p>TE process flow diagram.</p>
Full article ">Figure 6
<p>Coefficient of variation plot of IDV (7) in each subgroup. The process variables corresponding to the subgroup numbers in the figure correspond to the sorting of subgroup members in <a href="#entropy-27-00213-t007" class="html-table">Table 7</a>: (<b>a</b>) subgroup 1; (<b>b</b>) subgroup 2; (<b>c</b>) subgroup 3; (<b>d</b>) subgroup 4; (<b>e</b>) subgroup 5.</p>
Full article ">Figure 6 Cont.
<p>Coefficient of variation plot of IDV (7) in each subgroup. The process variables corresponding to the subgroup numbers in the figure correspond to the sorting of subgroup members in <a href="#entropy-27-00213-t007" class="html-table">Table 7</a>: (<b>a</b>) subgroup 1; (<b>b</b>) subgroup 2; (<b>c</b>) subgroup 3; (<b>d</b>) subgroup 4; (<b>e</b>) subgroup 5.</p>
Full article ">Figure 7
<p>Normalized information flow.</p>
Full article ">Figure 8
<p>Transfer entropy plot of IDV (7) in each subgroup. The process variables corresponding to the subgroup numbers in the figure correspond to the sorting of subgroup members in <a href="#entropy-27-00213-t007" class="html-table">Table 7</a>: (<b>a</b>) subgroup 1; (<b>b</b>) subgroup 2; (<b>c</b>) subgroup 3; (<b>d</b>) subgroup 4; (<b>e</b>) subgroup 5.</p>
Full article ">Figure 8 Cont.
<p>Transfer entropy plot of IDV (7) in each subgroup. The process variables corresponding to the subgroup numbers in the figure correspond to the sorting of subgroup members in <a href="#entropy-27-00213-t007" class="html-table">Table 7</a>: (<b>a</b>) subgroup 1; (<b>b</b>) subgroup 2; (<b>c</b>) subgroup 3; (<b>d</b>) subgroup 4; (<b>e</b>) subgroup 5.</p>
Full article ">Figure 8 Cont.
<p>Transfer entropy plot of IDV (7) in each subgroup. The process variables corresponding to the subgroup numbers in the figure correspond to the sorting of subgroup members in <a href="#entropy-27-00213-t007" class="html-table">Table 7</a>: (<b>a</b>) subgroup 1; (<b>b</b>) subgroup 2; (<b>c</b>) subgroup 3; (<b>d</b>) subgroup 4; (<b>e</b>) subgroup 5.</p>
Full article ">Figure 9
<p>Actual changes in monitoring variables in TEP: (<b>a</b>) reactor pressure; (<b>b</b>) stripper pressure; (<b>c</b>) stripper temp.</p>
Full article ">
15 pages, 3884 KiB  
Article
Research on Development Progress and Test Evaluation of Post-Quantum Cryptography
by Meng Zhang, Jing Wang, Junsen Lai, Mingfu Dong, Zhenzhong Zhu, Ryan Ma and Jun Yang
Entropy 2025, 27(2), 212; https://doi.org/10.3390/e27020212 - 18 Feb 2025
Viewed by 218
Abstract
With the rapid development of quantum computing technology, traditional cryptographic systems are facing unprecedented challenges. Post-Quantum Cryptography (PQC), as a new cryptographic technology that can resist attacks from quantum computers, has received widespread attention in recent years. This paper first analyzes the threat [...] Read more.
With the rapid development of quantum computing technology, traditional cryptographic systems are facing unprecedented challenges. Post-Quantum Cryptography (PQC), as a new cryptographic technology that can resist attacks from quantum computers, has received widespread attention in recent years. This paper first analyzes the threat of quantum computing to existing cryptographic systems, then introduces in detail the main technical routes of PQC and its standardization process. Then, a testing and evaluation system for PQC is proposed and relevant tests are carried out. Finally, suggestions for future development are put forward. Full article
(This article belongs to the Special Issue Quantum Information: Working towards Applications)
Show Figures

Figure 1

Figure 1
<p>The security threat of quantum computing to traditional encryption system.</p>
Full article ">Figure 2
<p>PQC Testing and Evaluation System.</p>
Full article ">Figure 3
<p>Test topology diagram.</p>
Full article ">Figure 4
<p>Results of TLSv1.3 handshake. (<b>a</b>) Client Hello protocol message. (<b>b</b>) Server Hello protocol message.</p>
Full article ">Figure 5
<p>End-to-end latency results.</p>
Full article ">Figure 6
<p>Results of connections per second. (<b>a</b>) Number of established connections per second. (<b>b</b>) Number of failed connections per second.</p>
Full article ">Figure 7
<p>Results of traffic. (<b>a</b>) Results of downstream traffic. (<b>b</b>) Results of average latency.</p>
Full article ">
11 pages, 1005 KiB  
Article
OTFS Radar Waveform Design Based on Information Theory
by Qilong Miao, Ling Kuang, Ge Zhang and Yu Shao
Entropy 2025, 27(2), 211; https://doi.org/10.3390/e27020211 - 17 Feb 2025
Viewed by 127
Abstract
In this work, we consider the waveform design for radar systems based on orthogonal time–frequency space (OTFS). The conditional mutual information (CMI), chosen as a promising metric for assessing the radar cognitive capability, serves as the criterion for OTFS waveform design. After formulating [...] Read more.
In this work, we consider the waveform design for radar systems based on orthogonal time–frequency space (OTFS). The conditional mutual information (CMI), chosen as a promising metric for assessing the radar cognitive capability, serves as the criterion for OTFS waveform design. After formulating the OTFS waveform design problem based on maximizing CMI, we propose an equivalent waveform processing approach by minimizing the autocorrelation sidelobes and cross-correlations (ASaCC) of the OTFS transmitting matrix. Simulation results demonstrate that superior performance in target information extraction is achieved by the optimized OTFS waveforms compared to random waveforms. Full article
(This article belongs to the Section Information Theory, Probability and Statistics)
Show Figures

Figure 1

Figure 1
<p>Block diagram of OTFS-based radar.</p>
Full article ">Figure 2
<p>Multi-CAN results for <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>64</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> of (<b>a</b>) autocorrelation coefficients, and (<b>b</b>) cross-correlation coefficients normalized by the maximum value of the autocorrelation coefficients.</p>
Full article ">Figure 3
<p>CMI corresponds to the designed <math display="inline"><semantics> <mrow> <mi>χ</mi> </mrow> </semantics></math> of different iterations.</p>
Full article ">
26 pages, 3014 KiB  
Article
Beamforming Design for STAR-RIS-Assisted NOMA with Binary and Coupled Phase-Shifts
by Yongfei Liu, Yuhuan Wang and Weizhang Xu
Entropy 2025, 27(2), 210; https://doi.org/10.3390/e27020210 - 17 Feb 2025
Viewed by 140
Abstract
This paper investigates the joint optimization of active and passive beamforming in simultaneously transmitting and reflecting reconfigurable intelligent surface (STAR-RIS)-assisted non-orthogonal multiple access (NOMA) systems, with the aim of maximizing system throughput and improving overall performance. To achieve this goal, we propose an [...] Read more.
This paper investigates the joint optimization of active and passive beamforming in simultaneously transmitting and reflecting reconfigurable intelligent surface (STAR-RIS)-assisted non-orthogonal multiple access (NOMA) systems, with the aim of maximizing system throughput and improving overall performance. To achieve this goal, we propose an iterative and efficient algorithmic framework. For active beamforming optimization, the fractional programming (FP) method is employed to reformulate the non-convex optimization problem into a convex problem, making it more tractable. Additionally, Nesterov’s extrapolation technique is introduced to enhance the convergence rate and reduce computational overhead. For the phase optimization of the STAR-RIS, a binary phase design method is proposed, which reformulates the binary phase optimization problem as a segmentation problem on the unit circle. This approach enables a closed form solution that can be derived in linear time. Simulation results demonstrate that the proposed algorithmic framework outperforms existing benchmark algorithms in terms of both system throughput and computational efficiency, verifying its effectiveness and practicality in STAR-RIS-assisted NOMA systems. Full article
(This article belongs to the Special Issue Entropy and Time–Frequency Signal Processing)
Show Figures

Figure 1

Figure 1
<p>System model of a STAR-RIS-assisted NOMA network.</p>
Full article ">Figure 2
<p>The binary phase beamforming.</p>
Full article ">Figure 3
<p>Sum-rate performance over iterations for <math display="inline"><semantics> <mrow> <mi>K</mi> <mo>=</mo> <mn>6</mn> <mo>,</mo> <mi>M</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Sum-rate performance by number of elements.</p>
Full article ">Figure 5
<p>Sum-rate performance according to user numbers.</p>
Full article ">Figure 6
<p>Amplitude control of transmitted and reflected signals.</p>
Full article ">Figure 7
<p>Performance comparison: average sum rate vs. element counts and runtime.</p>
Full article ">Figure 8
<p>Performance comparison: average sum rate vs. number of users and runtime.</p>
Full article ">
19 pages, 4812 KiB  
Article
Exploring Causal Network Complexity in Industrial Linkages: A Comparative Study
by Yongmei Ding, Chao Huang and Xubo Feng
Entropy 2025, 27(2), 209; https://doi.org/10.3390/e27020209 - 17 Feb 2025
Viewed by 168
Abstract
Industrial linkages play a crucial role in sustaining industrial agglomerations, driving economic growth, and shaping the spatial architecture of economic systems. This study explores the complexity of causal networks within the industrial ecosystems of China and the United States, using high-frequency economic data [...] Read more.
Industrial linkages play a crucial role in sustaining industrial agglomerations, driving economic growth, and shaping the spatial architecture of economic systems. This study explores the complexity of causal networks within the industrial ecosystems of China and the United States, using high-frequency economic data to compare the interdependencies and causal structures across key sectors. By employing the partial cross mapping (PCM) technique, we capture the dynamic interactions and intricate linkages among industries, providing a detailed analysis of inter-industry causality. Utilizing data from 32 Chinese industries and 11 United States industries spanning 2015 to 2023, our findings reveal that the United States, as a global leader in technology and finance, exhibits a diversified and service-oriented industrial structure, where financial and technology sectors are pivotal to economic propagation. In contrast, China’s industrial network shows higher centrality in heavy industries and manufacturing sectors, underscoring its focus on industrial output and export-led growth. A comparative analysis of the network topology and resilience highlights that China’s industrial structure enhances network stability and interconnectivity, fostering robust inter-industry linkages, whereas the limited nodal points in the United States network constrain its resilience. These insights into causal network complexity offer a comprehensive perspective on the structural dynamics and resilience of the economic systems in both countries. Full article
(This article belongs to the Section Complexity)
Show Figures

Figure 1

Figure 1
<p>Heat map of PCM causal intensity for the United States industry in 2022 (the size of the circle represents the degree of correlation).</p>
Full article ">Figure 2
<p>Causal networks of industries for China and the United States (2015–2023).</p>
Full article ">Figure 2 Cont.
<p>Causal networks of industries for China and the United States (2015–2023).</p>
Full article ">Figure 2 Cont.
<p>Causal networks of industries for China and the United States (2015–2023).</p>
Full article ">Figure 3
<p>Radar chart for network topology features by year.</p>
Full article ">Figure 4
<p>Comparison of network topology features of China and the United States.</p>
Full article ">Figure 5
<p>Causal network diagram of PCM of China’s Shenwan industries.</p>
Full article ">Figure 6
<p>Global network connectivity efficiency.</p>
Full article ">Figure 7
<p>Maximum connectivity subgraph scale efficiency.</p>
Full article ">
28 pages, 6007 KiB  
Article
Improving the CRCC-DHR Reliability: An Entropy-Based Mimic-Defense-Resource Scheduling Algorithm
by Xinghua Wu, Mingzhe Wang, Yun Cai, Xiaolin Chang and Yong Liu
Entropy 2025, 27(2), 208; https://doi.org/10.3390/e27020208 - 16 Feb 2025
Viewed by 151
Abstract
With more China railway business information systems migrating to the China Railway Cloud Center (CRCC), the attack surface is expanding and there are increasing security threats for the CRCC to deal with. Cyber Mimic Defense (CMD) technology, as an active defense strategy, can [...] Read more.
With more China railway business information systems migrating to the China Railway Cloud Center (CRCC), the attack surface is expanding and there are increasing security threats for the CRCC to deal with. Cyber Mimic Defense (CMD) technology, as an active defense strategy, can counter these threats by constructing a Dynamic Heterogeneous Redundancy (DHR) architecture. However, there are at least two challenges posed to the DHR deployment, namely, the limited number of available schedulable heterogeneous resources and memorization-based attacks. This paper aims to address these two challenges to improve the CRCC-DHR reliability and then facilitate the DHR deployment. By reliability, we mean that the CRCC-DHR with the limited number of available heterogeneous resources can effectively resist memorization-based attacks. We first propose three metrics for assessing the reliability of the CRCC-DHR architecture. Then, we propose an incomplete-information-based game model to capture the relationships between attackers and defenders. Finally, based on the proposed metrics and the captured relationship, we propose a redundant-heterogeneous-resources scheduling algorithm, called the Entropy Weight Scheduling Algorithm (REWS). We evaluate the capability of REWS with the three existing algorithms through simulations. The results show that REWS can achieve a better reliability than the other algorithms. In addition, REWS demonstrates a lower time complexity compared with the existing algorithms. Full article
(This article belongs to the Special Issue Entropy Method for Decision Making with Uncertainty)
Show Figures

