[go: up one dir, main page]

Next Issue
Volume 22, April
Previous Issue
Volume 22, February
 
 
entropy-logo

Journal Browser

Journal Browser

Entropy, Volume 22, Issue 3 (March 2020) – 115 articles

Cover Story (view full-size image): Many efforts have been made to develop models and parameters to predict the glass-forming ability of different alloy systems. By combining basic thermodynamic calculations and empirically verified correlations between structural, thermodynamic, and kinetic parameters and the phase formation behavior of alloys, a fast computational model for the prediction of glass-forming alloy compositions and their critical cooling rates has been developed. In our report, we present the developed model and compare its behavior, as predicted, to that of the well-known bulk metallic glass-forming systems PdNiP, CuMgCa, CuZrTi, and CuZrTiNi. An algorithm has been constructed to pinpoint compositions with a minimum critical cooling rate to find promising new glass-forming compositions. View this paper.
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
14 pages, 2541 KiB  
Review
New Invariant Expressions in Chemical Kinetics
by Gregory S. Yablonsky, Daniel Branco, Guy B. Marin and Denis Constales
Entropy 2020, 22(3), 373; https://doi.org/10.3390/e22030373 - 24 Mar 2020
Cited by 6 | Viewed by 3192
Abstract
This paper presents a review of our original results obtained during the last decade. These results have been found theoretically for classical mass-action-law models of chemical kinetics and justified experimentally. In contrast with the traditional invariances, they relate to a special battery of [...] Read more.
This paper presents a review of our original results obtained during the last decade. These results have been found theoretically for classical mass-action-law models of chemical kinetics and justified experimentally. In contrast with the traditional invariances, they relate to a special battery of kinetic experiments, not a single experiment. Two types of invariances are distinguished and described in detail: thermodynamic invariants, i.e., special combinations of kinetic dependences that yield the equilibrium constants, or simple functions of the equilibrium constants; and “mixed” kinetico-thermodynamic invariances, functions both of equilibrium constants and non-thermodynamic ratios of kinetic coefficients. Full article
(This article belongs to the Special Issue Entropies: Between Information Geometry and Kinetics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Concentration profiles of A (black) and B (blue), starting from pure A (solid) and from pure B (dashed), assuming a single step reversible reaction A ⇄ B with k<sup>+</sup>= 2 s<sup>−1</sup>, k<sup>−</sup>= 1 s<sup>−1</sup>. The ratio between B<sub>A</sub>(t) (solid blue) and A<sub>B</sub>(t) (black dashed) is the equilibrium constant.</p>
Full article ">Figure 2
<p>From top to bottom, the Scaled Incremental Conversion (SIC) plots of C<sub>B</sub>, B<sub>B</sub>, B<sub>A</sub> and C<sub>A</sub>. k<sub>1</sub><sup>+</sup> = 4.5 s<sup>−1</sup>, k<sub>1</sub><sup>−</sup> = 10 s<sup>−1</sup>, k<sub>2</sub><sup>+</sup> = 6.5 s<sup>−1</sup>, k<sub>2</sub><sup>−</sup> = 12 s<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Concentration profiles of A<sub>BC</sub>, B<sub>AC</sub> and C<sub>AB</sub>. The first two, A<sub>BC</sub> and B<sub>AC</sub>, must be read with the left Y-scale. The last one, C<sub>AB</sub>, must be read with the right Y-scale. k<sub>1</sub><sup>+</sup> = 1.75 s<sup>−1</sup>, k<sub>1</sub><sup>−</sup> = 3.00 s<sup>−1</sup>, k<sub>2</sub><sup>+</sup> = 1.50 s<sup>−1</sup>, k<sub>2</sub><sup>−</sup> = 0.50 s<sup>−1</sup>.</p>
Full article ">Figure 4
<p>Corrected concentration profiles of A<sub>BC</sub>, B<sub>AC</sub> and C<sub>AB</sub>. From top to bottom, A<sub>BC</sub> − A<sub>eq</sub>, B<sub>AC</sub> − B<sub>eq</sub> and C<sub>AB</sub> − C<sub>eq</sub>. k<sub>1</sub><sup>+</sup> = 1.75 s<sup>−1</sup>, k<sub>1</sub><sup>−</sup> = 3.00 s<sup>−1</sup>, k<sub>2</sub><sup>+</sup> = 1.50 s<sup>−1</sup>, k<sub>2</sub><sup>−</sup> = 0.50 s<sup>−1</sup>.</p>
Full article ">
32 pages, 5963 KiB  
Review
On the Evidence of Thermodynamic Self-Organization during Fatigue: A Review
by Mehdi Naderi
Entropy 2020, 22(3), 372; https://doi.org/10.3390/e22030372 - 24 Mar 2020
Cited by 7 | Viewed by 5799
Abstract
In this review paper, the evidence and application of thermodynamic self-organization are reviewed for metals typically with single crystals subjected to cyclic loading. The theory of self-organization in thermodynamic processes far from equilibrium is a cutting-edge theme for the development of a new [...] Read more.
In this review paper, the evidence and application of thermodynamic self-organization are reviewed for metals typically with single crystals subjected to cyclic loading. The theory of self-organization in thermodynamic processes far from equilibrium is a cutting-edge theme for the development of a new generation of materials. It could be interpreted as the formation of globally coherent patterns, configurations and orderliness through local interactivities by “cascade evolution of dissipative structures”. Non-equilibrium thermodynamics, entropy, and dissipative structures connected to self-organization phenomenon (patterning, orderliness) are briefly discussed. Some example evidences are reviewed in detail to show how thermodynamics self-organization can emerge from a non-equilibrium process; fatigue. Evidences including dislocation density evolution, stored energy, temperature, and acoustic signals can be considered as the signature of self-organization. Most of the attention is given to relate an analogy between persistent slip bands (PSBs) and self-organization in metals with single crystals. Some aspects of the stability of dislocations during fatigue of single crystals are discussed using the formulation of excess entropy generation. Full article
(This article belongs to the Special Issue Review Papers for Entropy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Entropy exchange in an open system. <span class="html-italic">dS<sub>e</sub></span> is entropy flow through the boundaries of the open system and <span class="html-italic">dS<sub>i</sub></span> is the entropy production within the open system.</p>
Full article ">Figure 2
<p>Schematic of the cyclic stress-strain curve along with the evolution of dislocation pattern of copper single crystal for single slip [<a href="#B92-entropy-22-00372" class="html-bibr">92</a>] (with permission) (Images on this original figure quoted from [<a href="#B104-entropy-22-00372" class="html-bibr">104</a>,<a href="#B108-entropy-22-00372" class="html-bibr">108</a>,<a href="#B117-entropy-22-00372" class="html-bibr">117</a>,<a href="#B118-entropy-22-00372" class="html-bibr">118</a>]).</p>
Full article ">Figure 3
<p>Schematic of macroscopic evolution of dislocation configuration during fatigue: (<b>a</b>) dislocation sources; (<b>b</b>) dipole segment; (<b>c</b>) vein structure; and (<b>d</b>) persistent slip band (PSB)-ladder structure [<a href="#B93-entropy-22-00372" class="html-bibr">93</a>]. (with permission).</p>
Full article ">Figure 4
<p>PSBs in fatigued copper; (<b>a</b>) SEM of surface; (<b>b</b>) transmission electron microscopy (TEM) micrograph of PSBs with ladder structure between matrix of dipolar veins [<a href="#B92-entropy-22-00372" class="html-bibr">92</a>,<a href="#B113-entropy-22-00372" class="html-bibr">113</a>] (with permission).</p>
Full article ">Figure 5
<p>TEM image of dislocation patterns in polycrystalline copper-5%aluminum under cyclic plastic strain of 0.001: (<b>a</b>) PSB wall; (<b>b</b>) labyrinth; (<b>c</b>) cell wall; (<b>d</b>) wall structure [<a href="#B115-entropy-22-00372" class="html-bibr">115</a>] (with permission).</p>
Full article ">Figure 6
<p>The effect of orientation on the dislocation configurations of face-centered cubic (FCC) single crystals [<a href="#B121-entropy-22-00372" class="html-bibr">121</a>] (with permission).</p>
Full article ">Figure 7
<p>(<b>a</b>) hysteresis loop of first fatigue cycle with seven measurement points; (<b>b</b>) stress evolution pertained to peak hysteresis loop points of tension (point 7) and compression (point 4); (<b>c</b>) dislocation density and dislocation wall spacing as a function of the number of cycles at 1% strain [<a href="#B98-entropy-22-00372" class="html-bibr">98</a>,<a href="#B170-entropy-22-00372" class="html-bibr">170</a>] (with permission).</p>
Full article ">Figure 8
<p>(<b>a</b>) The evolution of stress and temperature differences between points 7 and 4 during fatigue life; (<b>b</b>) effective heat conduction rate versus fatigue cycles; comparison of the ratio of stored energy percentage and stress evolution during fatigue. (<b>a</b>,<b>b</b>) are obtained from [<a href="#B98-entropy-22-00372" class="html-bibr">98</a>] (with permission).</p>
Full article ">Figure 9
<p>Comparison of cyclic hardening and acoustic emission (AE) signals for copper single crystal [<a href="#B171-entropy-22-00372" class="html-bibr">171</a>]. AE features are energy, median frequency, and pulse position. Bauschinger energy parameter is related to the area of the hysteresis loop (with permission).</p>
Full article ">Figure 10
<p>TEM image of dislocation arrangement in polycrystalline copper at the stress level of 102 Mpa (<b>a</b>) without current pulse and fatigue failure life at 8 × 10<sup>5</sup> cycles, (<b>b</b>) with current pulse and fatigue failure life at 2.1 × 10<sup>6</sup> cycles [<a href="#B189-entropy-22-00372" class="html-bibr">189</a>] (with permission).</p>
Full article ">Figure 11
<p>(<b>a</b>) Variation of V<sub>H</sub> with cumulative strain for different plastic shear strain. Arrows show the first appearance of PSBs on surfaces [<a href="#B106-entropy-22-00372" class="html-bibr">106</a>] (with permission); (<b>b</b>) the derivative of V<sub>H</sub> with respect to cumulative plastic strain. Plastic shear strain amplitudes for J127, J130, J128, J129 specimen are 0.78, 1.45, 1.95, 3.02 (×10<sup>−3</sup>).</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) typical surface morphology evolution versus fatigued; (<b>d</b>) evolution of cyclic stress (σ<sub>a</sub>), hysteresis loop shape parameter V<sub>H</sub>, internal friction Q<sup>−1</sup>, and energy parameter β<sub>E</sub> for copper single crystal oriented for easy glide. Dashed line corresponds to bifurcation point as the onset of vein rupture and PSBs emergence. The figure is quoted from [<a href="#B172-entropy-22-00372" class="html-bibr">172</a>] (with permission).</p>
Full article ">
4 pages, 175 KiB  
Editorial
Entropy in Foundations of Quantum Physics
by Marcin Pawłowski
Entropy 2020, 22(3), 371; https://doi.org/10.3390/e22030371 - 24 Mar 2020
Viewed by 3138
Abstract
Entropy can be used in studies on foundations of quantum physics in many different ways, each of them using different properties of this mathematical object [...] Full article
(This article belongs to the Special Issue Entropy in Foundations of Quantum Physics)
28 pages, 4982 KiB  
Article
Theory, Analysis, and Applications of the Entropic Lattice Boltzmann Model for Compressible Flows
by Nicolò Frapolli, Shyam Chikatamarla and Ilya Karlin
Entropy 2020, 22(3), 370; https://doi.org/10.3390/e22030370 - 24 Mar 2020
Cited by 16 | Viewed by 5210
Abstract
The entropic lattice Boltzmann method for the simulation of compressible flows is studied in detail and new opportunities for extending operating range are explored. We address limitations on the maximum Mach number and temperature range allowed for a given lattice. Solutions to both [...] Read more.
The entropic lattice Boltzmann method for the simulation of compressible flows is studied in detail and new opportunities for extending operating range are explored. We address limitations on the maximum Mach number and temperature range allowed for a given lattice. Solutions to both these problems are presented by modifying the original lattices without increasing the number of discrete velocities and without altering the numerical algorithm. In order to increase the Mach number, we employ shifted lattices while the magnitude of lattice speeds is increased in order to extend the temperature range. Accuracy and efficiency of the shifted lattices are demonstrated with simulations of the supersonic flow field around a diamond-shaped and NACA0012 airfoil, the subsonic, transonic, and supersonic flow field around the Busemann biplane, and the interaction of vortices with a planar shock wave. For the lattices with extended temperature range, the model is validated with the simulation of the Richtmyer–Meshkov instability. We also discuss some key ideas of how to reduce the number of discrete speeds in three-dimensional simulations by pruning of the higher-order lattices, and introduce a new construction of the corresponding guided equilibrium by entropy minimization. Full article
(This article belongs to the Special Issue Entropies: Between Information Geometry and Kinetics)
Show Figures