Figure 1

Figure 1
<p>The architecture of the CRCC.</p>
Full article ">Figure 2
<p>Dynamic Heterogeneous Redundancy model structures.</p>
Full article ">Figure 3
<p>Attack chain.</p>
Full article ">Figure 4
<p>(<b>a</b>) Trend of information entropy loss during the scheduling of redundant systems under infinite resources. (<b>b</b>) Trend of information entropy loss during the scheduling of redundant systems under finite resources.</p>
Full article ">Figure 5
<p>The flow chart of the algorithm.</p>
Full article ">Figure 6
<p>(<b>a</b>) Weights and scheduling time functions for <math display="inline"><semantics> <mi>η</mi> </semantics></math> of 1. (<b>b</b>) Weights and scheduling times functions for <math display="inline"><semantics> <mi>η</mi> </semantics></math> of 0.25.</p>
Full article ">Figure 7
<p>(<b>a</b>) CRS algorithm scheduling period for m = 3. (<b>b</b>) HCDC algorithm scheduling period for <span class="html-italic">m</span> = 3. (<b>c</b>) HHAC algorithm scheduling period for <span class="html-italic">m</span> = 3. (<b>d</b>) REWS algorithm scheduling period for <span class="html-italic">m</span> = 3.</p>
Full article ">Figure 8
<p>(<b>a</b>) Average number of scheduling states for the CRS algorithm for <span class="html-italic">m</span> = 3. (<b>b</b>) Average number of scheduling states for the HCDC algorithm for <span class="html-italic">m</span> = 3. (<b>c</b>) Average number of scheduling states for the HHAC algorithm for <span class="html-italic">m</span> = 3. (<b>d</b>) Average number of scheduling states for the REWS algorithm for m = 3.</p>
Full article ">Figure 8 Cont.
<p>(<b>a</b>) Average number of scheduling states for the CRS algorithm for <span class="html-italic">m</span> = 3. (<b>b</b>) Average number of scheduling states for the HCDC algorithm for <span class="html-italic">m</span> = 3. (<b>c</b>) Average number of scheduling states for the HHAC algorithm for <span class="html-italic">m</span> = 3. (<b>d</b>) Average number of scheduling states for the REWS algorithm for m = 3.</p>
Full article ">Figure 9
<p>(<b>a</b>) CRS algorithm scheduling period for <span class="html-italic">m</span> = 4. (<b>b</b>) HCDC algorithm scheduling period for <span class="html-italic">m</span> = 4. (<b>c</b>) HHAC algorithm scheduling period for <span class="html-italic">m</span> = 4. (<b>d</b>) REWS algorithm scheduling period for <span class="html-italic">m</span> = 4.</p>
Full article ">Figure 10
<p>(<b>a</b>) Average number of scheduling states for the CRS algorithm for <span class="html-italic">m</span> = 4. (<b>b</b>) Average number of scheduling states for the HCDC algorithm for <span class="html-italic">m</span> = 4. (<b>c</b>) Average number of scheduling states for the HHAC algorithm for <span class="html-italic">m</span> = 4. (<b>d</b>) Average number of scheduling states for the REWS algorithm for <span class="html-italic">m</span> = 4.</p>
Full article ">Figure 11
<p>(<b>a</b>) Experiments on different algorithms against memorization-based attacks for <span class="html-italic">m</span> = 3. (<b>b</b>) Experiments on different algorithms against memorization-based attacks for <span class="html-italic">m</span> = 4.</p>
Full article ">
60 pages, 720 KiB  
Article
New Formulas of Feedback Capacity for AGN Channels with Memory: A Time-Domain Sufficient Statistic Approach
by Charalambos D. Charalambous, Christos Kourtellaris and Stelios Louka
Entropy 2025, 27(2), 207; https://doi.org/10.3390/e27020207 - 15 Feb 2025
Viewed by 243
Abstract
Recently, several papers identified technical issues related to equivalent time-domain and frequency-domain “characterization of the nblock or transmission” feedback capacity formula and its asymptotic limit, the feedback capacity, of additive Gaussian noise (AGN) channels, first introduce by Cover and Pombra in [...] Read more.
Recently, several papers identified technical issues related to equivalent time-domain and frequency-domain “characterization of the nblock or transmission” feedback capacity formula and its asymptotic limit, the feedback capacity, of additive Gaussian noise (AGN) channels, first introduce by Cover and Pombra in 1989 (IEEE Transactions on Information Theory). The main objective of this paper is to derive new results on the Cover and Pombra characterization of the nblock feedback capacity formula, and to clarify the main points of confusion regarding the time-domain results that appeared in the literature. The first part of this paper derives new equivalent time-domain sequential characterizations of feedback capacity of AGN channels driven by non-stationary and non-ergodic Gaussian noise. It is shown that the optimal channel input processes of the new equivalent sequential characterizations are expressed as functionals of a sufficient statistic and a Gaussian orthogonal innovations process. Further, the Cover and Pombra nblock capacity formula is expressed as a functional of two generalized matrix difference Riccati equations (DREs) of the filtering theory of Gaussian systems, contrary to results that appeared in the literature and involve only one DRE. It is clarified that prior literature deals with a simpler problem that presupposes the state of the noise is known to the encoder and the decoder. In the second part of this paper, the existence of the asymptotic limit of the nblock feedback capacity formula is shown to be equivalent to the convergence properties of solutions of the two generalized DREs. Further, necessary and or sufficient conditions are identified for the existence of asymptotic limits, for stable and unstable Gaussian noise, when the optimal input distributions are asymptotically time-invariant but not necessarily stationary. This paper contains an in-depth analysis, with various examples, and identifies the technical conditions on the feedback code and state space noise realization, so that the time-domain capacity formulas that appeared in the literature, for AGN channels with stationary noises, are indeed correct. Full article
(This article belongs to the Section Information Theory, Probability and Statistics)
18 pages, 1334 KiB  
Article
Transient Dynamics and Homogenization in Incoherent Collision Models
by Göktuğ Karpat and Barış Çakmak
Entropy 2025, 27(2), 206; https://doi.org/10.3390/e27020206 - 15 Feb 2025
Viewed by 204
Abstract
Collision models have attracted significant attention in recent years due to their versatility to simulate open quantum systems in different dynamical regimes. They have been used to study various interesting phenomena such as the dynamical emergence of non-Markovian memory effects and the spontaneous [...] Read more.
Collision models have attracted significant attention in recent years due to their versatility to simulate open quantum systems in different dynamical regimes. They have been used to study various interesting phenomena such as the dynamical emergence of non-Markovian memory effects and the spontaneous establishment of synchronization in open quantum systems. In such models, the repeated pairwise interactions between the system and the environment and also the possible coupling between different environmental units are typically modeled using the coherent partial SWAP (PSWAP) operation as it is known to be a universal homogenizer. In this study, we investigate the dynamical behavior of incoherent collision models, where the interactions between different units are modeled by the incoherent controlled SWAP (CSWAP) operation, which is also a universal homogenizer. Even though the asymptotic dynamics of the open system in cases of both coherent and incoherent swap interactions appear to be identical, its transient dynamics turns out to be significantly different. Here, we present a comparative analysis of the consequences of having coherent or incoherent couplings in collision models, namely, PSWAP or CSWAP interactions, respectively, for the emergence of memory effects for a single-qubit system and for the onset synchronization between a pair of qubits, both of which are strictly determined by the transient dynamics of the open system. Full article
(This article belongs to the Special Issue Simulation of Open Quantum Systems)
Show Figures