Figure 1

Figure 1
<p>Visualization of the <span class="html-italic">D</span>2<span class="html-italic">Q</span>49 lattice.</p>
Full article ">Figure 2
<p>Deviation of the equilibrium moments from the Maxwell–Boltzmann values as a function of the turbulent and the advection Mach number <math display="inline"><semantics> <msub> <mi>Ma</mi> <mi mathvariant="normal">t</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>Ma</mi> <mi mathvariant="normal">a</mi> </msub> </semantics></math>, respectively, for three different lattices, <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>5</mn> <mn>2</mn> </msup> </semantics></math>, <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>7</mn> <mn>2</mn> </msup> </semantics></math>, and <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>11</mn> <mn>2</mn> </msup> </semantics></math>. (<b>a</b>) The <span class="html-italic">x</span>-component of the third-order moment <math display="inline"><semantics> <mrow> <msubsup> <mi>q</mi> <mi>x</mi> <mi>eq</mi> </msubsup> <mo>=</mo> <msubsup> <mo>∑</mo> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> </mrow> <mi>n</mi> </msubsup> <msub> <mi>v</mi> <mrow> <mi>i</mi> <mi>x</mi> </mrow> </msub> <msubsup> <mi>v</mi> <mi>i</mi> <mn>2</mn> </msubsup> <msubsup> <mi>f</mi> <mi>i</mi> <mi>eq</mi> </msubsup> </mrow> </semantics></math>; (<b>b</b>) The <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>x</mi> </mrow> </semantics></math>-component of fourth-order moment <math display="inline"><semantics> <mrow> <msubsup> <mi>R</mi> <mrow> <mi>x</mi> <mi>x</mi> </mrow> <mi>eq</mi> </msubsup> <mo>=</mo> <msubsup> <mo>∑</mo> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> </mrow> <mi>n</mi> </msubsup> <msub> <mi>v</mi> <mrow> <mi>i</mi> <mi>x</mi> </mrow> </msub> <msub> <mi>v</mi> <mrow> <mi>i</mi> <mi>x</mi> </mrow> </msub> <msubsup> <mi>v</mi> <mi>i</mi> <mn>2</mn> </msubsup> <msubsup> <mi>f</mi> <mi>i</mi> <mi>eq</mi> </msubsup> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Deviation of the equilibrium moments from the Maxwell–Boltzmann values as a function of the reduced temperature <math display="inline"><semantics> <mi>θ</mi> </semantics></math>. (<b>a</b>) The <span class="html-italic">x</span>-component of the third-order moment <math display="inline"><semantics> <msubsup> <mi>q</mi> <mi>x</mi> <mi>eq</mi> </msubsup> </semantics></math>; (<b>b</b>) The <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>x</mi> </mrow> </semantics></math>-component of the fourth-order moment <math display="inline"><semantics> <msubsup> <mi>R</mi> <mrow> <mi>x</mi> <mi>x</mi> </mrow> <mi>eq</mi> </msubsup> </semantics></math>. Results for the <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>5</mn> <mn>2</mn> </msup> </semantics></math> lattice at the local Mach <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi>loc</mi> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and for the <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>7</mn> <mn>2</mn> </msup> </semantics></math> lattice at local Mach <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi>loc</mi> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi>loc</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> are shown.</p>
Full article ">Figure 4
<p>Equilibrium populations as a function of the Mach number. (<b>a</b>) <math display="inline"><semantics> <msubsup> <mi>f</mi> <mn>0</mn> <mi>eq</mi> </msubsup> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <msubsup> <mi>f</mi> <mn>2</mn> <mi>eq</mi> </msubsup> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <msubsup> <mi>f</mi> <mrow> <mo>−</mo> <mn>2</mn> </mrow> <mi>eq</mi> </msubsup> </semantics></math>. Line: Third-order polynomial approximation; Dash: Fourth-order polynomial approximation; Symbol: Accurate numerical evaluation.</p>
Full article ">Figure 5
<p>Visualization of the shifted <span class="html-italic">D</span>2<span class="html-italic">Q</span>49 lattice with <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Deviation of the equilibrium moments from the Maxwell–Boltzmann values as a function of the <span class="html-italic">x</span>-component of the local velocity. (<b>a</b>) The third-order moment <math display="inline"><semantics> <msubsup> <mi>q</mi> <mi>x</mi> <mi>eq</mi> </msubsup> </semantics></math>; (<b>b</b>) The fourth-order moment <math display="inline"><semantics> <msubsup> <mi>R</mi> <mrow> <mi>x</mi> <mi>x</mi> </mrow> <mi>eq</mi> </msubsup> </semantics></math>. Line: Non-shifted lattice, <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>; Dash: Shifted lattice, <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>; Dotted-dash: Shifted lattice, <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Vortex advection comparison at different advection Mach number. Top row: symmetric lattice. Bottom row: shifted lattice with <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Grid convergence study for the Green–Taylor vortex. Results are shown for non-shifted lattice and lattices with the shift <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Deviation of the equilibrium moments from the Maxwell–Boltzmann values as a function of Mach number. (<b>a</b>) The third-order moment <math display="inline"><semantics> <msubsup> <mi>q</mi> <mi>x</mi> <mi>eq</mi> </msubsup> </semantics></math>; (<b>b</b>) The fourth-order moment <math display="inline"><semantics> <msubsup> <mi>R</mi> <mrow> <mi>x</mi> <mi>x</mi> </mrow> <mi>eq</mi> </msubsup> </semantics></math>. Line: The <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>7</mn> <mn>2</mn> </msup> </semantics></math>-0123 lattice. Dash: The <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>7</mn> <mn>2</mn> </msup> </semantics></math>-0124 lattice.</p>
Full article ">Figure 10
<p>Deviation of the equilibrium moments from the Maxwell–Boltzmann values as a function of the reduced temperature <math display="inline"><semantics> <mi>θ</mi> </semantics></math>, at the local Mach number <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi>loc</mi> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>. (<b>a</b>) The third-order moment <math display="inline"><semantics> <msubsup> <mi>q</mi> <mi>x</mi> <mi>eq</mi> </msubsup> </semantics></math>. (<b>b</b>) The fourth-order moment <math display="inline"><semantics> <msubsup> <mi>R</mi> <mrow> <mi>x</mi> <mi>x</mi> </mrow> <mi>eq</mi> </msubsup> </semantics></math>. Line: The <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>5</mn> <mn>2</mn> </msup> </semantics></math>-0123 lattice; Dash: The <span class="html-italic">D</span>2<span class="html-italic">Q</span><math display="inline"><semantics> <msup> <mn>7</mn> <mn>2</mn> </msup> </semantics></math>-0124 lattice.</p>
Full article ">Figure 11
<p>Steady state solution of the Mach number distribution around a diamond-shaped airfoil at the inlet <math display="inline"><semantics> <mrow> <mi>Ma</mi> <mo>=</mo> <mn>2.2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Re</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>AoA</mi> <mo>=</mo> <msup> <mn>3</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>Drag coefficient <math display="inline"><semantics> <msub> <mi>c</mi> <mi>d</mi> </msub> </semantics></math> as a function of the free stream Mach number for the Busemann biplane simulations. Reference: [<a href="#B30-entropy-22-00370" class="html-bibr">30</a>]. Inset: snapshots of the pressure distribution around the biplane for three different Mach numbers: <math display="inline"><semantics> <mrow> <mi>Ma</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, top; <math display="inline"><semantics> <mrow> <mi>Ma</mi> <mo>=</mo> <mn>1.7</mn> </mrow> </semantics></math>, bottom left; <math display="inline"><semantics> <mrow> <mi>Ma</mi> <mo>=</mo> <mn>2.0</mn> </mrow> </semantics></math>, bottom right. The pressure is shown between <math display="inline"><semantics> <mrow> <msub> <mi>p</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> (white) and <math display="inline"><semantics> <mrow> <msub> <mi>p</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <mn>1.9</mn> </mrow> </semantics></math> (pink), in lattice units.</p>
Full article ">Figure 13
<p>Snapshot of the temperature around the NACA0012 airfoil with a free stream Mach of <math display="inline"><semantics> <mrow> <mi>Ma</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, Reynolds number <math display="inline"><semantics> <mrow> <mi>Re</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>4</mn> </msup> </mrow> </semantics></math>, and angle of attack <math display="inline"><semantics> <mrow> <mi>AoA</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 14
<p>Pressure coefficient in front of the airfoil, on the airfoil surface and behind the airfoil for the simulation of the NACA0012 airfoil at free stream Mach of <math display="inline"><semantics> <mrow> <mi>Ma</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, a Reynolds number <math display="inline"><semantics> <mrow> <mi>Re</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>4</mn> </msup> </mrow> </semantics></math>, and an angle of attack <math display="inline"><semantics> <mrow> <mi>AoA</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 15
<p>Snapshot of the density in the vortex–shock interaction simulation. Left: DNS [<a href="#B32-entropy-22-00370" class="html-bibr">32</a>]; Right: ELBM. <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi mathvariant="normal">a</mi> </msub> <mo>=</mo> <mn>1.2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi mathvariant="normal">v</mi> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Re</mi> <mo>=</mo> <mn>400</mn> </mrow> </semantics></math>. Contour levels are from <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.92</mn> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <mn>1.55</mn> </mrow> </semantics></math> with an increment of <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>ρ</mi> <mo>=</mo> <mn>0.0053</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>The sound pressure distribution <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>p</mi> </mrow> </semantics></math> in the radial and tangential directions for the case <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi mathvariant="normal">a</mi> </msub> <mo>=</mo> <mn>1.2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi mathvariant="normal">v</mi> </msub> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Re</mi> <mo>=</mo> <mn>800</mn> </mrow> </semantics></math>. (<b>a</b>) The radial sound pressure distribution <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>p</mi> </mrow> </semantics></math> measured at an angle <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mo>−</mo> <msup> <mn>45</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> with respect to the <span class="html-italic">x</span>-axis. Times: <math display="inline"><semantics> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>8</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>. (<b>b</b>) The tangential sound pressure distribution <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>p</mi> </mrow> </semantics></math> measured at two radii, <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>3.7</mn> </mrow> </semantics></math>. Time: <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math>. Symbol: ELBM. Line: DNS [<a href="#B32-entropy-22-00370" class="html-bibr">32</a>].</p>
Full article ">Figure 16 Cont.
<p>The sound pressure distribution <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>p</mi> </mrow> </semantics></math> in the radial and tangential directions for the case <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi mathvariant="normal">a</mi> </msub> <mo>=</mo> <mn>1.2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi mathvariant="normal">v</mi> </msub> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Re</mi> <mo>=</mo> <mn>800</mn> </mrow> </semantics></math>. (<b>a</b>) The radial sound pressure distribution <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>p</mi> </mrow> </semantics></math> measured at an angle <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mo>−</mo> <msup> <mn>45</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> with respect to the <span class="html-italic">x</span>-axis. Times: <math display="inline"><semantics> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>8</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>. (<b>b</b>) The tangential sound pressure distribution <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>p</mi> </mrow> </semantics></math> measured at two radii, <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>3.7</mn> </mrow> </semantics></math>. Time: <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math>. Symbol: ELBM. Line: DNS [<a href="#B32-entropy-22-00370" class="html-bibr">32</a>].</p>
Full article ">Figure 17
<p>The decay of the peak sound pressure at different Mach numbers. Symbol: ELBM. Lines: DNS [<a href="#B32-entropy-22-00370" class="html-bibr">32</a>].</p>
Full article ">Figure 18
<p>Snapshot of the sound pressure, <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi mathvariant="normal">a</mi> </msub> <mo>=</mo> <mn>1.2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>Ma</mi> <mi mathvariant="normal">v</mi> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Re</mi> <mo>=</mo> <mn>400</mn> </mrow> </semantics></math>. Left: DNS [<a href="#B32-entropy-22-00370" class="html-bibr">32</a>], ⊕ and ⊖ indicate positive and negative sound pressure, respectively; Right: ELBM. Contour levels are from <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>p</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <mn>0.459</mn> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>p</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <mn>0.301</mn> </mrow> </semantics></math> with an increment of <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>ρ</mi> <mo>=</mo> <mn>0.027</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 19
<p>Snapshots of the evolution of the density for the Richtmyer–Meshkov instability. Time instants from left to right and from top to bottom; Top row: <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mo>−</mo> <mn>0.03</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0.09</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>1.13</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>2.29</mn> </mrow> </semantics></math>. Bottom row: <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>3.44</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>4.60</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>5.76</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>6.92</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 20
<p>Normalized amplitude growth rate for the Richtmyer–Meshkov instability simulation as a function of normalized time. Solid line: ELBM; Symbol: the experiment [<a href="#B34-entropy-22-00370" class="html-bibr">34</a>]; Dashed line: the WENO simulation [<a href="#B35-entropy-22-00370" class="html-bibr">35</a>]; Dotted line: the original analytical prediction of Richtmyer [<a href="#B36-entropy-22-00370" class="html-bibr">36</a>] for the initial growth.</p>
Full article ">Figure 21
<p>Snapshot of the flow field around the Onera M6 wing in a transonic flow at <math display="inline"><semantics> <mrow> <mi>Ma</mi> <mo>=</mo> <mn>0.839</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Re</mi> <mo>=</mo> <mn>11.7</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>AoA</mi> <mo>=</mo> <msup> <mn>3.06</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>—the streamlines colored by vorticity and the iso-surface of the sonic condition are shown.</p>
Full article ">Figure 22
<p>Pressure coefficient <math display="inline"><semantics> <msub> <mi>c</mi> <mi>p</mi> </msub> </semantics></math> on the Onera M6 wing at three wing sections in the stream-wise direction. (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>/</mo> <mi>b</mi> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>/</mo> <mi>b</mi> <mo>=</mo> <mn>0.65</mn> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>/</mo> <mi>b</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>. Line: ELBM. Symbol: Experiment and Euler solver [<a href="#B37-entropy-22-00370" class="html-bibr">37</a>].</p>
Full article ">Figure 23
<p>Pressure coefficient <math display="inline"><semantics> <msub> <mi>c</mi> <mi>p</mi> </msub> </semantics></math> on the Onera M6 wing at three sectional span-wise positions. (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>C</mi> <mo>=</mo> <mn>0.27</mn> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>C</mi> <mo>=</mo> <mn>0.69</mn> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>C</mi> <mo>=</mo> <mn>0.89</mn> </mrow> </semantics></math>. Line: ELBM. Symbol: Experiment and Euler solver [<a href="#B37-entropy-22-00370" class="html-bibr">37</a>].</p>
Full article ">Figure 24
<p>Visualization of the <math display="inline"><semantics> <mrow> <mi>D</mi> <mn>3</mn> <mi>Q</mi> <mn>39</mn> </mrow> </semantics></math> lattice. Red: <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, orange: <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>, yellow: <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, green: <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>8</mn> </mrow> </semantics></math>, blue: <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>9</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 25
<p>Vortex advection comparison at different advection Mach number. Top row: <math display="inline"><semantics> <mrow> <mi>D</mi> <mn>3</mn> <mi>Q</mi> <mn>343</mn> </mrow> </semantics></math> lattice with the standard entropic equilibrium. Bottom row: <math display="inline"><semantics> <mrow> <mi>D</mi> <mn>3</mn> <mi>Q</mi> <mn>39</mn> </mrow> </semantics></math> lattice with the guided entropic equilibrium.</p>
Full article ">
11 pages, 2355 KiB  
Article
PGNet: Pipeline Guidance for Human Key-Point Detection
by Feng Hong, Changhua Lu, Chun Liu, Ruru Liu, Weiwei Jiang, Wei Ju and Tao Wang
Entropy 2020, 22(3), 369; https://doi.org/10.3390/e22030369 - 24 Mar 2020
Cited by 9 | Viewed by 3877
Abstract
Human key-point detection is a challenging research field in computer vision. Convolutional neural models limit the number of parameters and mine the local structure, and have made great progress in significant target detection and key-point detection. However, the features extracted by shallow layers [...] Read more.
Human key-point detection is a challenging research field in computer vision. Convolutional neural models limit the number of parameters and mine the local structure, and have made great progress in significant target detection and key-point detection. However, the features extracted by shallow layers mainly contain a lack of semantic information, while the features extracted by deep layers contain rich semantic information but a lack of spatial information that results in information imbalance and feature extraction imbalance. With the complexity of the network structure and the increasing amount of computation, the balance between the time of communication and the time of calculation highlights the importance. Based on the improvement of hardware equipment, network operation time is greatly improved by optimizing the network structure and data operation methods. However, as the network structure becomes deeper and deeper, the communication consumption between networks also increases, and network computing capacity is optimized. In addition, communication overhead is also the focus of recent attention. We propose a novel network structure PGNet, which contains three parts: pipeline guidance strategy (PGS); Cross-Distance-IoU Loss (CIoU); and Cascaded Fusion Feature Model (CFFM). Full article
(This article belongs to the Special Issue Entropy in Image Analysis II)
Show Figures

Figure 1

Figure 1
<p>The proposed network to find key points of the human body.</p>
Full article ">Figure 2
<p>The number of positive bounding boxes after the NMS, grouped by their IoU with the matched ground truth. In traditional NMS (blue bar), a significant portion of accurately localized bounding boxes get mistakenly suppressed due to the misalignment of classification confidence and localization accuracy, while IoU-guided NMS (yellow bar) preserves more accurately localized bounding boxes [<a href="#B35-entropy-22-00369" class="html-bibr">35</a>].</p>
Full article ">Figure 3
<p>An of overview of proposed PGNet.ResNet-50 is used as the backbone. Using the cascaded fusion feature model (CFFM), the backbone network is divided into 5 stages, and the feature-guided network after the image is convolved is used to extract key-point features.</p>
Full article ">Figure 4
<p>An example pipeline-parallel assignment with four machines and an example timeline at one of machines, highlighting the temporal overlap of computation and activation/gradient communication.</p>
Full article ">Figure 5
<p><math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">L</mi> <mrow> <mi>C</mi> <mi>D</mi> <mi>I</mi> <mi>o</mi> <mi>U</mi> </mrow> </msub> </mrow> </semantics></math> Distribution of bounding boxes for iterative training.</p>
Full article ">Figure 6
<p>Comparison of epoch trained by this method and epoch of other training methods.</p>
Full article ">
18 pages, 7466 KiB  
Article
Entropy as a Measure of Attractiveness and Socioeconomic Complexity in Rio de Janeiro Metropolitan Area
by Maxime Lenormand, Horacio Samaniego, Júlio César Chaves, Vinícius da Fonseca Vieira, Moacyr Alvim Horta Barbosa da Silva and Alexandre Gonçalves Evsukoff
Entropy 2020, 22(3), 368; https://doi.org/10.3390/e22030368 - 23 Mar 2020
Cited by 16 | Viewed by 4966
Abstract
Defining and measuring spatial inequalities across the urban environment remains a complex and elusive task which has been facilitated by the increasing availability of large geolocated databases. In this study, we rely on a mobile phone dataset and an entropy-based metric to measure [...] Read more.
Defining and measuring spatial inequalities across the urban environment remains a complex and elusive task which has been facilitated by the increasing availability of large geolocated databases. In this study, we rely on a mobile phone dataset and an entropy-based metric to measure the attractiveness of a location in the Rio de Janeiro Metropolitan Area (Brazil) as the diversity of visitors’ location of residence. The results show that the attractiveness of a given location measured by entropy is an important descriptor of the socioeconomic status of the location, and can thus be used as a proxy for complex socioeconomic indicators. Full article
(This article belongs to the Special Issue Information Theory for Human and Social Processes)
Show Figures

Figure 1

Figure 1
<p><b>Rio de Janeiro Metropolitan Area (RJMA).</b> The RJMA is composed of 49 locations—16 municipalities outside the Capital represented in grey and 33 subdistricts inside the Capital grouped into 5 districts.</p>
Full article ">Figure 2
<p><b>Results of the clustering analysis.</b> Log-log scatter plot of (<b>a</b>) the attractiveness and (<b>b</b>) the radius of attraction in terms of the entropy index. The inset in (a) shows the relationship after removing one outlier (cluster C4). Each dot represents a location within the study area. Indicators have been averaged over the work-shift time period during weekdays.</p>
Full article ">Figure 3
<p>Map of the RJMA that display the spatial distribution of four clusters.</p>
Full article ">Figure 4
<p><b>Zoomed-in view of Rio de Janeiro city. (a)</b> Favela subdistricts (in purple) and business center (in orange) locations. We focus the discussion on five locations: (1) Complexo do Alemão, (2) Jacarezinho, (3) Rocinha, (4) Complexo da Maré, and (5) Cidade de Deus. (<b>b</b>) Clusters spatial distribution. (<b>c</b>) Municipal Human Development Index (MHDI) from 2013. (<b>d</b>) Social Progress Index (IPS) from 2016.</p>
Full article ">Figure 5
<p><b>Economic analysis.</b> Number of jobs (<b>a</b>) and income (in Brazilian Reals) (<b>b</b>) as a function of the entropy index. The entropy have been averaged over the work-shift time periods on weekdays.</p>
Full article ">Figure 6
<p><b>Sociodemographic analysis.</b> Percentage of primary school level education ((<b>a</b>), high school level education (<b>b</b>), black people (<b>c</b>), and white people (<b>d</b>) as a function of the entropy index. The entropy have been averaged over the work-shift time periods on weekdays.</p>
Full article ">Figure 7
<p><b>Social development indexes.</b> MHDI (<b>a</b>) and IPS (<b>b</b>) as a function of the entropy index. The entropy has been averaged over the work-shift time periods on weekdays.</p>
Full article ">Figure 8
<p><b>Temporal evolution of the three metrics.</b> From the top to the bottom—entropy, attractiveness, and radius of attraction as a function of time by cluster. The values are averaged by cluster and normalized by the value obtained for the work shift during weekdays. A similar plot displaying boxplots instead of average values is available in <a href="#app4-entropy-22-00368" class="html-app">Appendix D</a>.</p>
Full article ">Figure A1
<p>Spatial distribution of antennas with their respective Voronoi polygons.</p>
Full article ">Figure A2
<p>Spatial overlap between the aggregation of Voronoi cells and the districts’ spatial polygons.</p>
Full article ">Figure A3
<p><b>Number of calls per hour and the partition of time shifts.</b> Total number of calls made in the RJMA in 2014 (including weekdays and weekends).</p>
Full article ">Figure A4
<p>Number of mobile phone users with an identified residence in the RJMA as a function of the number of inhabitants in the 49 locations.</p>
Full article ">Figure A5
<p>Ratio between the within-group variance and the total variance as a function of the number of clusters.</p>
Full article ">Figure A6
<p><b>Global Domestic Product (GDP) as a function of the entropy index.</b> The entropy have been averaged over the work-shift time periods on weekdays.</p>
Full article ">Figure A7
<p><b>Temporal evolution of the three metrics.</b> From the top to the bottom, Tukey boxplots of the entropy, attractiveness, and radius of attraction as a function of time by cluster. The values are normalized by the value obtained for the work shift during weekdays.</p>
Full article ">
18 pages, 6821 KiB  
Article
Compound Fault Diagnosis of Rolling Bearing Based on Singular Negentropy Difference Spectrum and Integrated Fast Spectral Correlation
by Guiji Tang and Tian Tian
Entropy 2020, 22(3), 367; https://doi.org/10.3390/e22030367 - 23 Mar 2020
Cited by 17 | Viewed by 3285
Abstract
Compound fault diagnosis is challenging due to the complexity, diversity and non-stationary characteristics of mechanical complex faults. In this paper, a novel compound fault separation method based on singular negentropy difference spectrum (SNDS) and integrated fast spectral correlation (IFSC) is proposed. Firstly, the [...] Read more.
Compound fault diagnosis is challenging due to the complexity, diversity and non-stationary characteristics of mechanical complex faults. In this paper, a novel compound fault separation method based on singular negentropy difference spectrum (SNDS) and integrated fast spectral correlation (IFSC) is proposed. Firstly, the original signal was de-noised by SNDS which improved the noise reduction effect of singular difference spectrum by introducing negative entropy. Secondly, the de-noised signal was analyzed by fast spectral correlation. Finally, IFSC took the fourth-order energy as the index to determine the resonance band and separate the fault features of different single fault. The proposed method is applied to analyze the simulated compound signals and the experimental vibration signals, the results show that the proposed method has excellent performance in the separation of rolling bearing composite faults. Full article
(This article belongs to the Special Issue Entropy-Based Algorithms for Signal Processing)
Show Figures