Figure 1

Figure 1
<p>For the coherent PSWAP interaction between the system qubit s and the environment qubits <math display="inline"><semantics> <msub> <mi>e</mi> <mi>i</mi> </msub> </semantics></math> without intra–environment couplings, <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>e</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, (<b>a</b>) shows the evolution of the coherence <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>(</mo> <msub> <mi>ρ</mi> <mi>s</mi> </msub> <mo>)</mo> </mrow> </semantics></math> in the open system, the entropy <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>(</mo> <msub> <mi>ρ</mi> <mi>s</mi> </msub> <mo>)</mo> </mrow> </semantics></math> of the open system, and the fidelity <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>(</mo> <msub> <mi>ρ</mi> <mi>s</mi> </msub> <mo>,</mo> <msub> <mi>ρ</mi> <mi>e</mi> </msub> <mo>)</mo> </mrow> </semantics></math> between the open system and the initial state of the environment qubits <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mi>e</mi> </msub> <mo>=</mo> <mrow> <mo>|</mo> <mn>0</mn> <mo>〉</mo> <mo>〈</mo> <mn>0</mn> <mo>|</mo> </mrow> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>1100</mn> </mrow> </semantics></math> collisions. (<b>b</b>) displays the path of the open system state s through the Bloch ball starting from the state <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mi>s</mi> </msub> <mo>=</mo> <mrow> <mo>|</mo> <mo>+</mo> <mo>〉</mo> <mo>〈</mo> <mo>+</mo> <mo>|</mo> </mrow> </mrow> </semantics></math>. On the other hand, (<b>c</b>) and (<b>d</b>) display the same set of plots as in (<b>a</b>) and (<b>b</b>) when the interaction between the system and the environment is described by the incoherent CSWAP coupling with the same interaction parameters, that is, <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>e</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>While (<b>a</b>) shows the dynamics of the trace distance for the PSWAP–PSWAP collision model, together with the evolution of the non-Markovianity measure <math display="inline"><semantics> <msub> <mi mathvariant="script">N</mi> <mi>D</mi> </msub> </semantics></math> shown in the inset for <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>1200</mn> </mrow> </semantics></math> collisions, (<b>c</b>) displays the same set of plots in the case of the PSWAP–CSWAP collision model for same number of collisions. We take the initial system state pair as <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mo>±</mo> <mo>〉</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1</mn> <mo>/</mo> <msqrt> <mn>2</mn> </msqrt> <mo>)</mo> </mrow> <mrow> <mo>(</mo> <mo>|</mo> <mn>0</mn> <mo>〉</mo> </mrow> <mo>±</mo> <mrow> <mo>|</mo> <mn>1</mn> <mo>〉</mo> <mo>)</mo> </mrow> </mrow> </semantics></math>, and set the system–environment and the intra-environment coupling strengths identically in both models as <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>e</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.93</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. In (<b>b</b>) and (<b>d</b>), we demonstrate the paths of evolution of the Bloch vectors through the Bloch ball for the PSWAP–PSWAP and PSWAP–CSWAP models, respectively, starting from the initial system state <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mi>s</mi> </msub> <mo>=</mo> <mrow> <mo>|</mo> <mo>+</mo> <mo>〉</mo> <mo>〈</mo> <mo>+</mo> <mo>|</mo> </mrow> </mrow> </semantics></math>, with the same interaction parameters.</p>
Full article ">Figure 3
<p>Non-Markovianity diagrams for (<b>a</b>) the PSWAP-PSWAP and (<b>b</b>) the PSWAP-CSWAP collision models in terms of the system–environment and the intra-environment interaction parameters, <math display="inline"><semantics> <msub> <mi>γ</mi> <mrow> <mi>s</mi> <mi>e</mi> </mrow> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>γ</mi> <mrow> <mi>e</mi> <mi>e</mi> </mrow> </msub> </semantics></math>. For both models, we simulate the dynamics for <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>12,000</mn> </mrow> </semantics></math> iterations, and the state pair used in the calculation of the non-Markovianity measure <math display="inline"><semantics> <msub> <mi mathvariant="script">N</mi> <mi>D</mi> </msub> </semantics></math> is fixed as <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mo>±</mo> <mo>〉</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1</mn> <mo>/</mo> <msqrt> <mn>2</mn> </msqrt> <mo>)</mo> </mrow> <mrow> <mo>(</mo> <mo>|</mo> <mn>0</mn> <mo>〉</mo> </mrow> <mo>±</mo> <mrow> <mo>|</mo> <mn>1</mn> <mo>〉</mo> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>While (<b>a</b>) shows the dynamics of the trace distance for the CSWAP–CSWAP collision model, together with the evolution of the non–Markovianity measure <math display="inline"><semantics> <msub> <mi mathvariant="script">N</mi> <mi>D</mi> </msub> </semantics></math> shown in the inset for <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>1200</mn> </mrow> </semantics></math> collisions, (<b>c</b>) displays the same set of plots in the case of the CSWAP–PSWAP collision model for same number of collisions. We take the initial system state pair as <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mo>±</mo> <mo>〉</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1</mn> <mo>/</mo> <msqrt> <mn>2</mn> </msqrt> <mo>)</mo> </mrow> <mrow> <mo>(</mo> <mo>|</mo> <mn>0</mn> <mo>〉</mo> </mrow> <mo>±</mo> <mrow> <mo>|</mo> <mn>1</mn> <mo>〉</mo> <mo>)</mo> </mrow> </mrow> </semantics></math>, and set the system–environment and the intra-environment coupling strengths identically in both models as <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>e</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.93</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. In (<b>b</b>) and (<b>d</b>), we demonstrate the paths of evolution of the Bloch vectors through the Bloch ball for the CSWAP–CSWAP and CSWAP–PSWAP models, respectively, starting from the initial system state <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mi>s</mi> </msub> <mo>=</mo> <mrow> <mo>|</mo> <mo>+</mo> <mo>〉</mo> <mo>〈</mo> <mo>+</mo> <mo>|</mo> </mrow> </mrow> </semantics></math>, with the same interaction parameters.</p>
Full article ">Figure 5
<p>System particles are initialized as <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> <mo>=</mo> <mrow> <mo>|</mo> <mo>+</mo> <mo>〉</mo> <mo>|</mo> <mi>L</mi> <mo>〉</mo> <mo>〈</mo> <mi>L</mi> <mo>|</mo> <mo>〈</mo> <mo>+</mo> <mo>|</mo> </mrow> </mrow> </semantics></math> and resonant such that <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <mn>1</mn> </mrow> </semantics></math>, and both interact with a common environmental unit through a coherent PSWAP having strength <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.03</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. While (<b>a</b>) displays the dynamics of <math display="inline"><semantics> <mrow> <mo>〈</mo> <msubsup> <mi>σ</mi> <msub> <mi>s</mi> <mn>1</mn> </msub> <mi>x</mi> </msubsup> <mo>〉</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>〈</mo> <msubsup> <mi>σ</mi> <msub> <mi>s</mi> <mn>2</mn> </msub> <mi>x</mi> </msubsup> <mo>〉</mo> </mrow> </semantics></math>, (<b>b</b>) shows the corresponding Pearson coefficient <math display="inline"><semantics> <msub> <mi>C</mi> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </semantics></math> between these two data sets settling to <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> signaling anti-synchronization, which is plotted considering data windows of 100 collisions with partial overlaps of 50 collisions for N = 2500. In (<b>c</b>), we show the fidelity F between the state of the environmental units <math display="inline"><semantics> <msub> <mi>ρ</mi> <mi>e</mi> </msub> </semantics></math> and both the local states of the system particles <math display="inline"><semantics> <msub> <mi>ρ</mi> <msub> <mi>s</mi> <mn>1</mn> </msub> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>ρ</mi> <msub> <mi>s</mi> <mn>2</mn> </msub> </msub> </semantics></math> and their global state <math display="inline"><semantics> <msub> <mi>ρ</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> </semantics></math>. We observe a clear convergence towards <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, which indicates homogenization of system particles with the environment.</p>
Full article ">Figure 6
<p>System particles are initialized as <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> <mo>=</mo> <mrow> <mo>|</mo> <mo>+</mo> <mo>〉</mo> <mo>|</mo> <mi>L</mi> <mo>〉</mo> <mo>〈</mo> <mi>L</mi> <mo>|</mo> <mo>〈</mo> <mo>+</mo> <mo>|</mo> </mrow> </mrow> </semantics></math> and resonant such that <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <mn>1</mn> </mrow> </semantics></math>, and both interact with a common environmental unit through an incoherent CSWAP having strength <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.03</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. While (<b>a</b>) displays the dynamics of <math display="inline"><semantics> <mrow> <mo>〈</mo> <msubsup> <mi>σ</mi> <msub> <mi>s</mi> <mn>1</mn> </msub> <mi>x</mi> </msubsup> <mo>〉</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>〈</mo> <msubsup> <mi>σ</mi> <msub> <mi>s</mi> <mn>2</mn> </msub> <mi>x</mi> </msubsup> <mo>〉</mo> </mrow> </semantics></math>, (<b>b</b>) displays the corresponding Pearson coefficient <math display="inline"><semantics> <msub> <mi>C</mi> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </semantics></math> between these two data sets, which is plotted considering data windows of 100 collisions with partial overlaps of 50 collisions for N = 2500, showing no sign of synchronization. In (<b>c</b>), we show the fidelity F between the state of the environmental units <math display="inline"><semantics> <msub> <mi>ρ</mi> <mi>e</mi> </msub> </semantics></math> and both the local states of the system particles <math display="inline"><semantics> <msub> <mi>ρ</mi> <msub> <mi>s</mi> <mn>1</mn> </msub> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>ρ</mi> <msub> <mi>s</mi> <mn>2</mn> </msub> </msub> </semantics></math> and their global state <math display="inline"><semantics> <msub> <mi>ρ</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> </semantics></math>. We observe a clear convergence towards <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, which indicates the homogenization of system particles with the environment.</p>
Full article ">Figure 7
<p>System particles are initialized as <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> <mo>=</mo> <mrow> <mo>|</mo> <mo>+</mo> <mo>〉</mo> <mo>|</mo> <mi>L</mi> <mo>〉</mo> <mo>〈</mo> <mi>L</mi> <mo>|</mo> <mo>〈</mo> <mo>+</mo> <mo>|</mo> </mrow> </mrow> </semantics></math> and resonant such that <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <mn>1</mn> </mrow> </semantics></math>. As the first particle interacts with environmental units through incoherent CSWAP, the second one interacts with the same environment units via a coherent PSWAP interaction, with identical strength <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>s</mi> <mi>e</mi> </mrow> </msub> <mo>=</mo> <mn>0.03</mn> <mrow> <mo>(</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. While (<b>a</b>) displays the dynamics of <math display="inline"><semantics> <mrow> <mo>〈</mo> <msubsup> <mi>σ</mi> <msub> <mi>s</mi> <mn>1</mn> </msub> <mi>x</mi> </msubsup> <mo>〉</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>〈</mo> <msubsup> <mi>σ</mi> <msub> <mi>s</mi> <mn>2</mn> </msub> <mi>x</mi> </msubsup> <mo>〉</mo> </mrow> </semantics></math>, (<b>b</b>) displays the corresponding Pearson coefficient <math display="inline"><semantics> <msub> <mi>C</mi> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </semantics></math> between these two data sets, which is plotted considering data windows of 100 collisions with partial overlaps of 50 collisions for N = 2500, showing no sign of synchronization. In (<b>c</b>), we show the fidelity F between the state of the environmental units <math display="inline"><semantics> <msub> <mi>ρ</mi> <mi>e</mi> </msub> </semantics></math> and both the local states of the system particles <math display="inline"><semantics> <msub> <mi>ρ</mi> <msub> <mi>s</mi> <mn>1</mn> </msub> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>ρ</mi> <msub> <mi>s</mi> <mn>2</mn> </msub> </msub> </semantics></math> and their global state <math display="inline"><semantics> <msub> <mi>ρ</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> </semantics></math>. We observe a clear convergence towards <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, confirming the homogenization of system particles with the environment.</p>
Full article ">
15 pages, 3219 KiB  
Article
Earthquake Forecasting Based on b Value and Background Seismicity Rate in Yunnan Province, China
by Yuchen Zhang, Rui Wang, Haixia Shi, Miao Miao, Jiancang Zhuang, Ying Chang, Changsheng Jiang, Lingyuan Meng, Danning Li, Lifang Liu, Youjin Su, Zhenguo Zhang and Peng Han
Entropy 2025, 27(2), 205; https://doi.org/10.3390/e27020205 - 15 Feb 2025
Viewed by 278
Abstract
Characterized by frequent earthquakes and a dense population, Yunnan Province, China, faces significant seismic hazards and is a hot place for earthquake forecasting research. In a previous study, we evaluated the performance of the b value for 5-year seismic forecasting during 2000–2019 and [...] Read more.
Characterized by frequent earthquakes and a dense population, Yunnan Province, China, faces significant seismic hazards and is a hot place for earthquake forecasting research. In a previous study, we evaluated the performance of the b value for 5-year seismic forecasting during 2000–2019 and made a forward prediction of M ≥ 5.0 earthquakes in 2020–2024. In this study, with the forecast period having passed, we first revisit the results and assess the forward prediction performance. Then, the background seismicity rate, which may also offer valuable long-term forecasting information, is incorporated into earthquake prediction for Yunnan Province. To assess the effectiveness of the prediction, the Molchan Error Diagram (MED), Probability Gain (PG), and Probability Difference (PD) are employed. Using a 25-year catalog, the spatial b value and background seismicity rate across five temporal windows are calculated, and 86 M ≥ 5.0 earthquakes as prediction samples are examined. The predictive performance of the background seismicity rate and b value is comprehensively tested and shown to be useful for 5-year forecasting in Yunnan. The performance of the b value exhibits a positive correlation with the predicted earthquake magnitude. The synergistic effect of combining these two predictors is also revealed. Finally, using the threshold corresponding to the maximum PD, we integrate the forecast information of background seismicity rates and the b value. A forward prediction is derived for the period from January 2025 to December 2029. This study can be helpful for disaster preparedness and risk management in Yunnan Province, China. Full article
(This article belongs to the Special Issue Time Series Analysis in Earthquake Complex Networks)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The <span class="html-italic">b</span> value from January 2015 to December 2019 and earthquakes with M ≥ 5.0 from January 2020 to December 2024. The dot and star are scaled to the magnitude. (<b>b</b>) Temporal distribution of earthquakes in Yunnan Province from January 2000 to December 2024. (<b>c</b>) Temporal distribution of earthquakes in Yunnan Province from January 2020 to December 2024. (<b>d</b>) The MED of forecast performance based on the <span class="html-italic">b</span> value in (<b>a</b>). The marked numbers are the serial numbers in <a href="#entropy-27-00205-t001" class="html-table">Table 1</a>, and the size of the cross markers is scaled to the magnitude.</p>
Full article ">Figure 2
<p>The <span class="html-italic">b</span> value and background seismicity rate. (<b>a</b>–<b>e</b>) <span class="html-italic">b</span> value; (<b>f</b>,<b>j</b>) background seismicity rate. Results in (<b>a</b>,<b>f</b>) using catalog in 2000–2004 and forecasting moderate–large earthquakes in 2005–2009; (<b>b</b>,<b>g</b>) using catalog in 2005–2009 and forecasting moderate–large earthquakes in 2010–2014; (<b>c</b>,<b>h</b>) using catalog in 2010–2014 and forecasting moderate–large earthquakes in 2015–2019; (<b>d</b>,<b>i</b>) using catalog in 2015–2019 and forecasting moderate–large earthquakes in 2020–2024; (<b>e</b>,<b>j</b>) using catalog in 2020–2024. A dot represents an earthquake with 5.0 ≤ M &lt; 5.5. A star represents an earthquake with M ≥ 5.5. The sizes of the dots and stars are scaled to magnitude.</p>
Full article ">Figure 3
<p>Forecast performance based on <span class="html-italic">b</span> value and background seismicity rate during 2005–2024. (<b>a</b>–<b>c</b>) show the results of earthquakes with M ≥ 5.5. (<b>a</b>) MED; (<b>b</b>) <span class="html-italic">PG</span>; (<b>c</b>) <span class="html-italic">PD</span>. (<b>d</b>–<b>f</b>) are the results of earthquakes with M ≥ 5.0. (<b>d</b>) MED; (<b>e</b>) <span class="html-italic">PG</span>; (<b>f</b>) <span class="html-italic">PD</span>. The number of earthquake samples is <math display="inline"><semantics> <mrow> <mo>=</mo> <mn>29</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mi>M</mi> <mo>≥</mo> <mn>5.0</mn> </mrow> </msub> <mo>=</mo> <mn>86</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>The variation in forecast performance with earthquake magnitude. (<b>a</b>) Variation in maximum <span class="html-italic">PG</span> with the forecast magnitude; (<b>b</b>) variation in maximum <span class="html-italic">PD</span> with the forecast magnitude; (<b>c</b>) variation in <span class="html-italic">S</span> with the forecast magnitude.</p>
Full article ">Figure 5
<p>Forecast performance by combining <span class="html-italic">b</span> value and background seismicity rate during 2005–2024. The x-axis is the alarming rate of background seismicity corresponding to <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>t</mi> <mi>h</mi> <mi>r</mi> </mrow> <mrow> <mi>μ</mi> </mrow> </msub> </mrow> </semantics></math>, and the y-axis is the alarming rate of the <span class="html-italic">b</span> value corresponding to <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>t</mi> <mi>h</mi> <mi>r</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math>. (<b>a</b>) <span class="html-italic">PG</span> for M ≥ 5.5 earthquakes; (<b>b</b>) <span class="html-italic">PD</span> for M ≥ 5.5 earthquakes; (<b>c</b>) <span class="html-italic">PG</span> for M ≥ 5.0 earthquakes; (<b>d</b>) <span class="html-italic">PD</span> for M ≥ 5.0 earthquakes. The location of the maximum value (<span class="html-italic">PG</span> or <span class="html-italic">PD</span>) in each figure is marked with dots and detailed in <a href="#entropy-27-00205-t002" class="html-table">Table 2</a>. The cross in (<b>b</b>) is located at the alarming rate corresponding to the maximum <span class="html-italic">PD</span> in <a href="#entropy-27-00205-f003" class="html-fig">Figure 3</a>c.</p>
Full article ">Figure 6
<p>Alarmed regions for the period from January 2025 to December 2029 based on <span class="html-italic">b</span> value and background seismicity rate obtained during 2020–2024. (<b>a</b>) Alarmed area based on <span class="html-italic">b</span> value and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>t</mi> <mi>h</mi> <mi>r</mi> <mo>_</mo> <mi>P</mi> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math>, with 0.38 alarming rate; (<b>b</b>) alarmed area based on background seismicity rate and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>t</mi> <mi>h</mi> <mi>r</mi> <mo>_</mo> <mi>P</mi> <mi>D</mi> </mrow> <mrow> <mi>μ</mi> </mrow> </msub> </mrow> </semantics></math>, with 0.42 alarming rate; (<b>c</b>) alarmed area based on <span class="html-italic">b</span> value and background seismicity rate, with 0.20 alarming rate. The red edge squares show the alarmed grid cells.</p>
Full article ">
27 pages, 3309 KiB  
Article
Quantum Thermometry for Ultra-Low Temperatures Using Probe and Ancilla Qubit Chains
by Asghar Ullah, Vipul Upadhyay and Özgür E. Müstecaplıoğlu
Entropy 2025, 27(2), 204; https://doi.org/10.3390/e27020204 - 14 Feb 2025
Viewed by 316
Abstract
We propose a scheme to enhance the range and precision of ultra-low temperature measurements by employing a probe qubit coupled to a chain of ancilla qubits. Specifically, we analyze a qubit chain governed by Heisenberg XX and Dzyaloshinskii–Moriya (DM) interactions. The precision [...] Read more.
We propose a scheme to enhance the range and precision of ultra-low temperature measurements by employing a probe qubit coupled to a chain of ancilla qubits. Specifically, we analyze a qubit chain governed by Heisenberg XX and Dzyaloshinskii–Moriya (DM) interactions. The precision limits of temperature measurements are characterized by evaluating quantum Fisher information (QFI). Our findings demonstrate that the achievable precision bounds, as well as the number of peaks in the QFI as a function of temperature, can be controlled by adjusting the number of ancilla qubits and the system’s model parameters. These results are interpreted in terms of the influence of energy transitions on the range and the number of QFI peaks as a function of temperature. This study highlights the potential of the probe qubit–ancilla chain system as a powerful and precise tool for quantum thermometry in the ultra-low temperature regime. Full article
(This article belongs to the Special Issue Simulation of Open Quantum Systems)
Show Figures

Figure 1

Figure 1
<p>A schematic representation of our thermometry scheme is shown. The system consists of a probe qubit with a transition frequency <math display="inline"><semantics> <msub> <mi>ω</mi> <mi>p</mi> </msub> </semantics></math> located outside of the thermal sample. This qubit is used as a probe to measure the unknown temperature <span class="html-italic">T</span> of a sample. The measurement is facilitated by ancilla qubits with transition frequencies <math display="inline"><semantics> <msub> <mi>ω</mi> <mi>i</mi> </msub> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> <mo>,</mo> <mo>…</mo> <mo>,</mo> <mi>N</mi> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math>), which are immersed in the sample. The qubits are coupled via a combination of the Heisenberg <math display="inline"><semantics> <mrow> <mi>X</mi> <mi>X</mi> </mrow> </semantics></math> interaction (<math display="inline"><semantics> <msub> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>X</mi> <mi>X</mi> </mrow> </msub> </semantics></math>) with coupling strength <span class="html-italic">J</span>, and DM interaction (<math display="inline"><semantics> <msub> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>D</mi> <mi>M</mi> </mrow> </msub> </semantics></math>) characterized by strength <span class="html-italic">g</span>.</p>
Full article ">Figure 2
<p>(<b>a</b>) Energy transitions induced by the bath when the two qubits are resonant, such as <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <msub> <mi>ω</mi> <mi>a</mi> </msub> </mrow> </semantics></math>. The two transitions <math display="inline"><semantics> <msub> <mi>ω</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>ω</mi> <mn>2</mn> </msub> </semantics></math> are of almost the same order, such as <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mo>∼</mo> <msub> <mi>ω</mi> <mn>2</mn> </msub> </mrow> </semantics></math>. (<b>b</b>) shows the transitions induced when the two qubits are off-resonant, such as (<math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>≠</mo> <msub> <mi>ω</mi> <mi>a</mi> </msub> </mrow> </semantics></math>). In this case, the transition energies <math display="inline"><semantics> <msub> <mi>ω</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>ω</mi> <mn>2</mn> </msub> </semantics></math> are of different orders.</p>
Full article ">Figure 3
<p>The behavior of the first derivative of the population <math display="inline"><semantics> <mrow> <msup> <mi>p</mi> <mo>′</mo> </msup> <mrow> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> of the excited state of the probe qubit as a function of temperature <span class="html-italic">T</span> for (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>. We can see that there is an additional peak at lower temperatures when <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>≠</mo> <msub> <mi>ω</mi> <mi>a</mi> </msub> </mrow> </semantics></math>, while only one peak can be seen for the resonant qubits, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <msub> <mi>ω</mi> <mi>a</mi> </msub> </mrow> </semantics></math>. The parameters are set to <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>. All the system parameters are scaled with the probe qubit frequency <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>(<b>a</b>) The first derivative of the population <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mo>−</mo> <mo>′</mo> </msubsup> <mrow> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> as a function of <span class="html-italic">T</span> in the low-temperature limit, while (<b>b</b>) shows the first derivative of the population <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mo>+</mo> <mo>′</mo> </msubsup> <mrow> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> as a function of <span class="html-italic">T</span> in the high-temperature limit. In both plots, the blue dashed curve represents the exact expression for the first derivative of <span class="html-italic">p</span> as given in Equation (<a href="#FD11-entropy-27-00204" class="html-disp-formula">11</a>), while the solid red curve corresponds to the approximate expressions given in Equations (<a href="#FD14-entropy-27-00204" class="html-disp-formula">14</a>) and (<a href="#FD17-entropy-27-00204" class="html-disp-formula">17</a>), respectively. The parameters are set to <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>. All the system parameters are scaled with the probe qubit frequency <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>(<b>a</b>) QFI <math display="inline"><semantics> <msub> <mi mathvariant="script">F</mi> <mi>Q</mi> </msub> </semantics></math> as a function of an unknown parameter <span class="html-italic">T</span> for different values of coupling strength <span class="html-italic">g</span> at <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>. The solid magenta, orange dashed, and blue dot-dashed curves correspond to <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.03</mn> </mrow> </semantics></math>, respectively. We set <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>. (<b>b</b>) QFI for the resonant qubits case, such as <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <msub> <mi>ω</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>. The solid magenta, orange dashed, and blue dot-dashed curves correspond to <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.15</mn> </mrow> </semantics></math>, respectively. Here, we set <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>0.35</mn> </mrow> </semantics></math>. All the system parameters are scaled with the probe qubit frequency <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>QFI <math display="inline"><semantics> <msub> <mi mathvariant="script">F</mi> <mi>Q</mi> </msub> </semantics></math> as a function of <span class="html-italic">T</span> for <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>. The solid blue curves are obtained using exact QFI in Equation (<a href="#FD20-entropy-27-00204" class="html-disp-formula">20</a>), and the red dashed curve is plotted using approximate QFI in Equation (<a href="#FD24-entropy-27-00204" class="html-disp-formula">24</a>). All the system parameters are scaled with the probe qubit frequency <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>QFI of probe qubit as a function of temperature <span class="html-italic">T</span> for the case of two ancilla qubits attached to the bath (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>); (<b>a</b>–<b>c</b>) correspond to plots for <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.04</mn> <mo>,</mo> <mn>0.06</mn> <mo>,</mo> <mn>0.1</mn> </mrow> </semantics></math>, respectively. For each <math display="inline"><semantics> <msub> <mi>g</mi> <mn>1</mn> </msub> </semantics></math> value, we set <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.06</mn> <mo>,</mo> <mn>0.08</mn> <mo>,</mo> <mn>0.3</mn> </mrow> </semantics></math>, respectively. The remaining parameters are fixed as <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>. The solid red vertical lines indicate the temperature values calculated using Equation (<a href="#FD33-entropy-27-00204" class="html-disp-formula">33</a>) for the eigenvalues of the M matrix defined in Equation (<a href="#FD27-entropy-27-00204" class="html-disp-formula">27</a>). All the system parameters are scaled with the probe qubit frequency <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>QFI of the probe qubit as a function of temperature <span class="html-italic">T</span> on a log–log scale is shown for the case of three ancilla qubits attached to the bath (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>); (<b>a</b>–<b>c</b>) correspond to the plots for <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.0055</mn> <mo>,</mo> <mn>0.006</mn> <mo>,</mo> <mn>0.0065</mn> </mrow> </semantics></math>, respectively. For each <math display="inline"><semantics> <msub> <mi>g</mi> <mn>1</mn> </msub> </semantics></math> value, we set <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.0075</mn> <mo>,</mo> <mn>0.008</mn> <mo>,</mo> <mn>0.0085</mn> </mrow> </semantics></math>. respectively. The rest of the parameters are fixed to <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.004</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.06</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.08</mn> </mrow> </semantics></math>. The solid blue vertical lines indicate the temperature values calculated using Equation (<a href="#FD33-entropy-27-00204" class="html-disp-formula">33</a>) for the eigenvalues of the matrix in Equation (<a href="#FD27-entropy-27-00204" class="html-disp-formula">27</a>) associated with each peak. All the system parameters are scaled with the probe qubit frequency <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p><b>Top row:</b> (<b>a</b>) Transition energies <span class="html-italic">E</span> as a function of coupling strength <math display="inline"><semantics> <msub> <mi>g</mi> <mn>2</mn> </msub> </semantics></math> and (<b>b</b>) the corresponding QFI as a function of temperature <span class="html-italic">T</span> for <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, with the following parameter set: <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.004</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.007</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.06</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.005</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>. For plotting QFI, we consider the weak coupling strength of <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>. <b>Bottom row:</b> (<b>c</b>) Transition energies <span class="html-italic">E</span> as a function of coupling strength <math display="inline"><semantics> <msub> <mi>g</mi> <mn>2</mn> </msub> </semantics></math> and (<b>d</b>) the corresponding QFI as a function of temperature <span class="html-italic">T</span> for <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, with the following parameter set: <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.004</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.006</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.004</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>. For plotting QFI, we consider a strong coupling strength of <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>QFI of the probe qubit as a function of temperature <span class="html-italic">T</span> in the case of four ancilla qubits attached to the bath (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>). (<b>a</b>–<b>c</b>) represent the plots for <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.0005</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.00055</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.0007</mn> </mrow> </semantics></math>, respectively. The rest of the parameters are fixed to <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>4</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.004</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.0004</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>4</mn> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.06</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.005</mn> </mrow> </semantics></math>. The values of <span class="html-italic">J</span> are fixed as follows: <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>4</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.08</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.008</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.00095</mn> </mrow> </semantics></math>. The dashed blue vertical lines indicate the temperature values associated with each peak. All the system parameters are scaled with the probe qubit frequency <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure A1
<p>Fisher information calculated using (<a href="#FD19-entropy-27-00204" class="html-disp-formula">19</a>) and the optimal Fisher information based on the <math display="inline"><semantics> <msup> <mi>σ</mi> <mi>z</mi> </msup> </semantics></math> measurement, both plotted as functions of temperature <span class="html-italic">T</span>. The parameters are set to <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics></math>. All the system parameters are scaled with the probe qubit frequency <math display="inline"><semantics> <mrow> <msub> <mi>ω</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">
20 pages, 844 KiB  
Article
Maximum Entropy-Minimum Residual Model: An Optimum Solution to Comprehensive Evaluation and Multiple Attribute Decision Making
by Qi-Yi Tang and Yu-Xuan Lin
Entropy 2025, 27(2), 203; https://doi.org/10.3390/e27020203 - 14 Feb 2025
Viewed by 209
Abstract
To assess a subject with multiple factors or attributes, a comprehensive evaluation index, or say a composite indicator, is often constructed to make a holistic judgement. The key problem is to assign weights to the factors. There are various weighting methods in the [...] Read more.
To assess a subject with multiple factors or attributes, a comprehensive evaluation index, or say a composite indicator, is often constructed to make a holistic judgement. The key problem is to assign weights to the factors. There are various weighting methods in the literature, but a gold standard is lacking. Some weighting methods may lead to a trivial weight assignment that is one factor having a weight equal to 1 and the others all zero, while some methods generate a solution contradicting intuitive judgement, or even infeasible to calculate. This paper proposes a new model to generate weights based on the maximum entropy-minimum residual (MEMR) principle, directly estimating the relationship between factor weights and the composite indicator. The MEMR composite indicator extracts the common feature of multiple factors while preserving their diversity. This paper compares the MEMR model with other commonly used weighting methods in various case studies. The MEMR model has more robust, consistent, and interpretable results than others and is suitable for all comprehensive evaluation cases involving quantitative factors. The optimization technique of the proposed MEMR model and the related statistical tests are included as a package in the DPS (data processing system) software V21.05 for the convenience of application in all fields. Full article
(This article belongs to the Special Issue Number Theoretic Methods in Statistics: Theory and Applications)
Show Figures

Figure 1

Figure 1
<p>Weight coefficients of various weighting methods for the technological achievement index data.</p>
Full article ">Figure 2
<p>Relationship diagram of composite indicator estimates and factor weights and normalized values for different methods: (<b>a</b>) The relationships of MEMR CEI, factor weights, and normalized values. (<b>b</b>) The relationships of ET CEI, factor weights, and normalized values. (<b>c</b>) The relationships of CRITIC CEI, factor weights, and normalized values. (<b>d</b>) The relationships of SD CEI, factor weights, and normalized values.</p>
Full article ">
15 pages, 1714 KiB  
Article
SAluMC: Thwarting Side-Channel Attacks via Random Number Injection in RISC-V
by Shibo Dang, Yunlong Shao, Zhida Li, Adetokunbo Makanju and Thomas Aaron Gulliver
Entropy 2025, 27(2), 202; https://doi.org/10.3390/e27020202 - 14 Feb 2025
Viewed by 347
Abstract
As processor performance advances, the cache has become an essential component of computer architecture. Moreover, the rapid digital transformation of daily life has resulted in electronic devices storing greater amounts of sensitive information. Thus, device users are becoming more concerned about the security [...] Read more.
As processor performance advances, the cache has become an essential component of computer architecture. Moreover, the rapid digital transformation of daily life has resulted in electronic devices storing greater amounts of sensitive information. Thus, device users are becoming more concerned about the security of their personal information, so improving processor performance is no longer the sole priority. Hardware vulnerabilities are generally more difficult to detect and address compared to software viruses and related threats. A common technique for exploiting hardware vulnerabilities is through side-channel attacks. They can bypass software security to extract personal information directly from hardware components like the cache or registers. This paper introduces a novel architecture for the arithmetic logic unit (ALU) and associated memory controller (MC) based on the RISC-V microarchitecture to mitigate side-channel attacks. The proposed approach employs hardware-generated random numbers and has minimal design costs, negligible impact on the original system structure, seamless integration, and easy modification of internal components. Results are presented that show it is effective against side-channel attacks. Full article
(This article belongs to the Section Information Theory, Probability and Statistics)
Show Figures

Figure 1

Figure 1
<p>The proposed SAluMC architecture.</p>
Full article ">Figure 2
<p>The SALU architecture.</p>
Full article ">Figure 3
<p>The memory controller architecture.</p>
Full article ">Figure 4
<p>The SAluMC device architecture in the RISC-V microarchitecture of the five-stage pipeline.</p>
Full article ">Figure 5
<p>On-chip power for SAluMC with the five-stage pipeline core.</p>
Full article ">Figure 6
<p>SALU and ALU performance.</p>
Full article ">Figure 7
<p>LFSR random value range with different initial seeds and structures.</p>
Full article ">Figure 8
<p>Set numbers for one-way and four-way 8 KB cache sizes with 5000 and 1,000,000 iterations.</p>
Full article ">Figure 9
<p>Set numbers for one-way and four-way 64 KB cache sizes with 5000 and 1,000,000 iterations.</p>
Full article ">Figure 10
<p>Set number probabilities for one-way 8 KB and 64 KB caches.</p>
Full article ">Figure 10 Cont.
<p>Set number probabilities for one-way 8 KB and 64 KB caches.</p>
Full article ">
16 pages, 7069 KiB  
Article
Tradeoffs Between Richness and Bias of Augmented Data in Long-Tail Recognition
by Wei Dai, Yanbiao Ma, Jiayi Chen, Xiaohua Chen and Shuo Li
Entropy 2025, 27(2), 201; https://doi.org/10.3390/e27020201 - 14 Feb 2025
Viewed by 233
Abstract
In long-tail scenarios, models have a very high demand for high-quality data. Information augmentation, as an important class of data-centric methods, has been proposed to improve model performance by expanding the richness and quantity of samples in tail classes. However, the underlying mechanisms [...] Read more.
In long-tail scenarios, models have a very high demand for high-quality data. Information augmentation, as an important class of data-centric methods, has been proposed to improve model performance by expanding the richness and quantity of samples in tail classes. However, the underlying mechanisms behind the effectiveness of information augmentation methods remain underexplored. This has led to reliance on empirical and intricate fine-tuning in the use of information augmentation for long-tail recognition tasks. In this work, we simultaneously consider the richness gain and distribution shift introduced by information augmentation methods and propose effective information gain (EIG) to explore the mechanisms behind the effectiveness of these methods. We find that when the value of the effective information gain appropriately balances the richness gain and distribution shift, the performance of information augmentation methods is fully realized. Comprehensive experiments on long-tail benchmark datasets CIFAR-10-LT, CIFAR-100-LT, and ImageNet-LT demonstrate that using effective information gain to filter augmented data can further enhance model performance without any modifications to the model’s architecture. Therefore, in addition to proposing new model architectures, data-centric approaches also hold significant potential in the field of long-tail recognition. Full article
(This article belongs to the Section Information Theory, Probability and Statistics)
Show Figures