Figure 1

Figure 1
<p>Some application examples of rotating machinery [<a href="#B1-entropy-22-00367" class="html-bibr">1</a>].</p>
Full article ">Figure 2
<p>The schematic diagram of singular negentropy difference spectrum (SNDS).</p>
Full article ">Figure 3
<p>The flow chart of the proposed method.</p>
Full article ">Figure 4
<p>(<b>a</b>) Time domain waveform of the simulation signals; (<b>b</b>) the envelope spectrum of <span class="html-italic">x</span>(<span class="html-italic">t</span>).</p>
Full article ">Figure 5
<p>Negentropy of first 50 points and the singular negentropy difference spectrum (SNDS).</p>
Full article ">Figure 6
<p>Time domain waveform of the reconstructed signal.</p>
Full article ">Figure 7
<p>The fast spectral correlation spectrum of the reconstructed signal.</p>
Full article ">Figure 8
<p>Screening of resonance bands</p>
Full article ">Figure 9
<p>Outer ring fault after separation: (<b>a</b>) the fast spectral correlation spectrum; (<b>b</b>) the enhanced envelope spectrum.</p>
Full article ">Figure 10
<p>Inner ring fault after separation: (<b>a</b>) the fast spectral correlation spectrum; (<b>b</b>) the enhanced envelope spectrum.</p>
Full article ">Figure 11
<p>Singular value of first 50 points and singular difference spectrum (SDS).</p>
Full article ">Figure 12
<p>Time domain waveform of the reconstructed signal.</p>
Full article ">Figure 13
<p>The fast spectral correlation spectrum of the reconstructed signal.</p>
Full article ">Figure 14
<p>Screening of resonance bands</p>
Full article ">Figure 15
<p>Outer ring fault after separation: (<b>a</b>) the fast spectral correlation spectrum; (<b>b</b>) the enhanced envelope spectrum.</p>
Full article ">Figure 16
<p>Inner ring fault after separation: (<b>a</b>) the fast spectral correlation spectrum; (<b>b</b>) the enhanced envelope spectrum.</p>
Full article ">Figure 17
<p>Time-domain waveform after wavelet denoising.</p>
Full article ">Figure 18
<p>Kurtogram of the denoised signal.</p>
Full article ">Figure 19
<p>The results of separation: (<b>a</b>) the envelope spectrum of band 1; (<b>b</b>) the envelope spectrum of band 2.</p>
Full article ">Figure 20
<p>(<b>a</b>) Experiment stand; (<b>b</b>) sensors distribution and (<b>c</b>) rolling bearing with inner ring fault and outer ring fault.</p>
Full article ">Figure 21
<p>(<b>a</b>) Time domain waveform of the original signal; (<b>b</b>) the envelope spectrum of the original signal.</p>
Full article ">Figure 22
<p>Negentropy of the first 50 points and singular negentropy difference spectrum (SNDS).</p>
Full article ">Figure 23
<p>Time domain waveform of the reconstructed signal.</p>
Full article ">Figure 24
<p>The fast spectral correlation spectrum of the reconstructed signal.</p>
Full article ">Figure 25
<p>Screening of resonance bands.</p>
Full article ">Figure 26
<p>Outer ring fault after separation: (<b>a</b>) the fast spectral correlation spectrum; (<b>b</b>) the enhanced envelope spectrum.</p>
Full article ">Figure 27
<p>Inner ring fault after separation: (<b>a</b>) the fast spectral correlation spectrum; (<b>b</b>) the enhanced envelope spectrum.</p>
Full article ">Figure 28
<p>Singular value of the first 50 points and singular difference spectrum (SDS).</p>
Full article ">Figure 29
<p>Time domain waveform of the reconstructed signal.</p>
Full article ">Figure 30
<p>The fast spectral correlation spectrum of the reconstructed signal.</p>
Full article ">Figure 31
<p>Screening of resonance bands.</p>
Full article ">Figure 32
<p>Outer ring fault after separation: (<b>a</b>) the fast spectral correlation spectrum; (<b>b</b>) the enhanced envelope spectrum.</p>
Full article ">Figure 33
<p>Inner ring fault after separation: (<b>a</b>) the fast spectral correlation spectrum; (<b>b</b>) the enhanced envelope spectrum.</p>
Full article ">Figure 34
<p>Time-domain waveform after wavelet denoising.</p>
Full article ">Figure 35
<p>Kurtogram of denoised signal.</p>
Full article ">Figure 36
<p>The results of separation: (<b>a</b>) the envelope spectrum of band 1; (<b>b</b>) the envelope spectrum of band 2.</p>
Full article ">
12 pages, 2473 KiB  
Article
Optimal Digital Implementation of Fractional-Order Models in a Microcontroller
by Mariusz Matusiak, Marcin Bąkała and Rafał Wojciechowski
Entropy 2020, 22(3), 366; https://doi.org/10.3390/e22030366 - 23 Mar 2020
Cited by 4 | Viewed by 3710
Abstract
The growing number of operations in implementations of the non-local fractional differentiation operator is cumbersome for real applications with strict performance and memory storage requirements. This demands use of one of the available approximation methods. In this paper, the analysis of the classic [...] Read more.
The growing number of operations in implementations of the non-local fractional differentiation operator is cumbersome for real applications with strict performance and memory storage requirements. This demands use of one of the available approximation methods. In this paper, the analysis of the classic integer- (IO) and fractional-order (FO) models of the brushless DC (BLDC) micromotor mounted on a steel rotating arms, and next, the discretization and efficient implementation of the models in a microcontroller (MCU) is performed. Two different methods for the FO model are examined, including the approximation of the fractional-order operator s ν ( ν R ) using the Oustaloup Recursive filter and the numerical evaluation of the fractional differintegral operator based on the Grünwald–Letnikov definition and Short Memory Principle. The models are verified against the results of several experiments conducted on an ARM Cortex-M7-based STM32F746ZG unit. Additionally, some software optimization techniques for the Cortex-M microcontroller family are discussed. The described steps are universal and can also be easily adapted to any other microcontroller. The values for integral absolute error (IAE) and integral square error (ISE) performance indices, calculated on the basis of simulations performed in MATLAB, are used to evaluate accuracy. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Block scheme of the testing hardware platform (1-BLDC micromotor, 2-high-precision encoder, 3-adjustable arm, 5-controller).</p>
Full article ">Figure 2
<p>UAV arm testing platform with the BLDC motor and microcontroller (1-BLDC micromotor, 2-high-precision encoder, 3-adjustable arm, 4-rigid frame, 5-controllers, 6-power supply).</p>
Full article ">Figure 3
<p>Oustaloup approximations of order <span class="html-italic">N</span> for fractional-order models. Step responses.</p>
Full article ">Figure 4
<p>Oustaloup approximations of order <span class="html-italic">N</span> for fractional-order models. Bode diagram.</p>
Full article ">Figure 5
<p>Measured microcontroller outputs with implemented integer-order (1st and 2nd) models.</p>
Full article ">Figure 6
<p>Measured microcontroller outputs with fractional-order (ORA) models.</p>
Full article ">Figure 7
<p>Heaviside step response evaluated by the GL method and different memory lengths <span class="html-italic">N</span>.</p>
Full article ">Figure 8
<p>Software optimizations of the numerical algorithm for calculating output of model (<a href="#FD3-entropy-22-00366" class="html-disp-formula">3</a>) with the maximum number of CPU cycles allowed (red dotted line).</p>
Full article ">
13 pages, 381 KiB  
Article
Some Remarks about Entropy of Digital Filtered Signals
by Vinícius S. Borges, Erivelton G. Nepomuceno, Carlos A. Duque and Denis N. Butusov
Entropy 2020, 22(3), 365; https://doi.org/10.3390/e22030365 - 23 Mar 2020
Cited by 3 | Viewed by 2941
Abstract
The finite numerical resolution of digital number representation has an impact on the properties of filters. Much effort has been done to develop efficient digital filters investigating the effects in the frequency response. However, it seems that there is less attention to the [...] Read more.
The finite numerical resolution of digital number representation has an impact on the properties of filters. Much effort has been done to develop efficient digital filters investigating the effects in the frequency response. However, it seems that there is less attention to the influence in the entropy by digital filtered signals due to the finite precision. To contribute in such a direction, this manuscript presents some remarks about the entropy of filtered signals. Three types of filters are investigated: Butterworth, Chebyshev, and elliptic. Using a boundary technique, the parameters of the filters are evaluated according to the word length of 16 or 32 bits. It has been shown that filtered signals have their entropy increased even if the filters are linear. A significant positive correlation (p < 0.05) was observed between order and Shannon entropy of the filtered signal using the elliptic filter. Comparing to signal-to-noise ratio, entropy seems more efficient at detecting the increasing of noise in a filtered signal. Such knowledge can be used as an additional condition for designing digital filters. Full article
(This article belongs to the Section Information Theory, Probability and Statistics)
Show Figures

Figure 1

Figure 1
<p>Adaptation of “Schematic diagram of a general communication system” in [<a href="#B44-entropy-22-00365" class="html-bibr">44</a>]. In our case, we are interested in looking the channel as a filter and noise source as a consequence of finite precision implementation of the digital filters.</p>
Full article ">Figure 2
<p>Procedure of standardization of the vector <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>(</mo> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math> in panel (<b>a</b>) for entropy calculation using Equation (<a href="#FD15-entropy-22-00365" class="html-disp-formula">15</a>). The result <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>(</mo> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math> in panel (<b>b</b>) is composed of only integers within 0 to <math display="inline"><semantics> <mrow> <msup> <mn>2</mn> <mrow> <mi>W</mi> <mi>L</mi> </mrow> </msup> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math>. In this case, we have used <math display="inline"><semantics> <mrow> <mi>W</mi> <mi>L</mi> <mo>=</mo> <mn>8</mn> </mrow> </semantics></math> bits. Panel (<b>c</b>) shows a random signal with uniform distribution. Fifty runs of this signal produces an entropy of <math display="inline"><semantics> <mrow> <mi>H</mi> <mo>=</mo> <mn>7.59</mn> <mo>±</mo> <mn>0.03</mn> </mrow> </semantics></math>. Increasing the number of samples, this value approaches 8, as expected.</p>
Full article ">Figure 3
<p>Computation of Shannon entropy for three signals. All signals have been standardized according procedure described in Equation (<a href="#FD15-entropy-22-00365" class="html-disp-formula">15</a>). (<b>a</b>) Sine wave <math display="inline"><semantics> <mrow> <mn>2</mn> <mi>s</mi> <mi>i</mi> <mi>n</mi> <mo>(</mo> <mn>2</mn> <mi>π</mi> <mn>2</mn> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>. (<b>b</b>) Sine wave added with Gaussian noise of <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. (<b>c</b>) Sine wave added Gaussian noise of <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. The calculated entropy are (<b>a</b>) <math display="inline"><semantics> <mrow> <mn>5.66</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mn>5.95</mn> <mo>±</mo> <mn>0.03</mn> </mrow> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <mrow> <mn>6.23</mn> <mo>±</mo> <mn>0.04</mn> </mrow> </semantics></math>. The level of Gaussian noise is quite unseeingly; yet the entropy has been sensitive for the increasing of noise. Panels (<b>d</b>–<b>f</b>) show a zoom in the above figure to see the presence of a small level of noise in the signals of panels (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>) FFT of signal 2, <a href="#entropy-22-00365-t003" class="html-table">Table 3</a>—Complete signal; (<b>b</b>) FFT of signal 2 in <a href="#entropy-22-00365-t003" class="html-table">Table 3</a>—Ideally Filtered signal; (<b>c</b>) FFT of Chebyshev filter. The FFT computations the expected similarity between the signals. This is another point that makes relevant to investigate the effect of digital filter in the entropy of the filtered signal.</p>
Full article ">
11 pages, 2175 KiB  
Article
Research on the Node Importance of a Weighted Network Based on the K-Order Propagation Number Algorithm
by Pingchuan Tang, Chuancheng Song, Weiwei Ding, Junkai Ma, Jun Dong and Liya Huang
Entropy 2020, 22(3), 364; https://doi.org/10.3390/e22030364 - 22 Mar 2020
Cited by 15 | Viewed by 3823
Abstract
To describe both the global and local characteristics of a network more comprehensively, we propose the weighted K-order propagation number (WKPN) algorithm to extract the disease propagation based on the network topology to evaluate the node importance. Each node is set as [...] Read more.
To describe both the global and local characteristics of a network more comprehensively, we propose the weighted K-order propagation number (WKPN) algorithm to extract the disease propagation based on the network topology to evaluate the node importance. Each node is set as the source of infection, and the total number of infected nodes is defined as the K-order propagation number after experiencing the propagation time K. The simulation of the symmetric network with bridge nodes indicated that the WKPN algorithm was more effective for evaluation of the algorithm features. A deliberate attack strategy, which indicated an attack on the network according to the node importance from high to low, was employed to evaluate the WKPN algorithm in real networks. Compared with the other methods tested, the results demonstrate the applicability and advancement that a lower number of nodes, with a higher importance calculated by the K-order propagation number algorithm, has to achieve full damage to the network structure. Full article
(This article belongs to the Special Issue Entropic Forces in Complex Systems)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The weighted network topology with a symmetric network with bridge nodes.</p>
Full article ">Figure 2
<p>The networks graph structure: (<b>a</b>) the Science Museum visitor network; (<b>b</b>) the Facebook forum network; (<b>c</b>) the non-US airport route network; and (<b>d</b>) the US 500 busiest commercial airports network.</p>
Full article ">Figure 3
<p><span class="html-italic">K</span>-order structure entropy <math display="inline"><semantics> <msup> <mrow> <mi>H</mi> </mrow> <mi>K</mi> </msup> </semantics></math> varies with K: (<b>a</b>) the Science Museum visitor network; (<b>b</b>) the Facebook forum network; (<b>c</b>) the non-US airport route network; and (<b>d</b>) the US 500 busiest commercial airports network.</p>
Full article ">Figure 4
<p>The network efficiency decline rate varies with the attack times: (<b>a</b>) the Science Museum visitor network; (<b>b</b>) the Facebook forum network; (<b>c</b>) the non-US airport route network; and (<b>d</b>) the US 500 busiest commercial airports network.</p>
Full article ">Figure 5
<p>The node number of maximum sub-graph <math display="inline"><semantics> <mi>γ</mi> </semantics></math> in the network varies with the number of attacks: (<b>a</b>) the Science Museum visitor network; (<b>b</b>) the Facebook forum network; (<b>c</b>) the non-US airport route network; and (<b>d</b>) the US 500 busiest commercial airports network.</p>
Full article ">
10 pages, 5402 KiB  
Article
Numerical Analysis on Natural Convection Heat Transfer in a Single Circular Fin-Tube Heat Exchanger (Part 1): Numerical Method
by Jong Hwi Lee, Jong-Hyeon Shin, Se-Myong Chang and Taegee Min
Entropy 2020, 22(3), 363; https://doi.org/10.3390/e22030363 - 21 Mar 2020
Cited by 11 | Viewed by 4960
Abstract
In this research, unsteady three-dimensional incompressible Navier–Stokes equations are solved to simulate experiments with the Boussinesq approximation and validate the proposed numerical model for the design of a circular fin-tube heat exchanger. Unsteady time marching is proposed for a time sweeping analysis of [...] Read more.
In this research, unsteady three-dimensional incompressible Navier–Stokes equations are solved to simulate experiments with the Boussinesq approximation and validate the proposed numerical model for the design of a circular fin-tube heat exchanger. Unsteady time marching is proposed for a time sweeping analysis of various Rayleigh numbers. The accuracy of the natural convection data of a single horizontal circular tube with the proposed numerical method can be guaranteed when the Rayleigh number based on the tube diameter exceeds 400, which is regarded as the limitation of numerical errors due to instability. Moreover, the effective limit for a circular fin-tube heat exchanger is reached when the Rayleigh number based on the fin gap size ( Ra s ) is equal to or exceeds 100. This is because at low Rayleigh numbers, the air gap between the fins is isolated and rarely affected by natural convection of the outer air, where the fluid provides heat resistance. Thus, the fin acts favorably when Ra s exceeds 100. Full article
(This article belongs to the Section Thermodynamics)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of circular fin-tube heat exchanger studied in the present work.</p>
Full article ">Figure 2
<p>Numerical domain and grid in the present work: (<b>a</b>) computational domain, (<b>b</b>) grid (<b>c</b>) zoom in near the fin-tube.</p>
Full article ">Figure 3
<p>Comparison of Nusselt numbers with experimental correlation of circular tubes.</p>
Full article ">Figure 4
<p>Comparison of wall temperature with experimental data of circular tubes.</p>
Full article ">Figure 5
<p>Comparison of Nu<sub>L</sub> with experimental data of circular fin-tubes. (Experimental Data from Kang and Chang [<a href="#B10-entropy-22-00363" class="html-bibr">10</a>]).</p>
Full article ">Figure 6
<p>Velocity contours of the airflow in the present numerical experiment: (<b>a</b>) <span class="html-italic">s/D</span> = 0.119, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Ra</mi> </mrow> <mi>s</mi> </msub> </mrow> </semantics></math> = 15; (<b>b</b>) <span class="html-italic">s/D</span> = 0.256, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Ra</mi> </mrow> <mi>s</mi> </msub> </mrow> </semantics></math> = 150.</p>
Full article ">Figure 6 Cont.
<p>Velocity contours of the airflow in the present numerical experiment: (<b>a</b>) <span class="html-italic">s/D</span> = 0.119, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Ra</mi> </mrow> <mi>s</mi> </msub> </mrow> </semantics></math> = 15; (<b>b</b>) <span class="html-italic">s/D</span> = 0.256, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Ra</mi> </mrow> <mi>s</mi> </msub> </mrow> </semantics></math> = 150.</p>
Full article ">Figure 7
<p>Isotherm lines for air-sides in the present numerical experiment: (<b>a</b>) <span class="html-italic">s/D</span> = 0.119, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Ra</mi> </mrow> <mi>s</mi> </msub> </mrow> </semantics></math> = 15; (<b>b</b>) <span class="html-italic">s/D</span> = 0.256, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Ra</mi> </mrow> <mi>s</mi> </msub> </mrow> </semantics></math> = 150.</p>
Full article ">Figure 8
<p>Variation of fin efficiency, <span class="html-italic">η<sub>f</sub></span> with non-dimensional fin gap, <span class="html-italic">s</span>/<span class="html-italic">D</span>.</p>
Full article ">
36 pages, 1140 KiB  
Review
Particle Swarm Optimisation: A Historical Review Up to the Current Developments
by Diogo Freitas, Luiz Guerreiro Lopes and Fernando Morgado-Dias
Entropy 2020, 22(3), 362; https://doi.org/10.3390/e22030362 - 21 Mar 2020
Cited by 150 | Viewed by 14993
Abstract
The Particle Swarm Optimisation (PSO) algorithm was inspired by the social and biological behaviour of bird flocks searching for food sources. In this nature-based algorithm, individuals are referred to as particles and fly through the search space seeking for the global best position [...] Read more.
The Particle Swarm Optimisation (PSO) algorithm was inspired by the social and biological behaviour of bird flocks searching for food sources. In this nature-based algorithm, individuals are referred to as particles and fly through the search space seeking for the global best position that minimises (or maximises) a given problem. Today, PSO is one of the most well-known and widely used swarm intelligence algorithms and metaheuristic techniques, because of its simplicity and ability to be used in a wide range of applications. However, in-depth studies of the algorithm have led to the detection and identification of a number of problems with it, especially convergence problems and performance issues. Consequently, a myriad of variants, enhancements and extensions to the original version of the algorithm, developed and introduced in the mid-1990s, have been proposed, especially in the last two decades. In this article, a systematic literature review about those variants and improvements is made, which also covers the hybridisation and parallelisation of the algorithm and its extensions to other classes of optimisation problems, taking into consideration the most important ones. These approaches and improvements are appropriately summarised, organised and presented, in order to allow and facilitate the identification of the most appropriate PSO variant for a particular application. Full article
(This article belongs to the Special Issue Intelligent Tools and Applications in Engineering and Mathematics)
Show Figures