Figure 1

Figure 1
<p>In the traditional development cycle of AI models, researchers primarily focus on improving the model’s architecture or training techniques to enhance the performance of long-tail recognition. The common research paradigm involves assessing performance differences between different models, given training and test data. Data-centric long-tail learning, on the other hand, should keep the model constant and concentrate on improving the quality of the dataset.</p>
Full article ">Figure 2
<p>(<b>A</b>) The data manifold corresponding to a class has boundaries, and samples outside the manifold boundaries may not exist in the real world, such as dogs with long wings. (<b>B</b>) Grid regions represent the diversity gain in features brought about by augmented samples.</p>
Full article ">Figure 3
<p><b>Remix</b> improves upon Mixup by modifying the synthesized labels so that when samples from head classes and tail classes are mixed, the resulting label is <math display="inline"><semantics> <mrow> <mn>100</mn> <mo>%</mo> </mrow> </semantics></math> contributed by the tail class. <b>CMO</b> uses CutMix to paste patches from tail class samples onto head class samples, thereby generating augmented samples. <b>OFA</b> decomposes sample features into class-specific and class-generic features and combines the tail class’s specific features with the head class’s generic features to generate augmented samples for the tail class. <b>FDC</b> observes that similar classes have similar distribution statistics (variances) and, thus, transfers the variance of the most similar head class to the tail class to re-estimate the distribution, sampling augmented samples from the new distribution.</p>
Full article ">Figure 4
<p>The performances of information augmentation with different FDGs in CIFAR-10-LT.</p>
Full article ">Figure 5
<p>The performances of information augmentation with different FDGs in CIFAR-100-LT.</p>
Full article ">Figure 6
<p>Variation curve of the optimal EIG for augmented data across 10 CIFAR-10-LT datasets with different imbalance factors (IF = <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>∗</mo> <mn>10</mn> <mo>(</mo> <mi>i</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>,</mo> <mo>…</mo> <mo>,</mo> <mn>10</mn> <mo>)</mo> </mrow> </semantics></math>).</p>
Full article ">Figure 7
<p>(<b>Left</b>): Using the EIG to select augmented samples generated by Remix, CMO, OFA, and FDC in ImageNet-LT significantly improves the overall model performance. (<b>Right</b>): Performance improvement in tail classes, using EIG-selected augmented data.</p>
Full article ">
16 pages, 2425 KiB  
Article
Adaptive Trust Evaluation Model Based on Entropy Weight Method for Sensing Terminal Process
by Tao Li and Yanyi Zhang
Entropy 2025, 27(2), 200; https://doi.org/10.3390/e27020200 - 14 Feb 2025
Viewed by 249
Abstract
Remote sensing (RS) has been widely used for data acquisition, monitoring, control, and intelligent decision-making. However, most of them are unattended and easily become the target of attack, which means there are still some risks in the sensing terminal processes. Therefore, trust evaluation [...] Read more.
Remote sensing (RS) has been widely used for data acquisition, monitoring, control, and intelligent decision-making. However, most of them are unattended and easily become the target of attack, which means there are still some risks in the sensing terminal processes. Therefore, trust evaluation of the processes associated with the sensing terminal is necessary. The existing trust evaluation model based on the sensing terminal process has some defects, such as low performance, low precision, and difficulty in effectively identifying malicious processes in the sensing terminal. In this paper, an adaptive trust evaluation model based on the entropy weight method is proposed to detect the sensing terminal process (PB-ATEM). By establishing two kinds of trust values, the direct trust value and the reciprocal trust value, we can comprehensively judge whether the process is trustworthy. For the direct trust value, we can dynamically capture the auto-correlation value of processes and establish a dynamic reward and punishment function to improve the response to malicious processes. For reciprocal trust values, k-means cluster analysis is used to classify processes, and the optimal entropy weight method is proposed to calculate the direct trust weight and reciprocal trust weight of each process more accurately, which accelerates the exposure of malicious processes. Finally, these two trust values are weighted to obtain the integrated trust value. The simulation results show that PB-ATEM can respond quickly to malicious processes. Compared with the existing trust evaluation models, it has higher detection accuracy and better ability to identify malicious processes. Full article
(This article belongs to the Special Issue Information-Theoretic Cryptography and Security)
Show Figures

Figure 1

Figure 1
<p>PB-ATEM Framework.</p>
Full article ">Figure 2
<p>PB-ATEM working flow.</p>
Full article ">Figure 3
<p>Changes in trust values for different categories of processes.</p>
Full article ">Figure 4
<p>(<b>a</b>) PB-ATEM (with reward and punishment functions) vs. the Comparison models; (<b>b</b>) PB-ATEM (without reward and punishment functions) vs. the Comparison models. The data with green color from [<a href="#B12-entropy-27-00200" class="html-bibr">12</a>].</p>
Full article ">Figure 5
<p>Response speed of the three models to malicious processes. The data with green color from [<a href="#B12-entropy-27-00200" class="html-bibr">12</a>].</p>
Full article ">Figure 6
<p>The accuracy of the three models in detecting malicious processes in different proportions of malicious processes. The data with green color from [<a href="#B12-entropy-27-00200" class="html-bibr">12</a>].</p>
Full article ">
12 pages, 1184 KiB  
Article
Three-Phase Confusion Learning
by Filippo Caleca, Simone Tibaldi and Elisa Ercolessi
Entropy 2025, 27(2), 199; https://doi.org/10.3390/e27020199 - 14 Feb 2025
Viewed by 243
Abstract
The use of Neural Networks in quantum many-body theory has undergone a formidable rise in recent years. Among the many possible applications, their pattern recognition power can be utilized when dealing with the study of equilibrium phase diagrams. Learning by Confusion has emerged [...] Read more.
The use of Neural Networks in quantum many-body theory has undergone a formidable rise in recent years. Among the many possible applications, their pattern recognition power can be utilized when dealing with the study of equilibrium phase diagrams. Learning by Confusion has emerged as an interesting and unbiased scheme within this context. This technique involves systematically reassigning labels to the data in various ways, followed by training and testing the Neural Network. While random labeling results in low accuracy, the method reveals a peak in accuracy when the data are correctly and meaningfully partitioned, even if the correct labeling is initially unknown. Here, we propose a generalization of this confusion scheme for systems with more than two phases, for which it was originally proposed. Our construction relies on the use of a slightly different Neural Network: from a binary classifier, we move to a ternary one, which is more suitable to detect systems exhibiting three phases. After introducing this construction, we test it on free and interacting Kitaev chains and on the one-dimensional Extended Hubbard model, consistently achieving results that are compatible with previous works. Our work opens the way to wider use of Learning by Confusion, demonstrating once more the usefulness of Machine Learning to address quantum many-body problems. Full article
(This article belongs to the Section Statistical Physics)
Show Figures

Figure 1

Figure 1
<p><b>Confusion learning</b>. (<b>a</b>) We start by selecting a line of the phase diagram that may or may not cross a phase transition by fixing one parameter and changing the other one (in this example phase diagram, <math display="inline"><semantics> <mi>λ</mi> </semantics></math> is fixed and <math display="inline"><semantics> <mi>μ</mi> </semantics></math> is changed). (<b>b</b>) By sweeping a parameter <math display="inline"><semantics> <msub> <mi>μ</mi> <mi>c</mi> </msub> </semantics></math> in the discretized interval <math display="inline"><semantics> <mrow> <mo>[</mo> <mi>A</mi> <mo>,</mo> <mi>B</mi> <mo>]</mo> </mrow> </semantics></math>, we generate different labeling for our data, going from all zeros to all ones. (<b>c</b>) Scheme of the Convolutional Neural Network used in the process. Blue represents the input data; green, yellow, and purple indicate the intermediate layers; and, finally, the accuracy is read from the red square representing the output neuron. For each labeling, we train a Convolutional Neural Network and plot its accuracy. (<b>d</b>) We expect the canonical <span class="html-italic">V</span>-shape or <span class="html-italic">W</span>-shape in the case of no (<b>top panel</b>) or one (<b>middle panel</b>) phase transition, while the outcome in the presence of three or more phases is unknown, as shown in the (<b>lower panel</b>).</p>
Full article ">Figure 2
<p><b>Two-phase and three-phase learning on the Kitaev model</b>. The free Kitaev model: (<b>a</b>) phase diagram for <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>∈</mo> <mo>[</mo> <mo>−</mo> <mn>8</mn> <mo>,</mo> <mn>8</mn> <mo>]</mo> <mo>,</mo> <mo>Δ</mo> <mo>∈</mo> <mo>[</mo> <mo>−</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>]</mo> </mrow> </semantics></math>, presenting one trivial phase (TRI) and two topological phases (TOP<math display="inline"><semantics> <mrow> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math>/TOP<math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math>) for <math display="inline"><semantics> <mrow> <mo>|</mo> <mi>μ</mi> <mo>|</mo> <mo>≤</mo> <mn>2</mn> <mo>Δ</mo> </mrow> </semantics></math>. In red, at <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, the line chosen to test the models. (<b>b</b>) 2-phase learning applied to Kitaev. (<b>c</b>) 3-phase learning that predicts the two phase transitions at <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>μ</mi> <mi>c</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>μ</mi> <mi>c</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1.92</mn> <mo>,</mo> <mo>−</mo> <mn>1.92</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. Interacting Kitaev model: (<b>d</b>) phase diagram; in red is the section considered for confusion learning at <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. (<b>e</b>) 2-phase learning shows inconclusive results. (<b>f</b>) 3-phase learning shows a peak at two phase transition points, <math display="inline"><semantics> <mrow> <msubsup> <mi>V</mi> <mi>c</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>V</mi> <mi>c</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p><b>Two-phase and three-phase learning applied to Extended Hubbard with</b> <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="bold">r</mi> <mi mathvariant="bold">c</mi> </msub> <mo>=</mo> <mn mathvariant="bold">1</mn> <mo>,</mo> <mn mathvariant="bold">2</mn> </mrow> </semantics></math>. (<b>a</b>) Phase diagram of the <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> model showing the CDW and SDW sectors separated by the thin BOW phase. The black rectangle indicates the points where confusion learning was applied. (<b>b</b>) For this model, 2-phase learning detects a single phase transition at <math display="inline"><semantics> <mrow> <mi>V</mi> <mo>=</mo> <mn>2</mn> <mi>t</mi> </mrow> </semantics></math>, while (<b>c</b>) 3-phase learning shows a peak at the two close phase transition points <math display="inline"><semantics> <mrow> <msubsup> <mi>V</mi> <mi>c</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> <mo>≃</mo> <mn>1.9</mn> <mi>t</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mi>V</mi> <mi>c</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>≃</mo> <mn>2.1</mn> <mi>t</mi> </mrow> </semantics></math>. (<b>d</b>) Phase diagram for high <span class="html-italic">U</span> values of the <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> model with three phases; the region investigated with confusion learning is highlighted by the red rectangle. In this case, (<b>e</b>) 2-phase learning returns a plateau of high accuracy for all the values inside the <math display="inline"><semantics> <mrow> <mi>C</mi> <mi>L</mi> <msub> <mi>L</mi> <mrow> <mi>n</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math> phase, while (<b>f</b>) 3-phase learning shows a clear peak in accuracy at coordinates <math display="inline"><semantics> <mrow> <msubsup> <mi>V</mi> <mi>c</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> <mo>≃</mo> <mn>3.8</mn> <mi>t</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mi>V</mi> <mi>c</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>≃</mo> <mn>6.8</mn> <mi>t</mi> </mrow> </semantics></math>.</p>
Full article ">
30 pages, 4693 KiB  
Article
A Perturbative Approach to the Solution of the Thirring Quantum Cellular Automaton
by Alessandro Bisio, Paolo Perinotti, Andrea Pizzamiglio and Saverio Rota
Entropy 2025, 27(2), 198; https://doi.org/10.3390/e27020198 - 13 Feb 2025
Viewed by 404
Abstract
The Thirring Quantum Cellular Automaton (QCA) describes the discrete time dynamics of local fermionic modes that evolve according to one step of the Dirac cellular automaton, followed by the most general on-site number-preserving interaction, and serves as the QCA counterpart of the Thirring [...] Read more.
The Thirring Quantum Cellular Automaton (QCA) describes the discrete time dynamics of local fermionic modes that evolve according to one step of the Dirac cellular automaton, followed by the most general on-site number-preserving interaction, and serves as the QCA counterpart of the Thirring model in quantum field theory. In this work, we develop perturbative techniques for the QCA path sum approach, expanding both the number of interaction vertices and the mass parameter of the Thirring QCA. By classifying paths within the regimes of very light and very heavy particles, we computed the transition amplitudes in the two- and three-particle sectors to the first few orders. Our investigation into the properties of the Thirring QCA, addressing the combinatorial complexity of the problem, yielded some useful results applicable to the many-particle sector of any on-site number-preserving interactions in one spatial dimension. Full article
(This article belongs to the Special Issue Recent Advances and Challenges in Quantum Cellular Automata)
Show Figures

Figure 1

Figure 1
<p>Example of paths in the past causal cone of site <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>. The edges of the cone represent light-like paths. Space and time are expressed in arbitrary units.</p>
Full article ">Figure 2
<p>(<b>a</b>) Correspondence between different ways of representing the evolution of one particle at each step of the automaton. Each site accommodates two internal degrees of freedom: 0, or right mode (blue top half); and 1, or left mode (red bottom half). (<b>b</b>) Instance of the three representations for a path. The strings, whether representing lattice transitions or internal degrees of freedom, are ordered with time progressing from left to right, while on the lattice, time flows from bottom to top.</p>
Full article ">Figure 3
<p>(<b>a</b>) Representation of the evolution of the internal state of a single particle at each time step upon choosing its initial state. (<b>b</b>) Representation of the evolution of the internal states of two particles.</p>
Full article ">Figure 4
<p>We separate the total <span class="html-italic">T</span>-step process in a free evolution, given by the matrices <math display="inline"><semantics> <msub> <mi>W</mi> <mn>2</mn> </msub> </semantics></math> (lines), and an interacting one, described by the terms <math display="inline"><semantics> <msup> <mi mathvariant="script">U</mi> <mrow> <mo>(</mo> <mi>k</mi> <mo>)</mo> </mrow> </msup> </semantics></math>. Each <math display="inline"><semantics> <msup> <mi mathvariant="script">U</mi> <mrow> <mo>(</mo> <mi>k</mi> <mo>)</mo> </mrow> </msup> </semantics></math> contains <span class="html-italic">k</span> interactions (circles).</p>
Full article ">Figure 5
<p>Algorithm for the computation of <math display="inline"><semantics> <msup> <mi mathvariant="script">U</mi> <mrow> <mo>(</mo> <mi>k</mi> <mo>)</mo> </mrow> </msup> </semantics></math> for the case <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> at fixed initial and final particle states: (<b>a</b>) The region <span class="html-italic">D</span> where interactions can occur; (<b>b</b>) example of choice of two time steps wherein the interactions occur; (<b>c</b>) sum of all possible causally connected interaction sites (green dots and darker background) for the given choice of time steps; (<b>d</b>) sum of all possible ways of choosing time steps within region <span class="html-italic">D</span>.</p>
Full article ">Figure 6
<p>Interacting diagrams contributing to the second order in vertices expansion are highlighted in blue, while interacting diagrams contributing to the second order in mass expansion are highlighted in red.</p>
Full article ">Figure 7
<p><math display="inline"><semantics> <msub> <mi mathvariant="script">R</mi> <mi>v</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi mathvariant="script">R</mi> <mi>h</mi> </msub> </semantics></math> represent on binary matrices <math display="inline"><semantics> <msub> <mi>W</mi> <mrow> <mi>i</mi> <mi>j</mi> </mrow> </msub> </semantics></math> (lines) the reflections about the vertical <span class="html-italic">v</span> and horizontal <span class="html-italic">h</span> axes of the plane. Big circles represent the full process. Panel (<b>a</b>,<b>b</b>) shows the case where the relative position of the particles does (does not) change sign.</p>
Full article ">Figure 8
<p>Diagrams belonging to the class <math display="inline"><semantics> <mrow> <mo>(</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>)</mo> </mrow> </semantics></math> for some <span class="html-italic">T</span>-step process with <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. Dashed lines represent alternative paths connecting the same extremal points; notice that not all of them yield interactions. Equation (<a href="#FD19-entropy-27-00198" class="html-disp-formula">19</a>) accounts for this occurrence by excluding such paths from contributing to the interaction amplitude.</p>
Full article ">Figure 9
<p>All diagrams contributing to the low-mass regime process depicted in Figure up to the third perturbative order <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. The first diagram belongs to the class <math display="inline"><semantics> <msup> <mi mathvariant="script">F</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msup> </semantics></math>, while highlighted in turquoise are the 12 diagrams belonging to the class <math display="inline"><semantics> <msup> <mi mathvariant="script">F</mi> <mrow> <mo>(</mo> <mn>3</mn> <mo>)</mo> </mrow> </msup> </semantics></math> (dashed lines represent alternative paths connecting the same extremal points).</p>
Full article ">Figure 10
<p>Example of diagrams belonging to the class <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>−</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math> for an even number <span class="html-italic">T</span> of time steps. Depending on the border conditions, there can be an even (<b>a</b>) or odd (<b>b</b>) number of interactions. Green lines denote the tracks of paths both particles share, resulting in consecutive interactions. Dashed lines represent alternative paths connecting the same extremal points.</p>
Full article ">Figure 11
<p>Example of diagrams belonging to the classes (<b>a</b>) <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>−</mo> <mn>2</mn> <mo>)</mo> </mrow> </semantics></math> and (<b>b</b>) <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>T</mi> <mo>−</mo> <mn>1</mn> <mo>,</mo> <mi>T</mi> <mo>−</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>Paths violating Pauli’s exclusion principle in the multi-particle sector. Violation is highlighted in red in the two strings corresponding to the (non-interacting) paths.</p>
Full article ">
21 pages, 5719 KiB  
Article
Exergy Analysis of a Convective Heat Pump Dryer Integrated with a Membrane Energy Recovery Ventilator
by Anand Balaraman, Md Ashiqur Rahman, Davide Ziviani and David M. Warsinger
Entropy 2025, 27(2), 197; https://doi.org/10.3390/e27020197 - 13 Feb 2025
Viewed by 274
Abstract
To increase energy efficiency, heat pump dryers and membrane dryers have been proposed to replace conventional fossil fuel dryers. Both conventional and heat pump dryers require substantial energy for condensing and reheating, while “active” membrane systems require vacuum pumps that are insufficiently developed. [...] Read more.
To increase energy efficiency, heat pump dryers and membrane dryers have been proposed to replace conventional fossil fuel dryers. Both conventional and heat pump dryers require substantial energy for condensing and reheating, while “active” membrane systems require vacuum pumps that are insufficiently developed. Lower temperature dehumidification systems make efficient use of membrane energy recovery ventilators (MERVs) that do not need vacuum pumps, but their high heat losses and lack of vapor selectivity have prevented their use in industrial drying. In this work, we propose an insulating membrane energy recovery ventilator for moisture removal from drying exhaust air, thereby reducing sensible heat loss from the dehumidification process and reheating energy. The second law analysis of the proposed system is carried out and compared with a baseline convective heat pump dryer. Irreversibilities in each component under different ambient temperatures (5–35 °C) and relative humidity (5–95%) are identified. At an ambient temperature of 35 °C, the proposed system substantially reduces sensible heat loss (47–60%) in the dehumidification process, resulting in a large reduction in condenser load (45–50%) compared to the baseline system. The evaporator in the proposed system accounts for up to 59% less irreversibility than the baseline system. A maximum of 24.5% reduction in overall exergy input is also observed. The highest exergy efficiency of 10.2% is obtained at an ambient condition of 35 °C and 5% relative humidity, which is more than twice the efficiency of the baseline system under the same operating condition. Full article
(This article belongs to the Special Issue Thermodynamic Optimization of Energy Systems)
Show Figures