Figure 1

Figure 1
<p>Summary of the most important convergence improvements developed for pso.</p>
Full article ">Figure 2
<p>Summary of the most important architecture strategies developed for pso.</p>
Full article ">Figure 3
<p>Summary of the most important applications of pso.</p>
Full article ">
16 pages, 3140 KiB  
Article
Information Graphs Incorporating Predictive Values of Disease Forecasts
by Gareth Hughes, Jennifer Reed and Neil McRoberts
Entropy 2020, 22(3), 361; https://doi.org/10.3390/e22030361 - 20 Mar 2020
Cited by 4 | Viewed by 3012
Abstract
Diagrammatic formats are useful for summarizing the processes of evaluation and comparison of forecasts in plant pathology and other disciplines where decisions about interventions for the purpose of disease management are often based on a proxy risk variable. We describe a new diagrammatic [...] Read more.
Diagrammatic formats are useful for summarizing the processes of evaluation and comparison of forecasts in plant pathology and other disciplines where decisions about interventions for the purpose of disease management are often based on a proxy risk variable. We describe a new diagrammatic format for disease forecasts with two categories of actual status and two categories of forecast. The format displays relative entropies, functions of the predictive values that characterize expected information provided by disease forecasts. The new format arises from a consideration of earlier formats with underlying information properties that were previously unexploited. The new diagrammatic format requires no additional data for calculation beyond those used for the calculation of a receiver operating characteristic (ROC) curve. While an ROC curve characterizes a forecast in terms of sensitivity and specificity, the new format described here characterizes a forecast in terms of relative entropies based on predictive values. Thus it is complementary to ROC methodology in its application to the evaluation and comparison of forecasts. Full article
(This article belongs to the Special Issue Applications of Information Theory to Epidemiology)
Show Figures

Figure 1

Figure 1
<p>Biggerstaff’s likelihood ratios graph for Scenario B (reference) and Scenario C (comparison). The graph for Scenario B consists of a single point at 1–<span class="html-italic">specificity</span> = 0.156, <span class="html-italic">sensitivity</span> = 0.833 (see <a href="#entropy-22-00361-t002" class="html-table">Table 2</a>). The solid red line through (0, 0) and (0.156, 0.833) has slope = <span class="html-italic">sensitivity</span>/(1–<span class="html-italic">specificity</span>) = 5.333 = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math>. The dashed red line through (0.156, 0.833) and (1, 1) has slope = (1–<span class="html-italic">sensitivity</span>)/<span class="html-italic">specificity</span> = 0.198 = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math>. The graph for Scenario C consists of a single point at 1–<span class="html-italic">specificity</span> = 0.01, <span class="html-italic">sensitivity</span> = 0.39 (see <a href="#entropy-22-00361-t002" class="html-table">Table 2</a>). The solid green line through (0.01, 0.39) and (1, 1) has slope = <span class="html-italic">sensitivity</span>/(1–<span class="html-italic">specificity</span>) = 39.0 = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math>. The dashed green line through (0.156, 0.833) and (1, 1) has slope = (1–<span class="html-italic">sensitivity</span>)/<span class="html-italic">specificity</span> = 0.616 = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Biggerstaff’s likelihood ratios graph for Scenario D (reference) and Scenario E (comparison). The graph for Scenario D consists of a single point at 1–<span class="html-italic">specificity</span> = 0.156, <span class="html-italic">sensitivity</span> = 0.833 (see <a href="#entropy-22-00361-t002" class="html-table">Table 2</a>). The solid red line through (0, 0) and (0.156, 0.833) has slope = <span class="html-italic">sensitivity</span>/(1–<span class="html-italic">specificity</span>) = 5.333 = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math>. The dashed red line through (0.156, 0.833) and (1, 1) has slope = (1–<span class="html-italic">sensitivity</span>)/<span class="html-italic">specificity</span> = 0.198 = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math>. The graph for Scenario E consists of a single point at 1–<span class="html-italic">specificity</span> = 0.344, <span class="html-italic">sensitivity</span> = 0.944 (see <a href="#entropy-22-00361-t002" class="html-table">Table 2</a>). The solid blue line through (0, 0) and (0.344, 0.944) has slope = <span class="html-italic">sensitivity</span>/(1–<span class="html-italic">specificity</span>) = 2.744 = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math>. The dashed blue line through (0.344, 0.944) and (1, 1) has slope = (1–<span class="html-italic">sensitivity</span>)/<span class="html-italic">specificity</span> = 0.085 = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Biggerstaff’s likelihood ratios graphs for Scenarios A, B and D (<a href="#entropy-22-00361-t002" class="html-table">Table 2</a>). The slopes of the lines are the likelihood ratios <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 5.333 and <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.198, calculated from <a href="#entropy-22-00361-t002" class="html-table">Table 2</a>. Analysis shows that the lines themselves are also iso-information contours for the expected information contents of + and – forecasts. However, the calculated values of these expected information contents depend on the prior probability as well as on <span class="html-italic">sensitivity</span> and <span class="html-italic">specificity</span>. Making use of the available data on the prior probabilities allows us to calculate relative entropies in order to distinguish analytically between scenarios, but the likelihood ratios graph does not distinguish visually between scenarios with the same <span class="html-italic">sensitivity</span> and <span class="html-italic">specificity</span>.</p>
Full article ">Figure 4
<p>A version of Johnson’s log<sub>10</sub> likelihood ratios diagram for data from <a href="#entropy-22-00361-t002" class="html-table">Table 2</a>. Here <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.727 and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math>= −0.704 for Scenarios A, B and D (■). For Scenario C (<span style="color:#00B050">■</span>), <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 1.591 and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math>= −0.208. For Scenario E (<span style="color:#0070C0">■</span>), <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.438 and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = −1.071. Valid comparisons (i.e., for scenarios with equal prior probabilities) are Scenario B (reference) with Scenario C (comparison) and Scenario D (reference) with Scenario E (comparison).</p>
Full article ">Figure 5
<p>The “north-west” region of the figure is characterized by Equation (13), so relates to + predictions (which are correct for <span class="html-italic">c</span> subjects and incorrect for <span class="html-italic">nc</span> subjects). <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mi>L</mi> <mo>+</mo> </msub> </mrow> </semantics></math> contours are always straight lines with slope = 1. The solid red line indicates the contour for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.727 Hart, corresponding to Scenarios A, B, and D (<a href="#entropy-22-00361-t002" class="html-table">Table 2</a>). A correct + prediction has a large information content when <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> is small (B), and a small information content is when <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> is large (D) (the arrow indicates the direction of increasing <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> along the contour). As the information content <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <mrow> <mo>[</mo> <mrow> <mrow> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> <mrow> <msub> <mo>|</mo> <mo>+</mo> </msub> </mrow> </mrow> <mo>/</mo> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </mrow> </mrow> <mo>]</mo> </mrow> </mrow> </semantics></math> (on the vertical axis) becomes decreasingly positive, the information content <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <mrow> <mo>[</mo> <mrow> <mrow> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>n</mi> <mi>c</mi> </mrow> </msub> <mrow> <msub> <mo>|</mo> <mo>+</mo> </msub> </mrow> </mrow> <mo>/</mo> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>n</mi> <mi>c</mi> </mrow> </msub> </mrow> </mrow> </mrow> <mo>]</mo> </mrow> </mrow> </semantics></math> (on the horizontal axis) becomes increasingly negative. The “south-east” region of the figure is characterized by Equation (14), so relates to − predictions (which are correct for <span class="html-italic">nc</span> subjects and incorrect for <span class="html-italic">c</span> subjects). <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mi>L</mi> <mo>−</mo> </msub> </mrow> </semantics></math> contours are always straight lines with slope = 1. The dashed red line indicates the contour for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <msub> <mover accent="true"> <mi>L</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = −0.704 Hart, corresponding to Scenarios A, B, and D (<a href="#entropy-22-00361-t002" class="html-table">Table 2</a>). A correct − prediction has a large information content when <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>n</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> is small (D), and a small information content is when <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>n</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> is large (B) (the arrow indicates the direction of increasing <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>n</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> along the contour, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>n</mi> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo>−</mo> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math>). As the information content <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <mrow> <mo>[</mo> <mrow> <mrow> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>n</mi> <mi>c</mi> </mrow> </msub> <mrow> <msub> <mo>|</mo> <mo>−</mo> </msub> </mrow> </mrow> <mo>/</mo> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>n</mi> <mi>c</mi> </mrow> </msub> </mrow> </mrow> </mrow> <mo>]</mo> </mrow> </mrow> </semantics></math> (on the horizontal axis) becomes decreasingly positive, the information content <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>log</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <mrow> <mo>[</mo> <mrow> <mrow> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> <mrow> <msub> <mo>|</mo> <mo>−</mo> </msub> </mrow> </mrow> <mo>/</mo> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </mrow> </mrow> <mo>]</mo> </mrow> </mrow> </semantics></math> (on the vertical axis) becomes increasingly negative.</p>
Full article ">Figure 6
<p>Scenario A: from the data in <a href="#entropy-22-00361-t002" class="html-table">Table 2</a>, we calculate relative entropies <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.315, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.179 (both in nats) (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.36) (Equations (3) and (4)). Similarly, for Scenario B we calculate <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.171, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.024 nats (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.05) and for Scenario D, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.076, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.289 nats (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.85).</p>
Full article ">Figure 7
<p>The prior probability <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> contour for Scenarios A, B, and D (solid red line). The contour is calibrated at 0.1 intervals of <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math>, clockwise from the origin, 0.1 to 0.9 (+ symbol on curve). Scenarios B (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.05), A (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.36), and D (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.85) as characterized in <a href="#entropy-22-00361-t002" class="html-table">Table 2</a> are indicated (■). Also indicated on the prior probability contour: maximum <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.337 nats (▲) (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.245), maximum <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.317 nats (▲) (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.749), <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math>= 0.251 nats (●) (<math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> = 0.513).</p>
Full article ">Figure 8
<p>The dashed curve is the prior probability <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> contour showing the upper limit for performance of any binary predictor. The contour is calibrated at 0.1 intervals of <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> from upper left to lower right, 0.1 to 0.9 (+ symbol on curve). The maximum relative entropy for a + test result increases indefinitely as <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> approaches 0 while the maximum relative entropy for a – test result increases indefinitely as <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>p</mi> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> </mrow> </semantics></math> approaches 1. The prior probability contour for Scenarios A, B, and D from <a href="#entropy-22-00361-f007" class="html-fig">Figure 7</a> (solid red line) is also shown, for reference (note the rescaled axes).</p>
Full article ">Figure 9
<p>The prior probability contours for Scenarios C (solid green line) and E (solid blue line). Starting at the origin, the green prior probability contour passes through points (clockwise from origin): Scenario C, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 1.399, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.004 (prior = 0.05) (<span style="color:#00B050">■</span>); maximum <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 1.436 (prior = 0.073) (▲); maximum <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.029 (prior = 0.580) (▲). This contour does not coincide with the main diagonal of the plot other than at the origin. Starting at the origin, the blue prior probability contour passes through points (clockwise from origin): <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.080 (●) (prior = 0.109); maximum <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.126 (prior = 0.337) (▲); Scenario E, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>+</mo> </msub> </mrow> </semantics></math> = 0.039, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.700 (prior = 0.850) (<span style="color:#0070C0">■</span>); maximum <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>I</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> </msub> </mrow> </semantics></math> = 0.701 (prior = 0.842) (point obscured from view). The prior probability contour for Scenarios A, B, and D (solid red line) is included here for reference; clockwise from origin, points marked ■ indicate Scenarios B, A and D (see <a href="#entropy-22-00361-f007" class="html-fig">Figure 7</a> for details). The dashed curve shows the contour indicating the upper limit for performance of a binary predictor (see <a href="#entropy-22-00361-f008" class="html-fig">Figure 8</a> for details). Note the changes in the scales on the axes compared with <a href="#entropy-22-00361-f007" class="html-fig">Figure 7</a> and <a href="#entropy-22-00361-f008" class="html-fig">Figure 8</a>.</p>
Full article ">
12 pages, 3921 KiB  
Article
Bundled Causal History Interaction
by Peishi Jiang and Praveen Kumar
Entropy 2020, 22(3), 360; https://doi.org/10.3390/e22030360 - 20 Mar 2020
Cited by 4 | Viewed by 2910
Abstract
Complex systems arise as a result of the nonlinear interactions between components. In particular, the evolutionary dynamics of a multivariate system encodes the ways in which different variables interact with each other individually or in groups. One fundamental question that remains unanswered is: [...] Read more.
Complex systems arise as a result of the nonlinear interactions between components. In particular, the evolutionary dynamics of a multivariate system encodes the ways in which different variables interact with each other individually or in groups. One fundamental question that remains unanswered is: How do two non-overlapping multivariate subsets of variables interact to causally determine the outcome of a specific variable? Here, we provide an information-based approach to address this problem. We delineate the temporal interactions between the bundles in a probabilistic graphical model. The strength of the interactions, captured by partial information decomposition, then exposes complex behavior of dependencies and memory within the system. The proposed approach successfully illustrated complex dependence between cations and anions as determinants of pH in an observed stream chemistry system. In the studied catchment, the dynamics of pH is a result of both cations and anions through mainly synergistic effects of the two and their individual influences as well. This example demonstrates the potentially broad applicability of the approach, establishing the foundation to study the interaction between groups of variables in a range of complex systems. Full article
(This article belongs to the Section Information Theory, Probability and Statistics)
Show Figures

Figure 1

Figure 1
<p>(color online) Illustration of pairwise interaction (<b>a</b>) and different types of multivariate interactions (<b>b</b>–<b>d</b>): (<b>a</b>) the influence from a source variable to a target variable; and (<b>b</b>–<b>d</b>) the influences on the target variable from two individual variables, a group of variables, and two groups of variables, respectively.</p>
Full article ">Figure 2
<p>(color online) Illustration of bundled causal history analysis framework by using a system consisting of seven components: (<b>a</b>) the influence from immediate (<math display="inline"><semantics> <msubsup> <mover accent="true"> <mi>X</mi> <mo stretchy="false">→</mo> </mover> <mrow> <mi>t</mi> <mo>−</mo> <msub> <mi>τ</mi> <mi mathvariant="double-struck">C</mi> </msub> <mo>:</mo> <mi>t</mi> </mrow> <mrow> <mi>m</mi> <mi>n</mi> </mrow> </msubsup> </semantics></math>) and distant (<math display="inline"><semantics> <msubsup> <mover accent="true"> <mi>X</mi> <mo stretchy="false">→</mo> </mover> <mrow> <mo>:</mo> <mi>t</mi> <mo>−</mo> <msub> <mi>τ</mi> <mi mathvariant="double-struck">C</mi> </msub> </mrow> <mrow> <mi>m</mi> <mi>n</mi> </mrow> </msubsup> </semantics></math>) bundled causal histories to the target (<math display="inline"><semantics> <msubsup> <mi>X</mi> <mi>t</mi> <mi>tar</mi> </msubsup> </semantics></math>) in Equation (<a href="#FD3-entropy-22-00360" class="html-disp-formula">3</a>); and (<b>b</b>) the different components in Equation (<a href="#FD5-entropy-22-00360" class="html-disp-formula">5</a>) based on using the Markov property for DAG as well as the Order-1 approximation <math display="inline"><semantics> <msub> <mover accent="true"> <mi>F</mi> <mo stretchy="false">→</mo> </mover> <mn>1</mn> </msub> </semantics></math> for <math display="inline"><semantics> <msub> <mover accent="true"> <mi>F</mi> <mo stretchy="false">→</mo> </mover> <mrow/> </msub> </semantics></math>.</p>
Full article ">Figure 3
<p>(color online) Illustration of using Momentary Information Weighted Transitive Reduction (MIWTR) to reduce the dimensionality of <math display="inline"><semantics> <msub> <mover accent="true"> <mi>W</mi> <mo stretchy="false">→</mo> </mover> <msub> <mi>τ</mi> <mi mathvariant="double-struck">C</mi> </msub> </msub> </semantics></math> for the present state of <span class="html-italic">pH</span> influenced by the selected cation and anion groups, with <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mi mathvariant="double-struck">C</mi> </msub> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math>. (<b>a</b>) The used stream chemistry time-series data recorded in the Upper Hafren catchment in the United Kingdom [<a href="#B21-entropy-22-00360" class="html-bibr">21</a>]. (<b>b</b>) The estimated directed acyclic graph (DAG) using Tigramite algorithm [<a href="#B8-entropy-22-00360" class="html-bibr">8</a>,<a href="#B18-entropy-22-00360" class="html-bibr">18</a>,<a href="#B19-entropy-22-00360" class="html-bibr">19</a>,<a href="#B20-entropy-22-00360" class="html-bibr">20</a>], with the original identified <math display="inline"><semantics> <msub> <mover accent="true"> <mi>W</mi> <mo stretchy="false">→</mo> </mover> <msub> <mi>τ</mi> <mi mathvariant="double-struck">C</mi> </msub> </msub> </semantics></math> (orange nodes) and <math display="inline"><semantics> <mover accent="true"> <mi>V</mi> <mo stretchy="false">→</mo> </mover> </semantics></math> (blue nodes) in Equation (<a href="#FD5-entropy-22-00360" class="html-disp-formula">5</a>) (the edges removed by MIWTR are highlighted in magenta). (<b>c</b>) The reduced <math display="inline"><semantics> <msub> <mover accent="true"> <mi>W</mi> <mo stretchy="false">→</mo> </mover> <msub> <mi>τ</mi> <mi mathvariant="double-struck">C</mi> </msub> </msub> </semantics></math> (orange nodes) using MIWTR.</p>
Full article ">Figure 4
<p>(color online) Plots of partial information decomposition of the immediate (<math display="inline"><semantics> <msub> <mi mathvariant="script">J</mi> <mrow> <mi>m</mi> <mi>n</mi> </mrow> </msub> </semantics></math>) and distant (<math display="inline"><semantics> <msub> <mi mathvariant="script">D</mi> <mrow> <mi>m</mi> <mi>n</mi> </mrow> </msub> </semantics></math>) bundled causal histories based on the stream solute time-series data and the estimated directed acyclic graph for time-series in <a href="#entropy-22-00360-f002" class="html-fig">Figure 2</a> under the Order-0 (<b>left</b>) and Order-1 (<b>right</b>) of approximation for <math display="inline"><semantics> <msub> <mover accent="true"> <mi>F</mi> <mo stretchy="false">→</mo> </mover> <mrow/> </msub> </semantics></math> by using <span class="html-italic">k</span>-nearest-neighbor estimators with <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>∈</mo> <mo stretchy="false">{</mo> <mn>5</mn> <mo>,</mo> <mn>6</mn> <mo>,</mo> <mn>7</mn> <mo>,</mo> <mn>8</mn> <mo>,</mo> <mn>9</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>15</mn> <mo stretchy="false">}</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>(color online) Plots of partial information decomposition of the immediate (<math display="inline"><semantics> <msub> <mi mathvariant="script">J</mi> <mrow> <mi>m</mi> <mi>n</mi> </mrow> </msub> </semantics></math>) and distant (<math display="inline"><semantics> <msub> <mi mathvariant="script">D</mi> <mrow> <mi>m</mi> <mi>n</mi> </mrow> </msub> </semantics></math>) bundled causal histories based on the stream solute time-series data and the estimated directed acyclic graph for time-series in <a href="#entropy-22-00360-f003" class="html-fig">Figure 3</a>b with the Order-0 (<b>top</b>) and Order-1 (<b>bottom</b>) of approximation for <math display="inline"><semantics> <msub> <mover accent="true"> <mi>F</mi> <mo stretchy="false">→</mo> </mover> <mrow/> </msub> </semantics></math>. The <math display="inline"><semantics> <msub> <mi mathvariant="script">J</mi> <mrow> <mi>m</mi> <mi>n</mi> </mrow> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi mathvariant="script">D</mi> <mrow> <mi>m</mi> <mi>n</mi> </mrow> </msub> </semantics></math> (Equation (<a href="#FD4-entropy-22-00360" class="html-disp-formula">4</a>a,b)) components are separated by a black dotted line.</p>
Full article ">
16 pages, 414 KiB  
Article
Underdetermined DOA Estimation for Wideband Signals via Focused Atomic Norm Minimization
by Juan Shi, Qunfei Zhang, Weijie Tan, Linlin Mao, Lihuan Huang and Wentao Shi
Entropy 2020, 22(3), 359; https://doi.org/10.3390/e22030359 - 20 Mar 2020
Cited by 9 | Viewed by 3117
Abstract
In underwater acoustic signal processing, direction of arrival (DOA) estimation can provide important information for target tracking and localization. To address underdetermined wideband signal processing in underwater passive detection system, this paper proposes a novel underdetermined wideband DOA estimation method equipped with the [...] Read more.
In underwater acoustic signal processing, direction of arrival (DOA) estimation can provide important information for target tracking and localization. To address underdetermined wideband signal processing in underwater passive detection system, this paper proposes a novel underdetermined wideband DOA estimation method equipped with the nested array (NA) using focused atomic norm minimization (ANM), where the signal source number detection is accomplished by information theory criteria. In the proposed DOA estimation method, especially, after vectoring the covariance matrix of each frequency bin, each corresponding obtained vector is focused into the predefined frequency bin by focused matrix. Then, the collected averaged vector is considered as virtual array model, whose steering vector exhibits the Vandermonde structure in terms of the obtained virtual array geometries. Further, the new covariance matrix is recovered based on ANM by semi-definite programming (SDP), which utilizes the information of the Toeplitz structure. Finally, the Root-MUSIC algorithm is applied to estimate the DOAs. Simulation results show that the proposed method outperforms other underdetermined DOA estimation methods based on information theory in term of higher estimation accuracy. Full article
(This article belongs to the Special Issue Entropy and Information Theory in Acoustics)
Show Figures

Figure 1

Figure 1
<p>A two-level NA with <math display="inline"><semantics> <msub> <mi>M</mi> <mn>1</mn> </msub> </semantics></math> sensors in the inner sub-array and <math display="inline"><semantics> <msub> <mi>M</mi> <mn>2</mn> </msub> </semantics></math> sensors in the outer sub-array.</p>
Full article ">Figure 2
<p>Weight coefficient of co-array of NA with six sensors.</p>
Full article ">Figure 3
<p>Probability of detection versus SNRs.</p>
Full article ">Figure 4
<p>Normalized spatial spectrum for <math display="inline"><semantics> <mrow> <mi>K</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> sources when <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>N</mi> <mi>R</mi> <mo>=</mo> <mo>−</mo> <mn>5</mn> </mrow> </semantics></math> dB.</p>
Full article ">Figure 5
<p>Normalized spatial spectrum for <math display="inline"><semantics> <mrow> <mi>K</mi> <mo>=</mo> <mn>7</mn> </mrow> </semantics></math> sources when <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>N</mi> <mi>R</mi> <mo>=</mo> <mo>−</mo> <mn>5</mn> </mrow> </semantics></math> dB.</p>
Full article ">Figure 6
<p>RMSE of DOA estimation for <math display="inline"><semantics> <mrow> <mi>K</mi> <mo>=</mo> <mn>7</mn> </mrow> </semantics></math> sources versus SNRs when <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>RMSE of DOA estimation for <math display="inline"><semantics> <mrow> <mi>K</mi> <mo>=</mo> <mn>7</mn> </mrow> </semantics></math> sources versus the number of snapshots when <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>N</mi> <mi>R</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> dB.</p>
Full article ">
14 pages, 6139 KiB  
Article
Numerical Analysis on Natural Convection Heat Transfer in a Single Circular Fin-Tube Heat Exchanger (Part 2): Correlations for Limiting Cases
by Jong Hwi Lee, Young Woo Son and Se-Myong Chang
Entropy 2020, 22(3), 358; https://doi.org/10.3390/e22030358 - 20 Mar 2020
Cited by 4 | Viewed by 6205
Abstract
This research focused on the correlations associated with the physics of natural convection in circular fin-tube models. The limiting conditions are defined by two conditions. The lower limit ( D o / D → 1, s/D = finite value) corresponds to [...] Read more.
This research focused on the correlations associated with the physics of natural convection in circular fin-tube models. The limiting conditions are defined by two conditions. The lower limit ( D o / D → 1, s/D = finite value) corresponds to a horizontal circular tube, while the upper limit ( D o / D → ∞, s/D << 1) corresponds to vertical flat plates. In this paper, we proposed a corrected correlation based on empirical result. The circular fin-tube heat exchanger was divided into the A and B types, the categorizing criteria being D o / D = 1.2 , where D and D o refer to the diameter of the circular tube and the diameter of the circular fin, respectively. Moreover, with the computational fluid dynamics technique used to investigate the limiting conditions, the parametric range was extended substantially in this research for type B, namely 1.2 < D o / D ≤ 10. The complex correlation was also simplified to the form Nu L = C Ra s n , where C and n are the functions of the diameter ratio D o / D . Full article
(This article belongs to the Section Thermodynamics)
Show Figures

Figure 1

Figure 1
<p>Primary regime between two extreme conditions for natural convection in the circular fin-tube configuration. (from Kang &amp; Chang [<a href="#B9-entropy-22-00358" class="html-bibr">9</a>]).</p>
Full article ">Figure 2
<p>Schematic diagram of circular fin-tube heat exchanger studied in the present work.</p>
Full article ">Figure 3
<p>Schematic diagram of lowest cases on limit model.</p>
Full article ">Figure 4
<p>Variation of Nusselt number with Ra<sub>s</sub> in the lowest limit model.</p>
Full article ">Figure 5
<p>Variation of Nusselt number with Ra<span class="html-italic"><sub>D</sub></span>.</p>
Full article ">Figure 6
<p>Schematic diagram of highest cases on limit model.</p>
Full article ">Figure 7
<p>Variation of Nusselt number with Ra<sub>s</sub> in the highest limit model.</p>
Full article ">Figure 8
<p>The value of the power <span class="html-italic">n</span> according to the diameter ratio of the circular fin-tube.</p>
Full article ">Figure 9
<p>Variation of Nusselt number with <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Ra</mi> </mrow> <mi>s</mi> </msub> </mrow> </semantics></math>: Classification criteria for type A and B.</p>
Full article ">Figure 10
<p>Variation of Nusselt number and trend line with <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Ra</mi> </mrow> <mi>s</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Variation of C value with circular fin-tube diameter ratio.</p>
Full article ">Figure 12
<p>Variation of n value with circular fin-tube diameter ratio.</p>
Full article ">Figure 13
<p>Comparison of the present correlation with the numerical data.</p>
Full article ">Figure 14
<p>Comparison of the present correlation with experimental data: (<b>a</b>) D18, (<b>b</b>) D22, and (<b>c</b>) D28.</p>
Full article ">
38 pages, 479 KiB  
Article
The Design of Global Correlation Quantifiers and Continuous Notions of Statistical Sufficiency
by Nicholas Carrara and Kevin Vanslette
Entropy 2020, 22(3), 357; https://doi.org/10.3390/e22030357 - 19 Mar 2020
Cited by 2 | Viewed by 2388
Abstract
Using first principles from inference, we design a set of functionals for the purposes of ranking joint probability distributions with respect to their correlations. Starting with a general functional, we impose its desired behavior through the Principle of Constant Correlations (PCC), which constrains [...] Read more.
Using first principles from inference, we design a set of functionals for the purposes of ranking joint probability distributions with respect to their correlations. Starting with a general functional, we impose its desired behavior through the Principle of Constant Correlations (PCC), which constrains the correlation functional to behave in a consistent way under statistically independent inferential transformations. The PCC guides us in choosing the appropriate design criteria for constructing the desired functionals. Since the derivations depend on a choice of partitioning the variable space into n disjoint subspaces, the general functional we design is the n-partite information (NPI), of which the total correlation and mutual information are special cases. Thus, these functionals are found to be uniquely capable of determining whether a certain class of inferential transformations, ρ ρ , preserve, destroy or create correlations. This provides conceptual clarity by ruling out other possible global correlation quantifiers. Finally, the derivation and results allow us to quantify non-binary notions of statistical sufficiency. Our results express what percentage of the correlations are preserved under a given inferential transformation or variable mapping. Full article
18 pages, 2404 KiB  
Article
Impact of General Anesthesia Guided by State Entropy (SE) and Response Entropy (RE) on Perioperative Stability in Elective Laparoscopic Cholecystectomy Patients—A Prospective Observational Randomized Monocentric Study
by Anca Raluca Dinu, Alexandru Florin Rogobete, Sonia Elena Popovici, Ovidiu Horea Bedreag, Marius Papurica, Corina Maria Dumbuleu, Raluca Ramona Velovan, Daiana Toma, Corina Maria Georgescu, Lavinia Ioana Trache, Claudiu Barsac, Loredana Luca, Bettina Buzzi, Andra Maghiar, Mihai Alexandru Sandesc, Samir Rimawi, Madalin Marian Vaduva, Lavinia Melania Bratu, Paul Manuel Luminosu and Dorel Sandesc
Entropy 2020, 22(3), 356; https://doi.org/10.3390/e22030356 - 19 Mar 2020
Cited by 10 | Viewed by 6163
Abstract
Laparoscopic cholecystectomy is one of the most frequently performed interventions in general surgery departments. Some of the most important aims in achieving perioperative stability in these patients is diminishing the impact of general anesthesia on the hemodynamic stability and the optimization of anesthetic [...] Read more.
Laparoscopic cholecystectomy is one of the most frequently performed interventions in general surgery departments. Some of the most important aims in achieving perioperative stability in these patients is diminishing the impact of general anesthesia on the hemodynamic stability and the optimization of anesthetic drug doses based on the individual clinical profile of each patient. The objective of this study is the evaluation of the impact, as monitored through entropy (both state entropy (SE) and response entropy (RE)), that the depth of anesthesia has on the hemodynamic stability, as well as the doses of volatile anesthetic. A prospective, observational, randomized, and monocentric study was carried out between January and December 2019 in the Clinic of Anesthesia and Intensive Care of the “Pius Brînzeu” Emergency County Hospital in Timișoara, Romania. The patients included in the study were divided in two study groups: patients in Group A (target group) received multimodal monitoring, which included monitoring of standard parameters and of entropy (SE and RE); while the patients in Group B (control group) only received standard monitoring. The anesthetic dose in group A was optimized to achieve a target entropy of 40–60. A total of 68 patients met the inclusion criteria and were allocated to one of the two study groups: group A (N = 43) or group B (N = 25). There were no statistically significant differences identified between the two groups for both demographical and clinical characteristics (p > 0.05). Statistically significant differences were identified for the number of hypotensive episodes (p = 0.011, 95% CI: [0.1851, 0.7042]) and for the number of episodes of bradycardia (p < 0.0001, 95% CI: [0.3296, 0.7923]). Moreover, there was a significant difference in the Sevoflurane consumption between the two study groups (p = 0.0498, 95% CI: [−0.3942, 0.9047]). The implementation of the multimodal monitoring protocol, including the standard parameters and the measurement of entropy for determining the depth of anesthesia (SE and RE) led to a considerable improvement in perioperative hemodynamic stability. Furthermore, optimizing the doses of anesthetic drugs based on the individual clinical profile of each patient led to a considerable decrease in drug consumption, as well as to a lower incidence of hemodynamic side-effects. Full article
Show Figures