Figure 1

Figure 1
<p>Schematic of (<b>a</b>) the baseline heat pump dryer, and (<b>b</b>) the proposed dryer with a membrane energy recovery ventilator. The baseline uses a heat pump to condense water vapor, where its hot condenser coil is placed to reheat the air stream, and then rejects excess heat to the ambient air through an auxiliary condenser. The proposed dryer removes water vapor with a membrane that exchanges vapor with ambient air, reducing heat loss and reheating energy. The heat pump that reheats the dehumidified air stream draws its heat from the humidified ambient air stream, thereby reducing the temperature rise.</p>
Full article ">Figure 2
<p>Representation of the baseline and proposed system in the T-s diagram (<b>a</b>) air-side T-s diagram, and (<b>b</b>) refrigerant-side T-s diagram.</p>
Full article ">Figure 3
<p>Exergy destruction in the evaporator of the heat pump in the baseline (<b>left</b>) and present system (<b>right</b>) as ambient temperature and relative humidity are varied. The membrane’s sensible and latent effectiveness (0.7 and 0.7, respectively), dryer air inlet conditions (70 °C and 10% RH), dryer capacity (7 kg), dryer air velocity (1 m/s), superheat (5 °C) and subcooling (5 °C) in the heat pump are kept constant as specified in the modeling assumptions in <a href="#sec2dot2-entropy-27-00197" class="html-sec">Section 2.2</a>.</p>
Full article ">Figure 4
<p>Exergy destruction in the condenser of the heat pump in the baseline and present systems at various ambient temperature and relative humidity. The other variables are kept constant as mentioned in <a href="#entropy-27-00197-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>Exergy destruction in the compressor of heat pump in the baseline and present system at different ambient temperatures and relative humidities. The other variables are kept constant as mentioned in <a href="#entropy-27-00197-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 6
<p>Exergy destruction in the membrane dehumidifier in the present system at various ambient temperatures and relative humidities. The other variables are kept constant as mentioned in <a href="#entropy-27-00197-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 7
<p>(<b>a</b>) Exergy efficiency of the baseline and present system at various ambient temperatures and relative humidity levels. The other variables are kept constant, as mentioned in <a href="#entropy-27-00197-f003" class="html-fig">Figure 3</a>. (<b>b</b>) The exergy efficiency improvement of the proposed system over the baseline system.</p>
Full article ">Figure 8
<p>The influence of the membrane’s sensible effectiveness on the overall exergy destruction of the present system at different ambient temperatures (<b>left</b>). The influence of the membrane’s sensible effectiveness on the second law efficiency of the present system at different ambient temperatures (<b>right</b>). In both cases, ambient relative humidity and membrane latent effectiveness were kept constant at 5% and 0.7, respectively. The non-contour region represents that the present system cannot remove the desired moisture content and it can be considered the non-operating zone.</p>
Full article ">Figure 9
<p>Grassman diagram of the (<b>a</b>) baseline and (<b>b</b>) present system at 35 °C ambient temperature and 5% relative humidity. The other variables are kept constant, as mentioned in <a href="#entropy-27-00197-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 9 Cont.
<p>Grassman diagram of the (<b>a</b>) baseline and (<b>b</b>) present system at 35 °C ambient temperature and 5% relative humidity. The other variables are kept constant, as mentioned in <a href="#entropy-27-00197-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 10
<p>Exergy destruction in the major components of the natural gas-fired dryer, baseline heat pump dryer and the present systems at different drying temperatures (60 °C, 70 °C and 80 °C), with an ambient condition of 35 °C and 5% relative humidity. The other variables for the baseline and proposed system are kept constant, as mentioned in <a href="#entropy-27-00197-f003" class="html-fig">Figure 3</a>.</p>
Full article ">
20 pages, 932 KiB  
Article
Gradient-Based Multiple Robust Learning Calibration on Data Missing-Not-at-Random via Bi-Level Optimization
by Shuxia Gong and Chen Ma
Entropy 2025, 27(2), 196; https://doi.org/10.3390/e27020196 - 13 Feb 2025
Viewed by 314
Abstract
Recommendation systems (RS) have become integral to numerous digital platforms and applications, ranging from e-commerce to content streaming field. A critical problem in RS is that the ratings are missing not at random (MNAR), which is due to the users always giving feedback [...] Read more.
Recommendation systems (RS) have become integral to numerous digital platforms and applications, ranging from e-commerce to content streaming field. A critical problem in RS is that the ratings are missing not at random (MNAR), which is due to the users always giving feedback on items with self-selection. The biased selection of rating data results in inaccurate rating prediction for all user-item pairs. Doubly robust (DR) learning has been studied in many tasks in RS, which is unbiased when either a single imputation or a single propensity model is accurate. In addition, multiple robust (MR) has been proposed with multiple imputation models and propensity models, and is unbiased when there exists a linear combination of these imputation models and propensity models is correct. However, we claim that the imputed errors and propensity scores are miscalibrated in the MR method. In this paper, we propose a gradient-based calibrated multiple robust learning method to enhance the debiasing performance and reliability of the rating prediction model. Specifically, we propose to use bi-level optimization to solve the weights and model coefficients of each propensity and imputation model in MR framework. Moreover, we adopt the differentiable expected calibration error as part of the objective to optimize the model calibration quality directly. Experiments on three real-world datasets show that our method outperforms the state-of-the-art baselines. Full article
(This article belongs to the Special Issue Causal Inference in Recommender Systems)
Show Figures

Figure 1

Figure 1
<p>Categorization of model calibration methods: Post-hoc Calibration [<a href="#B46-entropy-27-00196" class="html-bibr">46</a>,<a href="#B47-entropy-27-00196" class="html-bibr">47</a>,<a href="#B48-entropy-27-00196" class="html-bibr">48</a>], Regularization [<a href="#B36-entropy-27-00196" class="html-bibr">36</a>,<a href="#B49-entropy-27-00196" class="html-bibr">49</a>,<a href="#B50-entropy-27-00196" class="html-bibr">50</a>,<a href="#B51-entropy-27-00196" class="html-bibr">51</a>,<a href="#B52-entropy-27-00196" class="html-bibr">52</a>], Uncertainty Estimation [<a href="#B53-entropy-27-00196" class="html-bibr">53</a>,<a href="#B54-entropy-27-00196" class="html-bibr">54</a>,<a href="#B55-entropy-27-00196" class="html-bibr">55</a>,<a href="#B56-entropy-27-00196" class="html-bibr">56</a>], and Hybrid Calibration [<a href="#B57-entropy-27-00196" class="html-bibr">57</a>,<a href="#B58-entropy-27-00196" class="html-bibr">58</a>].</p>
Full article ">Figure 2
<p>Comparison of joint calibration and individual calibration on the <b>Yahoo! R3</b> dataset, with different numbers of candidate propensity and imputation models.</p>
Full article ">Figure 3
<p>Comparison of joint calibration and individual calibration on the <b>KuaiRec</b> dataset, with different numbers of candidate propensity and imputation models.</p>
Full article ">Figure 4
<p>Impact of model calibration hyper-parameter <math display="inline"><semantics> <msub> <mi>λ</mi> <mrow> <mi>p</mi> <mi>r</mi> <mi>o</mi> <mi>p</mi> </mrow> </msub> </semantics></math> in multiple propensity calibration and <math display="inline"><semantics> <msub> <mi>λ</mi> <mrow> <mi>i</mi> <mi>m</mi> <mi>p</mi> </mrow> </msub> </semantics></math> in multiple imputation calibration on <b>Coat</b> dataset.</p>
Full article ">Figure 5
<p>Impact of model calibration hyper-parameter <math display="inline"><semantics> <msub> <mi>λ</mi> <mrow> <mi>p</mi> <mi>r</mi> <mi>o</mi> <mi>p</mi> </mrow> </msub> </semantics></math> in multiple propensity calibration and <math display="inline"><semantics> <msub> <mi>λ</mi> <mrow> <mi>i</mi> <mi>m</mi> <mi>p</mi> </mrow> </msub> </semantics></math> in multiple imputation calibration on <b>KuaiRec</b> dataset.</p>
Full article ">Figure 6
<p>Effect of varying <span class="html-italic">M</span> in soft binning strategy on prediction performance on <b>Coat</b> dataset.</p>
Full article ">Figure 7
<p>Effect of varying <span class="html-italic">M</span> in soft binning strategy on prediction performance on <b>KuaiRec</b> dataset.</p>
Full article ">
10 pages, 448 KiB  
Article
Revisiting Endoreversible Carnot Engine: Extending the Yvon Engine
by Xiu-Hua Zhao and Yu-Han Ma
Entropy 2025, 27(2), 195; https://doi.org/10.3390/e27020195 - 13 Feb 2025
Viewed by 441
Abstract
Curzon and Ahlborn’s 1975 paper, a pioneering work that inspired the birth of the field of finite-time thermodynamics, unveiled the efficiency at maximum power (EMP) of the endoreversible Carnot heat engine, now commonly referred to as the Curzon–Ahlborn (CA) engine. Historically, despite the [...] Read more.
Curzon and Ahlborn’s 1975 paper, a pioneering work that inspired the birth of the field of finite-time thermodynamics, unveiled the efficiency at maximum power (EMP) of the endoreversible Carnot heat engine, now commonly referred to as the Curzon–Ahlborn (CA) engine. Historically, despite the significance of the CA engine, similar findings had emerged at an earlier time, such as the Yvon engine proposed by J. Yvon in 1955 that shares the exact same EMP, that is, the CA efficiency ηCA. However, the special setup of the Yvon engine has circumscribed its broader influence. This paper extends the Yvon engine model to achieve a level of generality comparable to that of the CA engine. With the power expression of the extended Yvon engine, we directly explain the universality that ηCA is independent of the heat transfer coefficients between the working substance and the heat reservoirs. A rigorous comparison reveals that the extended Yvon engine and CA engine represent the steady-state and cyclic forms of the endoreversible Carnot heat engine, respectively, and are equivalent. Full article
(This article belongs to the Special Issue The First Half Century of Finite-Time Thermodynamics)
Show Figures