Figure 1

Figure 1
<p>Study flowchart and data processing methodology.</p>
Full article ">Figure 2
<p>State Entropy (SE) and response entropy (RE) expression. (<b>A</b>) correlation between SE and RE values (p = 0.7620, 95% CI: [0.9812, 0.9982], r = 0.9942, R<sup>2</sup> = 0.9884); (<b>B</b>) evolution of SE and RE in time; (<b>C</b>) expression of SE during general anesthesia (black dots, individual subject measurements; red lines, standard deviation; black lines, means); (<b>D</b>) expression and allocation of SE in time during general anesthesia (black dots, individual subject measurements; red lines, standard deviation; black lines, means).</p>
Full article ">Figure 3
<p>Statistical correlations for minimum alveolar concentration (MAC) and SE/RE. (<b>A</b>) correlations for SE and MAC (r = −0.9583, r<sup>2</sup> = 0.9184, 95% CI: [−0.9878, −0.8630]); (<b>B</b>) correlations for RE and MAC (r = −0.9519, r<sup>2</sup> = 0.9043, 95% CI: [−0.9855, −0.8402]); (<b>C</b>) Bland–Altman analysis of MAC values; and (<b>D</b>) MAC values over time.</p>
Full article ">Figure 4
<p>Statistical analysis for sevoflurane consumption. (<b>Left</b>) scatter plot with bar (mean with SD) for total consumption of Sevoflurane (mL), (group A: lower 95% CI of mean 122.5, upper 95% CI of mean 165.6, coefficient of variation 47.90%; group B: lower 95% CI of mean 160.9, upper 95% CI of mean 210.7, coefficient of variation 32.47%); (<b>Right</b>) scatter plot with bar (mean with SD) for Sevoflurane consumption/minute (mL/min), (group A: lower 95% CI of mean: 1.748, upper 95% CI of mean: 2.634, variation coefficient: 65.73%; group B: lower 95% CI: 2.040, upper 95% CI: 2.853, variation coefficient: 40.26%). The mean difference between the two groups was 0.2553 ± 0.3253 and the 95% CI was [−0.3942, 0.9047].</p>
Full article ">Figure 5
<p>Statistical and graphical analyses of perioperative hemodynamic changes. (<b>A</b>) number of bradycardia episodes; (<b>B</b>) number of tachycardia episodes; (<b>C</b>) number of hypotensive episodes; and (<b>D</b>) number of hypertensive episodes. Regarding the statistical analysis of the intraoperative hemodynamic events, significant statistical differences can be observed in the number of bradycardia (A) events (respectively, in the number of hypotension (C) events), where there was a decrease in the incidence for patients in group A. In contrast, regarding the number of tachycardia (B) events (respectively, of the number of hypertension (D) events), no statistically significant differences were observed between the two groups.</p>
Full article ">Figure 6
<p>(<b>A, B</b>) Statistical analysis of mean differences for heart rate (HR, bpm); (<b>C, D</b>) Statistical analysis of mean differences for blood pressure (SAB, mmHg).</p>
Full article ">
16 pages, 15166 KiB  
Article
Improved Practical Vulnerability Analysis of Mouse Data According to Offensive Security based on Machine Learning in Image-Based User Authentication
by Kyungroul Lee and Sun-Young Lee
Entropy 2020, 22(3), 355; https://doi.org/10.3390/e22030355 - 18 Mar 2020
Cited by 2 | Viewed by 3054
Abstract
The objective of this study was to verify the feasibility of mouse data exposure by deriving features to improve the accuracy of a mouse data attack technique using machine learning models. To improve the accuracy, the feature appearing between the mouse coordinates input [...] Read more.
The objective of this study was to verify the feasibility of mouse data exposure by deriving features to improve the accuracy of a mouse data attack technique using machine learning models. To improve the accuracy, the feature appearing between the mouse coordinates input from the user was analyzed, which is defined as a feature for machine learning models to derive a method of improving the accuracy. As a result, we found a feature where the distance between the coordinates is concentrated in a specific range. We verified that the mouse data is apt to being stolen more accurately when the distance is used as a feature. An accuracy of over 99% was achieved, which means that the proposed method almost completely classifies the mouse data input from the user and the mouse data generated by the defender. Full article
(This article belongs to the Special Issue Statistical Inference from High Dimensional Data)
Show Figures

Figure 1

Figure 1
<p>Mouse data transfer process.</p>
Full article ">Figure 2
<p>Distribution of the coordinates according to the elapsed time.</p>
Full article ">Figure 3
<p>Distribution of the coordinates according to the X coordinate.</p>
Full article ">Figure 4
<p>Distribution of the coordinates according to the Y coordinate.</p>
Full article ">Figure 5
<p>Distribution of the coordinates according to the distance between previous X coordinates and current X coordinates.</p>
Full article ">Figure 6
<p>Distribution of the coordinates according to the distance between previous Y coordinates and current Y coordinates.</p>
Full article ">Figure 7
<p>Performance evaluations of accuracy, precision, recall, F1-score, and AUC for datasets 1-1 to 1-4.</p>
Full article ">Figure 8
<p>Performance evaluations of accuracy, precision, recall, F1-score, and AUC for datasets 2-1 to 2-4.</p>
Full article ">Figure 9
<p>Performance evaluation of the proposed and existing methods in the training set, validation set, and test set. (<b>a</b>) dataset 1-1 and 1-2; (<b>b</b>) dataset 1-3 and 1-4; (<b>c</b>) dataset 2-1 and 2-2; (<b>d</b>) dataset 2-3 and 2-4.</p>
Full article ">Figure 9 Cont.
<p>Performance evaluation of the proposed and existing methods in the training set, validation set, and test set. (<b>a</b>) dataset 1-1 and 1-2; (<b>b</b>) dataset 1-3 and 1-4; (<b>c</b>) dataset 2-1 and 2-2; (<b>d</b>) dataset 2-3 and 2-4.</p>
Full article ">Figure 10
<p>Performance evaluation of the proposed and existing methods with respect to accuracy, precision, recall, F1-score, and AUC. (<b>a</b>) dataset 1-1 and 1-2; (<b>b</b>) dataset 1-3 and 1-4; (<b>c</b>) dataset 2-1 and 2-2; (<b>d</b>) dataset 2-3 and 2-4.</p>
Full article ">Figure 10 Cont.
<p>Performance evaluation of the proposed and existing methods with respect to accuracy, precision, recall, F1-score, and AUC. (<b>a</b>) dataset 1-1 and 1-2; (<b>b</b>) dataset 1-3 and 1-4; (<b>c</b>) dataset 2-1 and 2-2; (<b>d</b>) dataset 2-3 and 2-4.</p>
Full article ">Figure 10 Cont.
<p>Performance evaluation of the proposed and existing methods with respect to accuracy, precision, recall, F1-score, and AUC. (<b>a</b>) dataset 1-1 and 1-2; (<b>b</b>) dataset 1-3 and 1-4; (<b>c</b>) dataset 2-1 and 2-2; (<b>d</b>) dataset 2-3 and 2-4.</p>
Full article ">
17 pages, 4122 KiB  
Article
Radiative MHD Nanofluid Flow over a Moving Thin Needle with Entropy Generation in a Porous Medium with Dust Particles and Hall Current
by Iskander Tlili, Muhammad Ramzan, Seifedine Kadry, Hyun-Woo Kim and Yunyoung Nam
Entropy 2020, 22(3), 354; https://doi.org/10.3390/e22030354 - 18 Mar 2020
Cited by 37 | Viewed by 3743
Abstract
This paper investigated the behavior of the two-dimensional magnetohydrodynamics (MHD) nanofluid flow of water-based suspended carbon nanotubes (CNTs) with entropy generation and nonlinear thermal radiation in a Darcy–Forchheimer porous medium over a moving horizontal thin needle. The study also incorporated the effects of [...] Read more.
This paper investigated the behavior of the two-dimensional magnetohydrodynamics (MHD) nanofluid flow of water-based suspended carbon nanotubes (CNTs) with entropy generation and nonlinear thermal radiation in a Darcy–Forchheimer porous medium over a moving horizontal thin needle. The study also incorporated the effects of Hall current, magnetohydrodynamics, and viscous dissipation on dust particles. The said flow model was described using high order partial differential equations. An appropriate set of transformations was used to reduce the order of these equations. The reduced system was then solved by using a MATLAB tool bvp4c. The results obtained were compared with the existing literature, and excellent harmony was achieved in this regard. The results were presented using graphs and tables with coherent discussion. It was comprehended that Hall current parameter intensified the velocity profiles for both CNTs. Furthermore, it was perceived that the Bejan number boosted for higher values of Darcy–Forchheimer number. Full article
(This article belongs to the Special Issue Thermal Radiation and Entropy Analysis)
Show Figures

Figure 1

Figure 1
<p>The physical design of the flow problem.</p>
Full article ">Figure 2
<p>Impact of <math display="inline"><semantics> <mi>a</mi> </semantics></math> on (<b>a</b>) nanofluid velocity, (<b>b</b>) the velocity of the dust phase, (<b>c</b>) nanofluid temperature, and (<b>d</b>) temperature of the dust phase.</p>
Full article ">Figure 3
<p>Impact of <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>r</mi> </msub> </mrow> </semantics></math> on (<b>a</b>) nanofluid temperature and (<b>b</b>) temperature of the dust phase.</p>
Full article ">Figure 4
<p>Impact of <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>r</mi> </msub> </mrow> </semantics></math> on (<b>a</b>) nanofluid temperature and (<b>b</b>) temperature of the dust phase.</p>
Full article ">Figure 5
<p>Impact of <math display="inline"><semantics> <mrow> <msub> <mi>F</mi> <mi>r</mi> </msub> </mrow> </semantics></math> on (<b>a</b>) nanofluid velocity and (<b>b</b>) the velocity of the dust phase.</p>
Full article ">Figure 6
<p>Impact of <math display="inline"><semantics> <mi>M</mi> </semantics></math> on (<b>a</b>) nanofluid velocity, (<b>b</b>) the velocity of the dust phase, (<b>c</b>) nanofluid temperature, and (<b>d</b>) temperature of the dust phase.</p>
Full article ">Figure 7
<p>Impact of <math display="inline"><semantics> <mi>m</mi> </semantics></math> on (<b>a</b>) nanofluid velocity and (<b>b</b>) the velocity of the dust phase.</p>
Full article ">Figure 8
<p>Impact of <math display="inline"><semantics> <mi>m</mi> </semantics></math> on (<b>a</b>) entropy generation number and (<b>b</b>) Bejan number.</p>
Full article ">Figure 9
<p>Impact of <math display="inline"><semantics> <mrow> <msub> <mi>F</mi> <mi>r</mi> </msub> </mrow> </semantics></math> on (<b>a</b>) entropy generation number and (<b>b</b>) Bejan number.</p>
Full article ">Figure 10
<p>Impact of <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>r</mi> </msub> </mrow> </semantics></math> on (<b>a</b>) entropy generation number and (<b>b</b>) Bejan number.</p>
Full article ">
22 pages, 17436 KiB  
Article
Gaussian Curvature Entropy for Curved Surface Shape Generation
by Akihiro Okano, Taishi Matsumoto and Takeo Kato
Entropy 2020, 22(3), 353; https://doi.org/10.3390/e22030353 - 18 Mar 2020
Cited by 10 | Viewed by 5108
Abstract
The overall shape features that emerge from combinations of shape elements, such as “complexity” and “order”, are important in designing shapes of industrial products. However, controlling the features of shapes is difficult and depends on the experience and intuition of designers. Among these [...] Read more.
The overall shape features that emerge from combinations of shape elements, such as “complexity” and “order”, are important in designing shapes of industrial products. However, controlling the features of shapes is difficult and depends on the experience and intuition of designers. Among these features, “complexity” is said to have an influence on the “beauty” and “preference” of shapes. This research proposed a Gaussian curvature entropy as a “complexity” index of a curved surface shape. The proposed index is calculated based on Gaussian curvature, which is obtained by the sampling and quantization of a curved surface shape and validated by the sensory evaluation experiment while using two types of sample shapes. The result indicates the correspondence of the index to perceived “complexity” (the determination coefficient is greater than 0.8). Additionally, this research constructed a shape generation method that was based on the index as a car design supporting apparatus, in which the designers can refer many shapes generated by controlling “complexity”. The applicability of the proposed method was confirmed by the experiment while using the generated shapes. Full article
(This article belongs to the Special Issue Intelligent Tools and Applications in Engineering and Mathematics)
Show Figures

Figure 1

Figure 1
<p>Principal curvatures <span class="html-italic">κ</span><sub>max</sub> and <span class="html-italic">κ</span><sub>min</sub> at point <span class="html-italic">p</span>.</p>
Full article ">Figure 2
<p>Calculation of Gaussian curvature entropy: (<b>a</b>) Division of curved surface and calculation of dimensionless Gaussian curvature; (<b>b</b>) Quantization of dimensionless Gaussian curvature; and, (<b>c</b>) Calculation of occurrence probability and transition probability.</p>
Full article ">Figure 2 Cont.
<p>Calculation of Gaussian curvature entropy: (<b>a</b>) Division of curved surface and calculation of dimensionless Gaussian curvature; (<b>b</b>) Quantization of dimensionless Gaussian curvature; and, (<b>c</b>) Calculation of occurrence probability and transition probability.</p>
Full article ">Figure 3
<p>Calculation of Gaussian curvature at sample point <span class="html-italic">v</span>.</p>
Full article ">Figure 4
<p>Quantization of dimensionless Gaussian curvature at each point: (<b>a</b>) {<span class="html-italic">E</span>, <span class="html-italic">V</span>} = {30, 3}; and, (<b>b</b>) {<span class="html-italic">E</span>, <span class="html-italic">V</span>} = {15, 15}.</p>
Full article ">Figure 5
<p>Quantization of dimensionless Gaussian curvature on shapes without concavities when {<span class="html-italic">E</span>, <span class="html-italic">V</span>} = {20, 15}: (<b>a</b>) sphere; (<b>b</b>) distorted sphere.</p>
Full article ">Figure 6
<p>Generation of sample shapes A: (<b>a</b>) Curved shapes to extrude; (<b>b</b>) Sample shapes A.</p>
Full article ">Figure 7
<p>Generation of sample shapes B: (<b>a</b>) Curved shapes to rotate; (<b>b</b>) Sample shapes B.</p>
Full article ">Figure 8
<p>Relationship between parameters (<span class="html-italic">E</span> and <span class="html-italic">V</span>) and coefficient of determination <span class="html-italic">R</span><sup>2</sup> in sample shapes A.</p>
Full article ">Figure 9
<p>Relationship between Gaussian curvature entropy <span class="html-italic">H</span><sub>G</sub> and sensory evaluation values about “complexity” in sample shapes A.</p>
Full article ">Figure 10
<p>Relationship between parameters (<span class="html-italic">E</span> and <span class="html-italic">V</span>) and coefficient of determination <span class="html-italic">R</span><sup>2</sup> in sample shapes B.</p>
Full article ">Figure 11
<p>Relationship between Gaussian curvature entropy <span class="html-italic">H</span><sub>G</sub> and sensory evaluation values about “complexity” in sample shapes B.</p>
Full article ">Figure 12
<p>Shapes whose values of <span class="html-italic">H</span><sub>G</sub> differs while their sensory evaluation values are close: (<b>a</b>) shape 8; (<b>b</b>) shape 14.</p>
Full article ">Figure 13
<p>Examples of quantization when {<span class="html-italic">E</span>, <span class="html-italic">V</span>} = {200, 11}: (<b>a</b>) shape 11; (<b>b</b>) shape 13; and, (<b>c</b>) shape 6.</p>
Full article ">Figure 14
<p>Difference in quantization of dimensionless Gaussian curvature: (<b>a</b>) <span class="html-italic">V</span> = 3; (<b>b</b>) <span class="html-italic">V</span> = 4.</p>
Full article ">Figure 15
<p>Shapes whose relationship between the sensory evaluation values and <span class="html-italic">H</span><sub>G</sub> is reversed: (<b>a</b>) shape 9; (<b>b</b>) shape 11.</p>
Full article ">Figure 16
<p>Examples of quantization when {<span class="html-italic">E</span>, <span class="html-italic">V</span>} = {50, 19}: (<b>a</b>) shape 8; (<b>b</b>) shape 5; and, (<b>c</b>) shape 6.</p>
Full article ">Figure 17
<p>Shapes whose sensory evaluation values correspond to Gaussian curvature entropy rather than total absolute Gaussian curvature: (<b>a</b>) shape 10; (<b>b</b>) shape 13 in samples shapes B.</p>
Full article ">Figure 18
<p>Shapes whose sensory evaluation values correspond to total absolute Gaussian curvature rather than Gaussian curvature entropy: (<b>a</b>) shape 4; (<b>b</b>) shape 10 in sample shapes A.</p>
Full article ">Figure 19
<p>Shape generation method based on Gaussian curvature entropy: (<b>a</b>) Setting of an initial shape, <span class="html-italic">H</span><sub>G,target</sub> and <span class="html-italic">f</span><sub>u</sub>; (<b>b</b>) Generation of position vectors; and, (<b>c</b>) Generation and output of shapes.</p>
Full article ">Figure 20
<p>Initial shape of shape generation.</p>
Full article ">Figure 21
<p>Examples of generated shapes: (<b>a</b>) <span class="html-italic">H</span><sub>G</sub> = 0.30; (<b>b</b>) <span class="html-italic">H</span><sub>G</sub> = 0.46; and, (<b>c</b>) <span class="html-italic">H</span><sub>G</sub> = 0.62.</p>
Full article ">Figure 22
<p>Displaying method of generated shapes.</p>
Full article ">Figure 23
<p>Relationship between Gaussian curvature entropy <span class="html-italic">H</span><sub>G</sub> and sensory evaluation values about “complexity” in generated shapes.</p>
Full article ">Figure 24
<p>Quantization of dimensionless Gaussian curvature in shapes deviating from approximate curve: (<b>a</b>) shape 10; (<b>b</b>) shape 11; and, (<b>c</b>) shape 15.</p>
Full article ">Figure 25
<p>Relationship between Gaussian curvature entropy <span class="html-italic">H</span><sub>G</sub> considering front part of shapes and sensory evaluation values about “complexity” in generated shapes.</p>
Full article ">Figure A1
<p>Relationship between parameters (<span class="html-italic">E</span> and <span class="html-italic">V</span>) and coefficient of determination <span class="html-italic">R</span><sup>2</sup> in sample shapes A during cross-validation; (<b>a</b>) Group A1 and A2; (<b>b</b>) Group A1 and A3; (<b>c</b>) Group A2 and A3.</p>
Full article ">Figure A1 Cont.
<p>Relationship between parameters (<span class="html-italic">E</span> and <span class="html-italic">V</span>) and coefficient of determination <span class="html-italic">R</span><sup>2</sup> in sample shapes A during cross-validation; (<b>a</b>) Group A1 and A2; (<b>b</b>) Group A1 and A3; (<b>c</b>) Group A2 and A3.</p>
Full article ">Figure A2
<p>Relationship between parameters (<span class="html-italic">E</span> and <span class="html-italic">V</span>) and coefficient of determination <span class="html-italic">R</span><sup>2</sup> in sample shapes A during cross-validation; (<b>a</b>) Group B1 and B2; (<b>b</b>) Group B1 and B3; (<b>c</b>) Group B2 and B3.</p>
Full article ">Figure A2 Cont.
<p>Relationship between parameters (<span class="html-italic">E</span> and <span class="html-italic">V</span>) and coefficient of determination <span class="html-italic">R</span><sup>2</sup> in sample shapes A during cross-validation; (<b>a</b>) Group B1 and B2; (<b>b</b>) Group B1 and B3; (<b>c</b>) Group B2 and B3.</p>
Full article ">
8 pages, 263 KiB  
Article
Stability Analysis of the Explicit Difference Scheme for Richards Equation
by Fengnan Liu, Yasuhide Fukumoto and Xiaopeng Zhao
Entropy 2020, 22(3), 352; https://doi.org/10.3390/e22030352 - 18 Mar 2020
Cited by 4 | Viewed by 3560
Abstract
A stable explicit difference scheme, which is based on forward Euler format, is proposed for the Richards equation. To avoid the degeneracy of the Richards equation, we add a perturbation to the functional coefficient of the parabolic term. In addition, we introduce an [...] Read more.
A stable explicit difference scheme, which is based on forward Euler format, is proposed for the Richards equation. To avoid the degeneracy of the Richards equation, we add a perturbation to the functional coefficient of the parabolic term. In addition, we introduce an extra term in the difference scheme which is used to relax the time step restriction for improving the stability condition. With the augmented terms, we prove the stability using the induction method. Numerical experiments show the validity and the accuracy of the scheme, along with its efficiency. Full article
(This article belongs to the Special Issue Applications of Nonlinear Diffusion Equations)
Show Figures

Figure 1

Figure 1
<p>Variation of the pressure head with depth.</p>
Full article ">Figure 2
<p>Relationship between <math display="inline"><semantics> <msub> <mi>ϵ</mi> <mn>1</mn> </msub> </semantics></math> and errors.</p>
Full article ">
15 pages, 462 KiB  
Article
Weighted Mutual Information for Aggregated Kernel Clustering
by Nezamoddin N. Kachouie and Meshal Shutaywi
Entropy 2020, 22(3), 351; https://doi.org/10.3390/e22030351 - 18 Mar 2020
Cited by 7 | Viewed by 2963
Abstract
Background: A common task in machine learning is clustering data into different groups based on similarities. Clustering methods can be divided in two groups: linear and nonlinear. A commonly used linear clustering method is K-means. Its extension, kernel K-means, is a non-linear technique [...] Read more.
Background: A common task in machine learning is clustering data into different groups based on similarities. Clustering methods can be divided in two groups: linear and nonlinear. A commonly used linear clustering method is K-means. Its extension, kernel K-means, is a non-linear technique that utilizes a kernel function to project the data to a higher dimensional space. The projected data will then be clustered in different groups. Different kernels do not perform similarly when they are applied to different datasets. Methods: A kernel function might be relevant for one application but perform poorly to project data for another application. In turn choosing the right kernel for an arbitrary dataset is a challenging task. To address this challenge, a potential approach is aggregating the clustering results to obtain an impartial clustering result regardless of the selected kernel function. To this end, the main challenge is how to aggregate the clustering results. A potential solution is to combine the clustering results using a weight function. In this work, we introduce Weighted Mutual Information (WMI) for calculating the weights for different clustering methods based on their performance to combine the results. The performance of each method is evaluated using a training set with known labels. Results: We applied the proposed Weighted Mutual Information to four data sets that cannot be linearly separated. We also tested the method in different noise conditions. Conclusions: Our results show that the proposed Weighted Mutual Information method is impartial, does not rely on a single kernel, and performs better than each individual kernel specially in high noise. Full article
Show Figures

Figure 1

Figure 1
<p>NMI (Normalized Mutual Information) score vs. true positive rate (percentage of elements clustered in right groups).</p>
Full article ">Figure 2
<p>Clustering of the labeled training set for two inner circles (gray dots show entire data) obtained by Gaussian, polynomial, and tangent kernels.</p>
Full article ">Figure 3
<p>Clustering results obtained by Gaussian, polynomial, and tangent kernels along with aggregated results obtained by majority voting and WMI (Weighted Mutual Information) kernel clustering.</p>
Full article ">Figure 4
<p>Clustering results obtained by Gaussian, polynomial, and tangent kernels along with aggregated results obtained by majority voting and WMI kernel clustering for inner circles corrupted with high noise.</p>
Full article ">Figure 5
<p>Clustering results obtained by Gaussian, polynomial, and tangent kernels along with aggregated results obtained by majority voting and WMI kernel clustering for Two Moons.</p>
Full article ">Figure 6
<p>Clustering results obtained by Gaussian, polynomial, and tangent kernels along with aggregated results obtained by majority voting and WMI kernel clustering for Two Moons corrupted with high noise.</p>
Full article ">Figure 7
<p>Clustering results obtained by Gaussian, polynomial, and tangent kernels along with aggregated results obtained by majority voting and WMI kernel clustering for Iris data.</p>
Full article ">Figure 8
<p>Clustering of the labeled training set for chromosome one (gray dots show entire data) obtained by Gaussian, polynomial, and tangent kernels.</p>
Full article ">Figure 9
<p>Clustering results obtained by Gaussian, polynomial, and tangent kernels along with aggregated results obtained by majority voting and WMI kernel clustering for chromosome one.</p>
Full article ">
15 pages, 519 KiB  
Article
Invariant-Based Inverse Engineering for Fast and Robust Load Transport in a Double Pendulum Bridge Crane
by Ion Lizuain, Ander Tobalina, Alvaro Rodriguez-Prieto and Juan Gonzalo Muga
Entropy 2020, 22(3), 350; https://doi.org/10.3390/e22030350 - 18 Mar 2020
Cited by 3 | Viewed by 3461
Abstract
We set a shortcut-to-adiabaticity strategy to design the trolley motion in a double-pendulum bridge crane. The trajectories found guarantee payload transport without residual excitation regardless of the initial conditions within the small oscillations regime. The results are compared with exact dynamics to set [...] Read more.
We set a shortcut-to-adiabaticity strategy to design the trolley motion in a double-pendulum bridge crane. The trajectories found guarantee payload transport without residual excitation regardless of the initial conditions within the small oscillations regime. The results are compared with exact dynamics to set the working domain of the approach. The method is free from instabilities due to boundary effects or to resonances with the two natural frequencies. Full article
(This article belongs to the Special Issue Shortcuts to Adiabaticity)
Show Figures

Figure 1

Figure 1
<p>Double pendulum overhead crane scheme and relevant physical parameters.</p>
Full article ">Figure 2
<p>(Color online) Trolley trajectories <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> and velocities <math display="inline"><semantics> <mrow> <mi>v</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <mover accent="true"> <mi>x</mi> <mo>˙</mo> </mover> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for different final times: <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>f</mi> </msub> <mo>=</mo> <mn>2</mn> <mi>s</mi> </mrow> </semantics></math> (blue-dashed line), <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>f</mi> </msub> <mo>=</mo> <mn>4</mn> <mi>s</mi> </mrow> </semantics></math> (green-dot-dashed line), <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>f</mi> </msub> <mo>=</mo> <mn>8</mn> <mi>s</mi> </mrow> </semantics></math> (red-dotted line). Compare to the “long time behaviour” in (a) of (<a href="#FD38-entropy-22-00350" class="html-disp-formula">38</a>) (black-solid line). Other parameters are: <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math> m.</p>
Full article ">Figure 3
<p>(Color online) Time evolution of the suspension angles for a transport of <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math> m in a time <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>f</mi> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> s. We have numerically integrated the exact dynamical equations using the exact Lagrangian (<a href="#FD3-entropy-22-00350" class="html-disp-formula">3</a>) with different initial conditions: (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>5</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>5</mn> <mo>°</mo> </msup> </mrow> </semantics></math>; and (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>5</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>5</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, with <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>θ</mi> <mo>˙</mo> </mover> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mover accent="true"> <mi>θ</mi> <mo>˙</mo> </mover> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> in all cases. In the scale of the figure the results using the approximate Lagrangian (<a href="#FD5-entropy-22-00350" class="html-disp-formula">5</a>) or the exact one are indistinguishable. Other parameters are: <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math> m.</p>
Full article ">Figure 4
<p>(Color online) Maximum swing angles during the process as a function of the duration <math display="inline"><semantics> <msub> <mi>t</mi> <mi>f</mi> </msub> </semantics></math> (red dotted and blue dashed lines). For very rapid operations (small <math display="inline"><semantics> <msub> <mi>t</mi> <mi>f</mi> </msub> </semantics></math>), larger angles are involved and the harmonic approximation breaks down. Fictitious angle <math display="inline"><semantics> <msub> <mi>θ</mi> <mi>f</mi> </msub> </semantics></math> (black-solid line), which basically is a measure of the final excitation energy, see the main text, as a function of <math display="inline"><semantics> <msub> <mi>t</mi> <mi>f</mi> </msub> </semantics></math>. In the harmonic approximation this angle is zero by construction (<math display="inline"><semantics> <msub> <mi>θ</mi> <mrow> <mi>f</mi> <mo>,</mo> <mi>h</mi> </mrow> </msub> </semantics></math>, green-dashed-dotted line). System assumed initially in equilibrium. Other parameters are: <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math> m. The natural periods of the modes are <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>2</mn> <mi>π</mi> <mo>/</mo> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>11.048</mn> </mrow> </semantics></math> s and <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>2</mn> <mi>π</mi> <mo>/</mo> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>3.298</mn> </mrow> </semantics></math> s.</p>
Full article ">Figure 5
<p>(Color online) Difference between final and initial energy measured by the modulus of the fictitious angle <math display="inline"><semantics> <msub> <mi>θ</mi> <mi>f</mi> </msub> </semantics></math> (in °) as a function of the deviations from equilibrium configuration of either <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> or <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> after solving the exact dynamics with Lagrangian (<a href="#FD2-entropy-22-00350" class="html-disp-formula">2</a>). (<b>a</b>) Final fictitious angle for the inverse-engineered trolley trajectory (<a href="#FD37-entropy-22-00350" class="html-disp-formula">37</a>). (<b>b</b>) Final fictitious angle for the postulated cubic trajectory (<a href="#FD40-entropy-22-00350" class="html-disp-formula">40</a>). Other parameters are: <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math> m and <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>f</mi> </msub> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> s. The system is assummed initially at rest, <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>θ</mi> <mo>˙</mo> </mover> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mover accent="true"> <mi>θ</mi> <mo>˙</mo> </mover> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>(Color online) Comparison of two transport protocols, the inverse engineered trolley trajectory (<a href="#FD37-entropy-22-00350" class="html-disp-formula">37</a>) (red) and the cubic trajectory (<a href="#FD40-entropy-22-00350" class="html-disp-formula">40</a>) (blue). The dotted line is for segments where the maximal trolley velocity is larger than 2 m/s, whereas in solid line segments the maximal velocity is below that value. (<b>a</b>) Maximum trolley velocity during the process and (<b>b</b>) excitation at final time measured by the fictitious angle <math display="inline"><semantics> <msub> <mi>θ</mi> <mi>f</mi> </msub> </semantics></math>. Rest of parameters: <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>150</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math> kg, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>40</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> m and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>40</mn> </mrow> </semantics></math> m. System initially at equilibrium. The natural periods of the modes are <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>2</mn> <mi>π</mi> <mo>/</mo> <msub> <mi>ω</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>13.376</mn> </mrow> </semantics></math> s and <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>2</mn> <mi>π</mi> <mo>/</mo> <msub> <mi>ω</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1.538</mn> </mrow> </semantics></math> s.</p>
Full article ">
21 pages, 857 KiB  
Article
A Note on Wavelet-Based Estimator of the Hurst Parameter
by Liang Wu
Entropy 2020, 22(3), 349; https://doi.org/10.3390/e22030349 - 18 Mar 2020
Cited by 3 | Viewed by 2750
Abstract
The signals in numerous fields usually have scaling behaviors (long-range dependence and self-similarity) which is characterized by the Hurst parameter H. Fractal Brownian motion (FBM) plays an important role in modeling signals with self-similarity and long-range dependence. Wavelet analysis is a common [...] Read more.
The signals in numerous fields usually have scaling behaviors (long-range dependence and self-similarity) which is characterized by the Hurst parameter H. Fractal Brownian motion (FBM) plays an important role in modeling signals with self-similarity and long-range dependence. Wavelet analysis is a common method for signal processing, and has been used for estimation of Hurst parameter. This paper conducts a detailed numerical simulation study in the case of FBM on the selection of parameters and the empirical bias in the wavelet-based estimator which have not been studied comprehensively in previous studies, especially for the empirical bias. The results show that the empirical bias is due to the initialization errors caused by discrete sampling, and is not related to simulation methods. When choosing an appropriate orthogonal compact supported wavelet, the empirical bias is almost not related to the inaccurate bias correction caused by correlations of wavelet coefficients. The latter two causes are studied via comparison of estimators and comparison of simulation methods. These results could be a reference for future studies and applications in the scaling behavior of signals. Some preliminary results of this study have provided a reference for my previous studies. Full article
Show Figures