Figure 1

Figure 1
<p>Steady-state (<b>a</b>,<b>b</b>) and cyclic (<b>c</b>,<b>d</b>) endoreversible heat engines. (<b>a</b>) In the original Yvon engine, finite heat flux (denoted as <math display="inline"><semantics> <mover accent="true"> <mi>Q</mi> <mo>˙</mo> </mover> </semantics></math>) occurs only at the high-temperature end, where there is a temperature difference between the working substance and the hot reservoir. (<b>b</b>) The extended Yvon engine introduces temperature differences, and thus, heat fluxes, between the working substance and both the hot and cold reservoirs. (<b>c</b>,<b>d</b>) show the entropy (<span class="html-italic">S</span>)–temperature (<span class="html-italic">T</span>) diagrams of the endoreversible Carnot engine cycles with finite heat fluxes (denoted as <math display="inline"><semantics> <mover accent="true"> <mi>q</mi> <mo>˙</mo> </mover> </semantics></math>) along the high-temperature and both isothermal branches, respectively.</p>
Full article ">Figure 2
<p>Trade-off relations between power and efficiency for the Curzon-Ahlborn engine with <math display="inline"><semantics> <mrow> <msub> <mi>η</mi> <mi mathvariant="normal">C</mi> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> (dash-dotted curve), <math display="inline"><semantics> <mrow> <msub> <mi>η</mi> <mi mathvariant="normal">C</mi> </msub> <mo>=</mo> <mn>0.7</mn> </mrow> </semantics></math> (dashed curve), and <math display="inline"><semantics> <mrow> <msub> <mi>η</mi> <mi mathvariant="normal">C</mi> </msub> <mo>=</mo> <mn>0.9</mn> </mrow> </semantics></math> (solid curve). The triangles and circle mark the maximum power and maximum efficiency, respectively.</p>
Full article ">
11 pages, 546 KiB  
Technical Note
The Density Profile of a Neutron Star
by Allan D. Woodbury
Entropy 2025, 27(2), 194; https://doi.org/10.3390/e27020194 - 13 Feb 2025
Viewed by 221
Abstract
The problem posed in this study is to determine the density distribution within an ideal spherically symmetric neutron star based on only two constraints: the volumetrically averaged density and a moment of inertia factor, f. In order to deal with the above, [...] Read more.
The problem posed in this study is to determine the density distribution within an ideal spherically symmetric neutron star based on only two constraints: the volumetrically averaged density and a moment of inertia factor, f. In order to deal with the above, it is recognized that space within these objects is heavily curved, and thus lengths, densities, and the moment of inertia have to be adjusted for relativistic effects. For the first time, the minimum relative entropy methodology (MRE) is used to find the expected value of a series of effective densities within a neutron star. In numerical experiments, we use the data from the star PSR J0737-3039A, which has a mass of 2.6×1030 kg and a radius of 13.75 km. Here, the factor f is based on a range of values of moments of inertia (MOI): 1.30–1.63 ×1045 g cm2. For f=0.324, at no time do densities cross over 1×1015 gm/cc. For the most part, densities > 6×1014 gm/cc are shown at radial dimensions of less than about 4 km. When f=0.258, densities closer to the core are pushed higher, as one might expect, and peak at slightly over 4×1015 gm/cc. If recent values of MOI are more appropriate at 1.15×1045 g cm2, this then suggests core densities greater than 4×1015 gm/cc. These various density models lead to quantitative statements about qualitative interpretations, and as time goes on, any internal density models should satisfy the two constraints posed. Also, since the model presented here is probabilistic, it can be established that density at a certain depth is constrained within a certain confidence limit. The expected values of densities for PSR J0737-3039A are in reasonable agreement with current conceptual neutron star models but are highly sensitive to assumed MOI values. It is emphasized that the probabilities and the mean values of density obtained are conditional on the imposed moments, namely, M and f, and also the radius R. Full article
(This article belongs to the Section Multidisciplinary Applications)
Show Figures

Figure 1

Figure 1
<p>Conceptual model of a neutron star and its interior. Credit NASA/B. Link, <a href="https://heasarc.gsfc.nasa.gov" target="_blank">https://heasarc.gsfc.nasa.gov</a>.</p>
Full article ">Figure 2
<p>Star J0737-3039A. Expected values of density for different values of moment of inertia (MOI) factors, <span class="html-italic">f</span>. The dashed line is the expected value of prior probability, which is a uniform prior <math display="inline"><semantics> <mrow> <mn>2.4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>14</mn> </msup> </mrow> </semantics></math> gm/cc in the radial direction.</p>
Full article ">Figure 3
<p>SR J0737-3039A. Expected values of density for two different values of moment of inertia factor <span class="html-italic">f</span>. The dashed line is the expected value of prior probability, which, in this case, is a zoned conceptual model loosely based on the conceptual model of <a href="#entropy-27-00194-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 4
<p>Assumed conceptual model of a neutron star studied by [<a href="#B20-entropy-27-00194" class="html-bibr">20</a>]. The expected value of density for assumed values of moment of inertia factor <span class="html-italic">f</span>. The dashed line is the expected value of prior probability from nuclear model V18. The dashed line is a uniform prior in the radial direction.</p>
Full article ">
17 pages, 5697 KiB  
Article
Alkali Halide Aqueous Solutions Under Pressure: A Non-Equilibrium Molecular Dynamics Investigation of Thermal Transport and Thermodiffusion
by Guansen Zhao and Fernando Bresme
Entropy 2025, 27(2), 193; https://doi.org/10.3390/e27020193 - 13 Feb 2025
Viewed by 338
Abstract
Thermal gradients induce thermodiffusion in aqueous solutions, a non-equilibrium effect arising from the coupling of thermal and mass fluxes. While thermal transport processes have garnered significant attention under standard conditions, thermal transport at high pressures and temperatures, typical of the Earth’s crust, has [...] Read more.
Thermal gradients induce thermodiffusion in aqueous solutions, a non-equilibrium effect arising from the coupling of thermal and mass fluxes. While thermal transport processes have garnered significant attention under standard conditions, thermal transport at high pressures and temperatures, typical of the Earth’s crust, has escaped scrutiny. Non-equilibrium thermodynamics theory and non-equilibrium molecular dynamics simulations provide an excellent means to quantify thermal transport under extreme conditions and establish a connection between the behaviour of the solutions and their microscopic structure. Here, we investigate the thermal conductivity and thermal diffusion of NaCl and LiCl solutions in the GPa pressure regime, targeting temperatures between 300 K and 1000 K at 1 molal concentration. We employ non-equilibrium molecular dynamics simulations along with the Madrid-2019 and TIP4P/2005 force fields. The thermal conductivity of the solutions increases significantly with pressure, and following the behaviour observed at standard pressure, the thermal conductivity is lower than that of pure water. The reduction in thermal conductivity is significant in the GPa pressure regime, ∼3% for 1 molal NaCl and LiCl solutions. We demonstrate that under GPa pressure conditions, the solutions feature thermophobic behaviour, with ions migrating towards colder regions. The pronounced impact of pressure is more evident in LiCl solutions, which display a thermophilic to thermophobic “transition” at pressures above 0.25 GPa. We discuss a correlation between the solution’s thermophobicity and the disruption of the water hydrogen bond structure at high pressure, where the water structure resembles that observed in simple liquids. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>Top</b>) Snapshot of the simulations cell employed in the NEMD simulation of 1 molal LiCl aqueous solution. The red, white, blue and green spheres represent the oxygen, hydrogen atoms, and <math display="inline"><semantics> <msup> <mi>Li</mi> <mo>+</mo> </msup> </semantics></math> and <math display="inline"><semantics> <msup> <mi>Cl</mi> <mo>−</mo> </msup> </semantics></math> ions, respectively. The thermostatting regions are highlighted in red (hot, center of the box) and blue (cold, box edges). (<b>Bottom</b>) Temperature profile obtained with our NEMD simulation method.</p>
Full article ">Figure 2
<p>Energy exchanged at the hot and cold thermostats during a representative NEMD run. The blue and orange lines indicate the energy exchange in the cold and hot regions. The negative of the energy exchanged at the hot thermostats has been plotted in green for an easier comparison and test of energy conservation.</p>
Full article ">Figure 3
<p>(<b>a</b>) Thermal conductivities and (<b>b</b>) Soret coefficients at 290 K temperature, 10<sup>−4</sup>, 0.1, 0.5, 1, 1.5, and 2 GPa pressures for 1 m NaCl and LiCl solutions. The square represents the thermal conductivity at 298 K and 31 bar reported in Ref. [<a href="#B21-entropy-27-00193" class="html-bibr">21</a>]. The green points in (<b>a</b>) represent our thermal conductivity data for pure water at 290 K temperature. The red triangles represent the thermal conductivity and Soret coefficient of 1 m LiCl solutions at 1000 K and 2.6 GPa pressure. The numerical data are compiled in <a href="#entropy-27-00193-t0A2" class="html-table">Table A2</a>.</p>
Full article ">Figure 4
<p>Oxygen-oxygen radial distribution functions of pure water at 290 K and various pressures (10<sup>−4</sup>, 0.1, 0.5, 1, 1.5, and 2 GPa).</p>
Full article ">Figure 5
<p>Spatial distribution functions <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mrow> <mi>O</mi> <mi>O</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mo>Ω</mo> <mo>)</mo> </mrow> </mrow> </semantics></math> for the 1 m NaCl (<b>a</b>) and LiCl (<b>b</b>) solutions at 290 K and <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </semantics></math>, 0.1, 0.5, 1, 1.5, and 2 GPa pressures. The central oxygen, colored in red, is shown for reference. Image generated with OVITO [<a href="#B58-entropy-27-00193" class="html-bibr">58</a>].</p>
Full article ">Figure 6
<p>Perturbation of the oxygen-oxygen radial distribution function in 1 m NaCl and LiCl solutions at 290 K and pressures of 10<sup>−4</sup>, 0.1, 0.5, 1, 1.5, and 2 GPa. The different RDFs are presented using the O-O RDFs of water at standard pressure, i.e., <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>g</mi> <mrow> <mi>O</mi> <mi>O</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>g</mi> <mrow> <mi>O</mi> <mi>O</mi> <mo>,</mo> <mi>s</mi> <mi>o</mi> <mi>l</mi> <mi>u</mi> <mi>t</mi> <mi>i</mi> <mi>o</mi> <mi>n</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> <mo>−</mo> <msub> <mi>g</mi> <mrow> <mi>O</mi> <mi>O</mi> <mo>,</mo> <mi>p</mi> <mi>u</mi> <mi>r</mi> <mi>e</mi> <mi>w</mi> <mi>a</mi> <mi>t</mi> <mi>e</mi> <mi>r</mi> <mspace width="3.33333pt"/> <mo>(</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> <mspace width="3.33333pt"/> <mi>G</mi> <mi>P</mi> <mi>a</mi> <mo>)</mo> </mrow> </msub> <mrow> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Oxygen-ion radial distribution functions for 1 m NaCl (<b>a</b>) and LiCl (<b>b</b>) solutions at 290 K and 10<sup>−4</sup>, 0.1, 0.5, 1, 1.5, and 2 GPa pressures. The left and right panels show the oxygen-cation and oxygen-anion RDFs, respectively.</p>
Full article ">Figure 8
<p>(<b>a</b>) Isobaric thermal expansion coefficient, (<b>b</b>) isothermal compressibility, and (<b>c</b>) speed of sound of pure water and 1 m NaCl, and 1 m LiCl solutions at 290 K. The data correspond to the following pressures: 10<sup>−4</sup>, 0.1, 0.5, 1, 1.5, and 2 GPa. The purple triangles represent the experimental data from Ref. [<a href="#B60-entropy-27-00193" class="html-bibr">60</a>] at 10<sup>−4</sup>, 0.1, 0.5, and 0.8 GPa pressures and 290 K.</p>
Full article ">Figure A1
<p>Equation of state from equilibrium NPT simulations (green points) at 10<sup>−4</sup>, 0.1, 0.5, 1, 1.5, and 2 GPa pressures and 290 K temperature. The simulation results are compared with experimental data (purple curve) from 10<sup>−4</sup> to 0.8 GPa [<a href="#B60-entropy-27-00193" class="html-bibr">60</a>].</p>
Full article ">Figure A2
<p>Oxygen-oxygen radial distribution functions for pure water, 1 m NaCl, and LiCl solutions at 290 K and various pressures (10<sup>−4</sup>, 0.1, 0.5, 1, and 2 GPa).</p>
Full article ">Figure A3
<p>Sodium-chloride (blue curve) and lithium-chloride (orange curve) radial distribution functions and corresponding running coordination numbers (dashed line) for NaCl and LiCl solutions at 290 K temperature and 2 GPa pressure. The inset represents the zoomed-in region for the radial distance between 0.2 and 0.35 nm, providing a closer view of the first minimum in the radial distribution functions for Li-Cl.</p>
Full article ">
20 pages, 8250 KiB  
Article
Fault Diagnosis of Wind Turbine Gearbox Based on Improved Multivariate Variational Mode Decomposition and Ensemble Refined Composite Multivariate Multiscale Dispersion Entropy
by Xin Xia, Xiaolu Wang and Weilin Chen
Entropy 2025, 27(2), 192; https://doi.org/10.3390/e27020192 - 13 Feb 2025
Viewed by 266
Abstract
Wind turbine planetary gearboxes have complex structures and operating environments, which makes it difficult to extract fault features effectively. In addition, it is difficult to achieve efficient fault diagnosis. To improve the efficiency of feature extraction and fault diagnosis, a fault diagnosis method [...] Read more.
Wind turbine planetary gearboxes have complex structures and operating environments, which makes it difficult to extract fault features effectively. In addition, it is difficult to achieve efficient fault diagnosis. To improve the efficiency of feature extraction and fault diagnosis, a fault diagnosis method based on improved multivariate variational mode decomposition (IMVMD) and ensemble refined composite multivariate multiscale dispersion entropy (ERCmvMDE) with multi-channel vibration data is proposed. Firstly, the IMVMD is proposed to obtain the optimal parameters of the MVMD, which would make the MVMD more effective. Secondly, the ERCmvMDE is proposed to extract rich and effective feature information. Finally, the fault diagnosis of the planetary gearbox is achieved using the least squares support vector machine (LSSVM) with features consisting of ERCmvMDE. Simulations and experimental studies indicate that the proposed method performs feature extraction well and obtains higher fault diagnosis accuracy. Full article
Show Figures

Figure 1

Figure 1
<p>The flowchart of the proposed fault diagnosis method.</p>
Full article ">Figure 2
<p>The time-domain waveforms of a multi-channel signal.</p>
Full article ">Figure 3
<p>The Sy values for different <span class="html-italic">K</span> (<span class="html-italic">α</span> = 2000).</p>
Full article ">Figure 4
<p>The Sy values for different <span class="html-italic">α</span> (<span class="html-italic">K</span> = 3).</p>
Full article ">Figure 5
<p>The final decomposition effects of IMVMD. (<b>a</b>) <span class="html-italic">x</span><sub>1</sub>, (<b>b</b>) <span class="html-italic">x</span><sub>2</sub>.</p>
Full article ">Figure 6
<p>The mean values and SDs of the results obtained by different methods with 20 groups of data. (<b>a</b>) RCmvMSE; (<b>b</b>) RCmvMFE; (<b>c</b>) RCmvMDE (mean); (<b>d</b>) RCmvMDE (maximum); (<b>e</b>) RCmvMDE (variance); (<b>f</b>) RCmvMDE (RMSA).</p>
Full article ">Figure 6 Cont.
<p>The mean values and SDs of the results obtained by different methods with 20 groups of data. (<b>a</b>) RCmvMSE; (<b>b</b>) RCmvMFE; (<b>c</b>) RCmvMDE (mean); (<b>d</b>) RCmvMDE (maximum); (<b>e</b>) RCmvMDE (variance); (<b>f</b>) RCmvMDE (RMSA).</p>
Full article ">Figure 7
<p>The time-domain waveforms of the vibration data of the planetary gearbox. (<b>A</b>) X direction; (<b>B</b>) Y direction.</p>
Full article ">Figure 8
<p>The mean values and SDs of the results obtained by different methods with 200 groups of sampling data. (<b>a</b>) RCmvMSE; (<b>b</b>) RCmvMFE; (<b>c</b>) RCmvMDE (mean); (<b>d</b>) RCmvMDE (maximum); (<b>e</b>) RCmvMDE (variance); (<b>f</b>) RCmvMDE (RMSA).</p>
Full article ">Figure 8 Cont.
<p>The mean values and SDs of the results obtained by different methods with 200 groups of sampling data. (<b>a</b>) RCmvMSE; (<b>b</b>) RCmvMFE; (<b>c</b>) RCmvMDE (mean); (<b>d</b>) RCmvMDE (maximum); (<b>e</b>) RCmvMDE (variance); (<b>f</b>) RCmvMDE (RMSA).</p>
Full article ">Figure 8 Cont.
<p>The mean values and SDs of the results obtained by different methods with 200 groups of sampling data. (<b>a</b>) RCmvMSE; (<b>b</b>) RCmvMFE; (<b>c</b>) RCmvMDE (mean); (<b>d</b>) RCmvMDE (maximum); (<b>e</b>) RCmvMDE (variance); (<b>f</b>) RCmvMDE (RMSA).</p>
Full article ">Figure 9
<p>The t-SEN visualizations of different feature extraction methods. (<b>a</b>) IMVMD+ RCmvMSE, (<b>b</b>) IMVMD + RCmvMFE, (<b>c</b>) IMVMD + RCmvMDE, (<b>d</b>) IMVMD + ERCmvMDE.</p>
Full article ">Figure 9 Cont.
<p>The t-SEN visualizations of different feature extraction methods. (<b>a</b>) IMVMD+ RCmvMSE, (<b>b</b>) IMVMD + RCmvMFE, (<b>c</b>) IMVMD + RCmvMDE, (<b>d</b>) IMVMD + ERCmvMDE.</p>
Full article ">Figure 10
<p>The confusion diagram of fault diagnosis with different parameters. (<b>a</b>) With default parameters (<span class="html-italic">K</span> = 5, <span class="html-italic">α</span> = 2000); (<b>b</b>) with optimal parameters (<span class="html-italic">K</span> = 4, <span class="html-italic">α</span> = 140).</p>
Full article ">Figure 11
<p>The confusion diagram of fault diagnosis with different feature extraction methods. (<b>a</b>) RCmvMSE; (<b>b</b>) RCmvMFE; (<b>c</b>) RCmvMDE; (<b>d</b>) ERCmvMDE.</p>
Full article ">
17 pages, 221 KiB  
Review
Complexity, Uncertainty, and Entropy: Applications to Food Sensory Perception and Other Complex Phenomena
by Luka Sturtewagen, Harald van Mil and Erik van der Linden
Entropy 2025, 27(2), 191; https://doi.org/10.3390/e27020191 - 13 Feb 2025
Viewed by 322
Abstract
Complexity has been studied in various areas of science, such as ecology and sensory science. One important aspect is the quantification of the complexity of a system. There exist a multitude of different approaches. One approach relates complexity to information, with its measure [...] Read more.
Complexity has been studied in various areas of science, such as ecology and sensory science. One important aspect is the quantification of the complexity of a system. There exist a multitude of different approaches. One approach relates complexity to information, with its measure introduced by Shannon. This is equal to the negative value of entropy, or uncertainty. In this review, we discuss the different approaches that are used to quantify and measure complexity in the realm of food sensory perception. We address how the food sensory field could benefit from an approach that is based on an information-theoretical footing, to allow for one quantitative measure, thus improving comparison and reproducibility among different laboratories. Reversely, the review is intended to inspire the physics community to explore complex phenomena, such as sensory perception, in terms of a general information-theoretical measure, such as entropy and other fields. Full article
21 pages, 1531 KiB  
Article
Dual-Regularized Feature Selection for Class-Specific and Global Feature Associations
by Chenchen Wang, Jun Wang, Yanfei Li, Chengkai Piao and Jinmao Wei
Entropy 2025, 27(2), 190; https://doi.org/10.3390/e27020190 - 13 Feb 2025
Viewed by 252
Abstract
Understanding feature associations is vital for selecting the most informative features. Existing methods primarily focus on global feature associations, which capture overall relationships across all samples. However, they often overlook class-specific feature interactions, which are essential for capturing locality features that may only [...] Read more.
Understanding feature associations is vital for selecting the most informative features. Existing methods primarily focus on global feature associations, which capture overall relationships across all samples. However, they often overlook class-specific feature interactions, which are essential for capturing locality features that may only be significant within certain classes. In this paper, we propose Dual-Regularized Feature Selection (DRFS), which incorporates two feature association regularizers to address both class-specific and global feature relationships. The class-specific regularizer captures the local geometric structure of features within each class. Meanwhile, the global regularizer utilizes a global feature similarity matrix to eliminate redundant features across classes. By combining these regularizers, DRFS selects features that preserve both local interactions within each class and global discriminative power, with each regularizer complementing the other to enhance feature selection. Experimental results on eight public real-world datasets demonstrate that DRFS outperforms existing methods in classification accuracy. Full article
(This article belongs to the Special Issue Information-Theoretic Methods in Data Analytics)
Show Figures

Figure 1

Figure 1
<p>Comparison of global feature association-based methods (<b>a</b>) and our approach (<b>b</b>). (<b>a</b>) Existing methods compute feature associations across the entire dataset to preserve the global feature manifold or eliminate redundant features. (<b>b</b>) Our method retains class-specific feature manifolds while removing global feature redundancies.</p>
Full article ">Figure 2
<p>An overview of the proposed DRFS method.</p>
Full article ">Figure 3
<p>Classification accuracy of the SVM classifier with different number of selected features on eight datasets.</p>
Full article ">Figure 4
<p>Classification accuracy of the 1-NN classifier with different number of selected features on eight datasets.</p>
Full article ">Figure 5
<p>Critical difference diagram by the Bonferroni–Dunn post hoc test (significance level of 0.05).</p>
Full article ">Figure 6
<p>Parameter sensitivity study of DRFS with respect to <math display="inline"><semantics> <mi>α</mi> </semantics></math> and <math display="inline"><semantics> <mi>β</mi> </semantics></math> across all datasets. The parameters <math display="inline"><semantics> <mi>α</mi> </semantics></math> and <math display="inline"><semantics> <mi>β</mi> </semantics></math> correspond to the sparse norm and dual regularizations, respectively.</p>
Full article ">Figure 7
<p>Parameter sensitivity study of the class-specific and global regularizations across all datasets. The parameters <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>2</mn> </msub> </semantics></math> correspond to the class-specific and global regularizations, respectively.</p>
Full article ">Figure 8
<p>Accuracy of DRFS with varying feature neighbors <span class="html-italic">k</span>. “# Feas” denotes the number of selected features.</p>
Full article ">Figure 9
<p>Convergence curves of DRFS across all datasets.</p>
Full article ">
13 pages, 2862 KiB  
Article
Solving Flexible Job-Shop Scheduling Problems Based on Quantum Computing
by Kaihan Fu, Jianjun Liu, Miao Chen and Huiying Zhang
Entropy 2025, 27(2), 189; https://doi.org/10.3390/e27020189 - 13 Feb 2025
Viewed by 389
Abstract
Flexible job-shop scheduling problems (FJSPs) represent one of the most complex combinatorial optimization challenges. Modern production systems and control processes demand rapid decision-making in scheduling. To address this challenge, we propose a quantum computing approach for solving FJSPs. We propose a quadratic unconstrained [...] Read more.
Flexible job-shop scheduling problems (FJSPs) represent one of the most complex combinatorial optimization challenges. Modern production systems and control processes demand rapid decision-making in scheduling. To address this challenge, we propose a quantum computing approach for solving FJSPs. We propose a quadratic unconstrained binary optimization (QUBO) model to minimize the makespan of FJSPs, with the scheduling scheme encoded in the ground state of the Hamiltonian operator. The model is solved using a coherent Ising machine (CIM). Numerical experiments are conducted to evaluate and validate the performance and effectiveness of the CIM. The results demonstrate that quantum computing holds significant potential for solving FJSPs more efficiently than traditional computational methods. Full article
(This article belongs to the Special Issue Quantum Information: Working towards Applications)
Show Figures

Figure 1

Figure 1
<p>Gantt chart of a solution for the example (3 jobs × 3 machines FJSP, makespan 16). Orange represents job 1, blue represents job 2, and green represents job 3.</p>
Full article ">Figure 2
<p>Structure and principle of a coherent Ising machine.</p>
Full article ">Figure 3
<p>Diagram of the total energy value of the Hamiltonian of benchmark SSFJS05.</p>
Full article ">Figure 4
<p>Diagram of the maximum cut of the optical quantum computer: (<b>a</b>) 17 × 17 Ising matrix; (<b>b</b>) 85 × 85 Ising matrix; (<b>c</b>) 160 × 160 Ising matrix; (<b>d</b>) 193 × 193 Ising matrix; (<b>e</b>) 127 × 127 Ising matrix. The blue or green dot on the circumference indicates the phase state of the optical qubit after coherence, the blue indicates that the phase is positive (spin variable σ is “1”), and the green indicates that the phase is negative (spin variable σ is “−1”).</p>
Full article ">Figure 5
<p>Gantt chart solved by the optical quantum computer of benchmark SSFJS05. Red represents job 1, yellow represents job 2, and orange represents job 3.</p>
Full article ">
21 pages, 335 KiB  
Article
On the Global Practical Exponential Stability of h-Manifolds for Impulsive Reaction–Diffusion Cohen–Grossberg Neural Networks with Time-Varying Delays
by Gani Stamov, Trayan Stamov, Ivanka Stamova and Cvetelina Spirova
Entropy 2025, 27(2), 188; https://doi.org/10.3390/e27020188 - 12 Feb 2025
Viewed by 387
Abstract
In this paper, we focus on h-manifolds related to impulsive reaction–diffusion Cohen–Grossberg neural networks with time-varying delays. By constructing a new Lyapunov-type function and a comparison principle, sufficient conditions that guarantee the global practical exponential stability of specific states are established. The [...] Read more.
In this paper, we focus on h-manifolds related to impulsive reaction–diffusion Cohen–Grossberg neural networks with time-varying delays. By constructing a new Lyapunov-type function and a comparison principle, sufficient conditions that guarantee the global practical exponential stability of specific states are established. The states of interest are determined by the so-called h-manifolds, i.e., manifolds defined by a specific function h, which is essential for various applied problems in imposing constraints on their dynamics. The established criteria are less restrictive for the variable domain and diffusion coefficients. The effect of some uncertain parameters on the stability behavior is also considered and a robust practical stability analysis is proposed. In addition, the obtained h-manifolds’ practical stability results are applied to a bidirectional associative memory (BAM) neural network model with impulsive perturbations and time-varying delays. Appropriate examples are discussed. Full article
(This article belongs to the Special Issue Dynamics in Complex Neural Networks, 2nd Edition)
13 pages, 359 KiB  
Article
Entropic Probability and Context States
by Benjamin Schumacher and Michael D. Westmoreland
Entropy 2025, 27(2), 187; https://doi.org/10.3390/e27020187 - 12 Feb 2025
Viewed by 257
Abstract
In a previous paper, we introduced an axiomatic system for information thermodynamics, deriving an entropy function that includes both thermodynamic and information components. From this function, we derived an entropic probability distribution for certain uniform collections of states. Here, we extend the concept [...] Read more.
In a previous paper, we introduced an axiomatic system for information thermodynamics, deriving an entropy function that includes both thermodynamic and information components. From this function, we derived an entropic probability distribution for certain uniform collections of states. Here, we extend the concept of entropic probability to more general collections, augmenting the states by reservoir and context states. This leads to an abstract concept of free energy and establishes a relation between free energy, information erasure, and generalized work. Full article
(This article belongs to the Section Thermodynamics)
Show Figures

Figure 1

Figure 1
<p>A simple Maxwell’s demon.</p>
Full article ">Figure 2
<p>Extraction of work by dividing gas enclosure into unequal volumes.</p>
Full article ">
15 pages, 1088 KiB  
Article
A Novel Evaluation of Income Class Boundaries Using Inflection Points of Probability Density Functions: A Case Study of Brazil
by Rafael Bittencourt, Hernane Borges de Barros Pereira, Marcelo A. Moret, Ivan C. Da Cunha Lima and Serge Galam
Entropy 2025, 27(2), 186; https://doi.org/10.3390/e27020186 - 12 Feb 2025
Viewed by 367
Abstract
Categorizing a population into different income classes is important for creating effective policies and analyzing markets. Our study develops a statistical method based on a nationwide survey of income distribution. We use these data to create a cumulative distribution function with a metalogistic [...] Read more.
Categorizing a population into different income classes is important for creating effective policies and analyzing markets. Our study develops a statistical method based on a nationwide survey of income distribution. We use these data to create a cumulative distribution function with a metalogistic distribution and its probability density function. We propose a new way to divide the population into income classes by using the inflection points of the probability density function as the class boundaries. As a case study, we apply this method to income data from Brazil between 2012 and 2022. We identify five income classes, with both their boundaries and the distribution of the population changing over time. To check our approach, we calculate the Gini coefficient and find that our results closely match official figures, with a root mean square deviation of less than 1%. By using individual income instead of family income, we avoid distortions caused by the fact that poorer families tend to be larger than wealthier ones. In the end, we identify five main income classes, with their boundaries shifting each year, reflecting the changing nature of income distribution in society. Full article
Show Figures

Figure 1

Figure 1
<p>Mean absolute error (MAE) for each year from 2012 to 2022, for metalog quantile functions with different numbers of parameters (from <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> up to <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 2
<p>Quantile probability function of the logarithm of monthly <span class="html-italic">per capita</span> income for the population segments in <a href="#entropy-27-00186-t001" class="html-table">Table 1</a>. The curves are derived from fitting a metalogistic distribution function with <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>The figure presents the average Lorenz curve derived from the metalogistic treatment with <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> for the period 2012–2022. The Gini coefficient associated with this curve is <math display="inline"><semantics> <mrow> <mi>G</mi> <mo>=</mo> <mn>0.528</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Metalogistic PDF for the presidential terms: the last three years of 2011–2014 (<b>a</b>), 2015–2018 (<b>b</b>), and 2019–2022 (<b>c</b>).</p>
Full article ">Figure 5
<p>Second derivative of the PDF through the years 2012 (<b>a</b>) up to 2022 (<b>k</b>). The ascending PDF rates are represented in blue, while the descending rates are in red. The points where the curves change colors show the inflection points that define the boundaries of classes in our model.</p>
Full article ">Figure 6
<p>The bar plots illustrate the distribution of the population across different income classes (<b>a</b>) and the income thresholds defining each class (<b>b</b>) from 2012 to 2022.</p>
Full article ">
Previous Issue
Back to TopTop