Figure 1

Figure 1
<p>The Bias, Std, RMSE for estimators: <math display="inline"><semantics> <msub> <mi>j</mi> <mn>1</mn> </msub> </semantics></math> is the lower bound of octaves <span class="html-italic">j</span>s. Std is the standard deviation, Bias <math display="inline"><semantics> <mrow> <mo>=</mo> <mi mathvariant="double-struck">E</mi> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> <mi>H</mi> </mrow> </semantics></math>, RMSE is the square root of MSE. The values of Std, Bias, and RMSE are the estimated versions of those for 1000 independent copies of FBM with length <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <msup> <mn>2</mn> <mn>18</mn> </msup> </mrow> </semantics></math>. The used wavelet is the Daubechies wavelet with <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> vanishing moments.</p>
Full article ">Figure 2
<p>The Bias, Std, RMSE for estimators: <span class="html-italic">n</span> is the data length. Std is the standard deviation, Bias <math display="inline"><semantics> <mrow> <mo>=</mo> <mi mathvariant="double-struck">E</mi> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> <mi>H</mi> </mrow> </semantics></math>, RMSE is the square root of MSE. The values of Std, Bias, and RMSE are the estimated versions of those for 1000 independent copies of FBM with length <span class="html-italic">n</span>. The lower bound of octaves <span class="html-italic">j</span>s is chosen <math display="inline"><semantics> <mrow> <msub> <mi>j</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. The used wavelet is the Daubechies wavelet with <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> vanishing moments.</p>
Full article ">Figure 3
<p>The Bias, Std, RMSE for estimators: <span class="html-italic">N</span> is the number of vanishing moments of Daubechies wavelet. Std is the standard deviation, Bias <math display="inline"><semantics> <mrow> <mo>=</mo> <mi mathvariant="double-struck">E</mi> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> <mi>H</mi> </mrow> </semantics></math>, RMSE is the square root of MSE. The values of Std, Bias and RMSE are the estimated versions of those for 1000 independent copies of FBM with length <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <msup> <mn>2</mn> <mn>18</mn> </msup> </mrow> </semantics></math>. The lower bound of octaves <span class="html-italic">j</span>s is chosen <math display="inline"><semantics> <mrow> <msub> <mi>j</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>The Bias, Std, RMSE for estimators: db3 stands for Daubechies wavelet with three vanishing moments, sym4 stands for Symlets wavelet with four vanishing moments, dmey stands for discrete Meyer wavelet, bior3.1 stands for biorthogonal spline wavelets with orders <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> (vanishing moments) and <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>d</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>. The values of Std, Bias and RMSE are the estimated versions of those for 1000 independent copies of FBM with length <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <msup> <mn>2</mn> <mn>18</mn> </msup> </mrow> </semantics></math>. The lower bound of octaves <span class="html-italic">j</span>s is chosen <math display="inline"><semantics> <mrow> <msub> <mi>j</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>The Bias and Std for estimators: M1 denotes the first estimator, M2 denotes the second estimator. Std is the standard deviation, Bias <math display="inline"><semantics> <mrow> <mo>=</mo> <mi mathvariant="double-struck">E</mi> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> <mi>H</mi> </mrow> </semantics></math>. The values of Std and Bias are the estimated versions of those for 1000 independent copies of FBM with length <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <msup> <mn>2</mn> <mn>18</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>The Bias for estimators: circ denotes the circulant embedding method for simulation of FBM, chol denotes the Cholesky method for simulation of FBM. Bias <math display="inline"><semantics> <mrow> <mo>=</mo> <mi mathvariant="double-struck">E</mi> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> <mi>H</mi> </mrow> </semantics></math>. The values of Bias are the estimated versions of those for 1000 independent copies of FBM with length <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <msup> <mn>2</mn> <mn>12</mn> </msup> </mrow> </semantics></math>. The lower bound of octaves <span class="html-italic">j</span>s is chosen <math display="inline"><semantics> <mrow> <msub> <mi>j</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. The used wavelet is the Daubechies wavelet with <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> vanishing moments.</p>
Full article ">Figure 7
<p>The Bias and Std for estimators: Init1 denotes the initialization by itself (or by (<a href="#FD27-entropy-22-00349" class="html-disp-formula">27</a>)), Init2 denotes the initialization denoted by (<a href="#FD29-entropy-22-00349" class="html-disp-formula">29</a>). Std is the standard deviation, Bias <math display="inline"><semantics> <mrow> <mo>=</mo> <mi mathvariant="double-struck">E</mi> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo>−</mo> <mi>H</mi> </mrow> </semantics></math>. The values of Std and Bias are the estimated versions of those for 1000 independent copies of FBM with length <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <msup> <mn>2</mn> <mn>18</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>The Bias, Std, RMSE for estimators: orig stands for original series without noise, gau stands for Gaussian noise, unm stands for uniform noise, cau stands for Cauchy noise. The values of Std, Bias and RMSE are the estimated versions of those for 1000 independent copies of FBM with noise. The data length <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <msup> <mn>2</mn> <mn>18</mn> </msup> </mrow> </semantics></math>. The lower bound of octaves <span class="html-italic">j</span>s is chosen <math display="inline"><semantics> <mrow> <msub> <mi>j</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. The used wavelet is the Daubechies wavelet with <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> vanishing moments</p>
Full article ">
2 pages, 148 KiB  
Editorial
Carnot Cycle and Heat Engine: Fundamentals and Applications
by Michel Feidt
Entropy 2020, 22(3), 348; https://doi.org/10.3390/e22030348 - 18 Mar 2020
Cited by 11 | Viewed by 3622
Abstract
After two years of exchange, this specific issue dedicated to the Carnot cycle and thermomechanical engines has been completed with ten papers including this editorial [...] Full article
(This article belongs to the Special Issue Carnot Cycle and Heat Engine Fundamentals and Applications)
25 pages, 2770 KiB  
Article
A Novel Microwave Treatment for Sleep Disorders and Classification of Sleep Stages Using Multi-Scale Entropy
by Daoshuang Geng, Daoguo Yang, Miao Cai and Lixia Zheng
Entropy 2020, 22(3), 347; https://doi.org/10.3390/e22030347 - 17 Mar 2020
Cited by 8 | Viewed by 4138
Abstract
The aim of this study was to develop an integrated system of non-contact sleep stage detection and sleep disorder treatment for health monitoring. Hence, a method of brain activity detection based on microwave scattering technology instead of scalp electroencephalogram was developed to evaluate [...] Read more.
The aim of this study was to develop an integrated system of non-contact sleep stage detection and sleep disorder treatment for health monitoring. Hence, a method of brain activity detection based on microwave scattering technology instead of scalp electroencephalogram was developed to evaluate the sleep stage. First, microwaves at a specific frequency were used to penetrate the functional sites of the brain in patients with sleep disorders to change the firing frequency of the activated areas of the brain and analyze and evaluate statistically the effects on sleep improvement. Then, a wavelet packet algorithm was used to decompose the microwave transmission signal, the refined composite multiscale sample entropy, the refined composite multiscale fluctuation-based dispersion entropy and multivariate multiscale weighted permutation entropy were obtained as features from the wavelet packet coefficient. Finally, the mutual information-principal component analysis feature selection method was used to optimize the feature set and random forest was used to classify and evaluate the sleep stage. The results show that after four times of microwave modulation treatment, sleep efficiency improved continuously, the overall maintenance was above 80%, and the insomnia rate was reduced gradually. The overall classification accuracy of the four sleep stages was 86.4%. The results indicate that the microwaves with a certain frequency can treat sleep disorders and detect abnormal brain activity. Therefore, the microwave scattering method is of great significance in the development of a new brain disease treatment, diagnosis and clinical application system. Full article
(This article belongs to the Section Entropy and Biology)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of experimental paradigm: (<b>a</b>) sleep disorder treatment and sleep staging detection system; (<b>b</b>) schematic diagram of microwave emission and recording.</p>
Full article ">Figure 2
<p>Comparison of pre-denoising and post-denoising of a microwave transmission signals in the first 5000 s sleep test. The signal was digitally bandpass filtered using a fourth-order Butterworth filter between 0.5 and 150 Hz. The forward and backward filtering is used to reduce the phase distortion. Then, the time series was digitally filtered using the hamming window FIR bandpass filters of order 200, with cutoff frequencies of 0.5 Hz and 40 Hz, which are usually used to analyze brain activity.</p>
Full article ">Figure 3
<p>Bar chart of mean satisfaction scores of all subjects.</p>
Full article ">Figure 4
<p>Comparison of sleep efficiency of four tests, * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 5
<p>SOL comparison chart for four tests, * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 6
<p>Variation of the power spectral density of brain activity under the modulation of a false machine and a true machine.</p>
Full article ">Figure 7
<p>Mean measurements of RCMSE, RCMFDE, and MMSWPE for each sleep stage at different wavelet packet coefficients.</p>
Full article ">Figure 8
<p>Mean measurements of RCMSE, RCMFDE, and MMSWPE at various stages of sleep.</p>
Full article ">Figure 9
<p>Mean classification precision and accuracy of the four types of sleep stages of the 24 subjects.</p>
Full article ">
24 pages, 511 KiB  
Article
The Truncated Cauchy Power Family of Distributions with Inference and Applications
by Maha A. Aldahlan, Farrukh Jamal, Christophe Chesneau, Mohammed Elgarhy and Ibrahim Elbatal
Entropy 2020, 22(3), 346; https://doi.org/10.3390/e22030346 - 17 Mar 2020
Cited by 38 | Viewed by 4457
Abstract
As a matter of fact, the statistical literature lacks of general family of distributions based on the truncated Cauchy distribution. In this paper, such a family is proposed, called the truncated Cauchy power-G family. It stands out for the originality of the involved [...] Read more.
As a matter of fact, the statistical literature lacks of general family of distributions based on the truncated Cauchy distribution. In this paper, such a family is proposed, called the truncated Cauchy power-G family. It stands out for the originality of the involved functions, its overall simplicity and its desirable properties for modelling purposes. In particular, (i) only one parameter is added to the baseline distribution avoiding the over-parametrization phenomenon, (ii) the related probability functions (cumulative distribution, probability density, hazard rate, and quantile functions) have tractable expressions, and (iii) thanks to the combined action of the arctangent and power functions, the flexible properties of the baseline distribution (symmetry, skewness, kurtosis, etc.) can be really enhanced. These aspects are discussed in detail, with the support of comprehensive numerical and graphical results. Furthermore, important mathematical features of the new family are derived, such as the moments, skewness and kurtosis, two kinds of entropy and order statistics. For the applied side, new models can be created in view of fitting data sets with simple or complex structure. This last point is illustrated by the consideration of the Weibull distribution as baseline, the maximum likelihood method of estimation and two practical data sets wit different skewness properties. The obtained results show that the truncated Cauchy power-G family is very competitive in comparison to other well implanted general families. Full article
Show Figures

Figure 1

Figure 1
<p>Plots of the pdf of the TCPW distribution for various values of the three parameters.</p>
Full article ">Figure 2
<p>Plots of the hrf of the TCPW distribution for various values of the three parameters.</p>
Full article ">Figure 3
<p>Plots of the mean and variance for the TCPW distribution: (<b>a</b>) for fixed <math display="inline"><semantics> <mi>λ</mi> </semantics></math> and <math display="inline"><semantics> <mi>θ</mi> </semantics></math> and varying <math display="inline"><semantics> <mi>α</mi> </semantics></math> and (<b>b</b>) for fixed <math display="inline"><semantics> <mi>θ</mi> </semantics></math> and <math display="inline"><semantics> <mi>α</mi> </semantics></math> and varying <math display="inline"><semantics> <mi>λ</mi> </semantics></math>.</p>
Full article ">Figure 4
<p>Plots of the MacGillivray skewness for selected values of the parameters when (<b>a</b>) <math display="inline"><semantics> <mi>α</mi> </semantics></math> increases and (<b>b</b>) <math display="inline"><semantics> <mi>θ</mi> </semantics></math> increases.</p>
Full article ">Figure 5
<p>Plots of Galton skewness for selected values of the parameters when (<b>a</b>) <math display="inline"><semantics> <mi>α</mi> </semantics></math> varies and (<b>b</b>) <math display="inline"><semantics> <mi>θ</mi> </semantics></math> varies.</p>
Full article ">Figure 6
<p>Plots of Moors kurtosis for selected values of the parameters when (<b>a</b>) <math display="inline"><semantics> <mi>α</mi> </semantics></math> varies and (<b>b</b>) <math display="inline"><semantics> <mi>θ</mi> </semantics></math> varies.</p>
Full article ">Figure 7
<p>Estimated (<b>a</b>) pdfs and (<b>b</b>) cdfs of the considered models for the first data set.</p>
Full article ">Figure 8
<p>Estimated (<b>a</b>) pdfs and (<b>b</b>) cdfs of the considered models for the second data set.</p>
Full article ">
34 pages, 667 KiB  
Article
On the Rate-Distortion Function of Sampled Cyclostationary Gaussian Processes
by Emeka Abakasanga, Nir Shlezinger and Ron Dabora
Entropy 2020, 22(3), 345; https://doi.org/10.3390/e22030345 - 17 Mar 2020
Cited by 2 | Viewed by 2484
Abstract
Man-made communications signals are typically modelled as continuous-time (CT) wide-sense cyclostationary (WSCS) processes. As modern processing is digital, it is applied to discrete-time (DT) processes obtained by sampling the CT processes. When sampling is applied to a CT WSCS process, the statistics of [...] Read more.
Man-made communications signals are typically modelled as continuous-time (CT) wide-sense cyclostationary (WSCS) processes. As modern processing is digital, it is applied to discrete-time (DT) processes obtained by sampling the CT processes. When sampling is applied to a CT WSCS process, the statistics of the resulting DT process depends on the relationship between the sampling interval and the period of the statistics of the CT process: When these two parameters have a common integer factor, then the DT process is WSCS. This situation is referred to as synchronous sampling. When this is not the case, which is referred to as asynchronous sampling, the resulting DT process is wide-sense almost cyclostationary (WSACS). The sampled CT processes are commonly encoded using a source code to facilitate storage or transmission over wireless networks, e.g., using compress-and-forward relaying. In this work, we study the fundamental tradeoff between rate and distortion for source codes applied to sampled CT WSCS processes, characterized via the rate-distortion function (RDF). We note that while RDF characterization for the case of synchronous sampling directly follows from classic information-theoretic tools utilizing ergodicity and the law of large numbers, when sampling is asynchronous, the resulting process is not information stable. In such cases, the commonly used information-theoretic tools are inapplicable to RDF analysis, which poses a major challenge. Using the information-spectrum framework, we show that the RDF for asynchronous sampling in the low distortion regime can be expressed as the limit superior of a sequence of RDFs in which each element corresponds to the RDF of a synchronously sampled WSCS process (yet their limit is not guaranteed to exist). The resulting characterization allows us to introduce novel insights on the relationship between sampling synchronization and the RDF. For example, we demonstrate that, differently from stationary processes, small differences in the sampling rate and the sampling time offset can notably affect the RDF of sampled CT WSCS processes. Full article
(This article belongs to the Special Issue Wireless Networks: Information Theoretic Perspectives)
Show Figures

Figure 1

Figure 1
<p>Source coding block diagram.</p>
Full article ">Figure 2
<p><math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>D</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> versus <span class="html-italic">n</span>; offset <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p><math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>D</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> versus <span class="html-italic">n</span>; offset <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mfrac> <mn>1</mn> <mn>16</mn> </mfrac> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p><math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>D</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> versus <math display="inline"><semantics> <mfrac> <msub> <mi>T</mi> <mi>ps</mi> </msub> <msub> <mi>T</mi> <mi mathvariant="normal">s</mi> </msub> </mfrac> </semantics></math>; offset <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p><math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>D</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> versus <math display="inline"><semantics> <mfrac> <msub> <mi>T</mi> <mi>ps</mi> </msub> <msub> <mi>T</mi> <mi mathvariant="normal">s</mi> </msub> </mfrac> </semantics></math>; offset <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mfrac> <mn>1</mn> <mn>16</mn> </mfrac> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p><math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>D</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> versus <span class="html-italic">D</span>; offset <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p><math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>D</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> versus <span class="html-italic">D</span>; offset <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mfrac> <mn>1</mn> <mn>16</mn> </mfrac> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p><math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>D</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> versus <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>dc</mi> </msub> <mo>=</mo> <mn>75</mn> <mo>%</mo> </mrow> </semantics></math>.</p>
Full article ">
19 pages, 4284 KiB  
Article
No-Reference Image Quality Assessment Based on Dual-Domain Feature Fusion
by Yueli Cui
Entropy 2020, 22(3), 344; https://doi.org/10.3390/e22030344 - 17 Mar 2020
Cited by 8 | Viewed by 3768
Abstract
Image quality assessment (IQA) aims to devise computational models to evaluate image quality in a perceptually consistent manner. In this paper, a novel no-reference image quality assessment model based on dual-domain feature fusion is proposed, dubbed as DFF-IQA. Firstly, in the spatial domain, [...] Read more.
Image quality assessment (IQA) aims to devise computational models to evaluate image quality in a perceptually consistent manner. In this paper, a novel no-reference image quality assessment model based on dual-domain feature fusion is proposed, dubbed as DFF-IQA. Firstly, in the spatial domain, several features about weighted local binary pattern, naturalness and spatial entropy are extracted, where the naturalness features are represented by fitting parameters of the generalized Gaussian distribution. Secondly, in the frequency domain, the features of spectral entropy, oriented energy distribution, and fitting parameters of asymmetrical generalized Gaussian distribution are extracted. Thirdly, the features extracted in the dual-domain are fused to form the quality-aware feature vector. Finally, quality regression process by random forest is conducted to build the relationship between image features and quality score, yielding a measure of image quality. The resulting algorithm is tested on the LIVE database and compared with competing IQA models. Experimental results on the LIVE database indicate that the proposed DFF-IQA method is more consistent with the human visual system than other competing IQA methods. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the proposed DFF-IQA framework.</p>
Full article ">Figure 2
<p>Pristine natural image and five distorted versions of it from the LIVE IQA database (“parrots” in LIVE database); from left column to right column are the input image, gradient image, LBP map, and histogram of WLBP. From top to bottom of the first column are pristine image with DMOS = 0, JPEG2000 compressed image with DMOS = 45.8920, JPEG compressed image with DMOS = 46.8606, white noise image with DMOS = 47.0386, fast-fading distorted image with DMOS = 44.0640, and Gaussian blur image with DMOS = 49.1911.</p>
Full article ">Figure 3
<p>(<b>a</b>) Distorted image (“parrots” in LIVE database, type of JP2K compressed image with DMOS = 45.8920); (<b>b</b>) Corresponding MSCN coefficients image of (<b>a</b>); (<b>c</b>) Statistical distribution of the distorted image; (<b>d</b>) MSCN histogram distribution of MSCN coefficients.</p>
Full article ">Figure 4
<p>Histogram of MSCN coefficients for a reference image and its various distorted versions. Distortions from the LIVE IQA database. org: original image (i.e., Pristine natural image). jp2k: JPEG2000. jpeg: JPEG compression. wn: additive white Gaussian noise. blur: Gaussian blur. ff: Rayleigh fast-fading channel simulation.</p>
Full article ">Figure 5
<p>Histograms of spatial and spectral entropy values for different types of distortion. The ordinate represents the coefficient normalized between 0 and 1.</p>
Full article ">Figure 6
<p>Various paired products computed in order to quantify neighboring statistical relationships. Pairwise products are computed along four orientations—horizontal, vertical, main-diagonal, and secondary-diagonal at a distance of 1 pixel.</p>
Full article ">Figure 7
<p>Histograms of paired products of MSCN coefficients of a natural undistorted image and various distorted versions of it. (<b>a</b>) Horizontal; (<b>b</b>) Vertical; (<b>c</b>) Main-diagonal; (<b>d</b>) Secondary-diagonal. Distortions from the LIVE IQA database. jp2k: JPEG2000. jpeg: JPEG compression. wn: additive white Gaussian noise. gblur: Gaussian blur. ff: Rayleigh fast-fading channel simulation.</p>
Full article ">Figure 8
<p>Some examples of reference scenes in the LIVE database. (<b>a</b>–<b>i</b>) shows some examples of reference scenes in the LIVE database, including “bikes scene”, “buildings scene”, “caps scene”, “lighthouse2 scene”, “monarch scene”, “ocean scene”, “parrots scene”, “plane scene” and “rapids scene” (not listed one by one due to layout reasons).</p>
Full article ">Figure 9
<p>Scatter plot of the proposed model on LIVE database. Each point (‘+’) represents one test image. The red curve shown in <a href="#entropy-22-00344-f009" class="html-fig">Figure 9</a> is obtained by a logistic function.</p>
Full article ">Figure 10
<p>Box plot of PLCC distributions of the compared IQA methods over 1000 trials on the LIVE database.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop