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

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (15)

Search Parameters:
Keywords = principle of maximum conformality

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 365 KiB  
Article
Approximate N5LO Higgs Boson Decay Width Γ(Hγγ)
by Yu-Feng Luo, Jiang Yan, Zhi-Fei Wu and Xing-Gang Wu
Symmetry 2024, 16(2), 173; https://doi.org/10.3390/sym16020173 - 1 Feb 2024
Cited by 1 | Viewed by 1041
Abstract
The precision and predictive power of perturbative QCD (pQCD) prediction depends on both a precise, convergent, fixed-order series and a reliable way of estimating the contributions of unknown higher-order (UHO) terms. It has been shown that by applying the principle of maximum conformality [...] Read more.
The precision and predictive power of perturbative QCD (pQCD) prediction depends on both a precise, convergent, fixed-order series and a reliable way of estimating the contributions of unknown higher-order (UHO) terms. It has been shown that by applying the principle of maximum conformality (PMC), which applies the renormalization group equation recursively to set the effective magnitude of αs of the process, the remaining conformal coefficients will be well matched with the corresponding αs at each order, leading to a scheme-and-scale invariant and more convergent perturbative series. The PMC series, being satisfied with the standard renormalization group invariance, has a rigorous foundation. Thus it not only can be widely applied to virtually all high-energy hadronic processes, but also can be a reliable platform for estimating UHO contributions. In this paper, by using the total decay width Γ(Hγγ) which has been calculated up to N4LO QCD corrections, we first derive its PMC series by using the PMC single-scale setting approach and then estimate its unknown N5LO contributions by using a Bayesian analysis. The newly suggested Bayesian-based approach estimates the magnitude of the UHO contributions based on an optimized analysis of the probability density distribution, and the predicted UHO contribution becomes more accurate when more loop terms have been known to tame the probability density function. Using the top-quark pole mass Mt = 172.69 GeV and the Higgs mass MH = 125.25 GeV as inputs, we obtain Γ(Hγγ)=9.56504keV, and the estimated N5LO contribution to the total decay width is ΔΓH=±1.65×104keV for the smallest credible interval of 95.5% degree of belief. Full article
(This article belongs to the Special Issue Symmetry on Multiboson Physics)
Show Figures

Figure 1

Figure 1
<p>The predicted values for the pQCD correction <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <msub> <mrow> <mo>|</mo> </mrow> <mi>PMC</mi> </msub> </mrow> </semantics></math> under the Padé approximation approach (PAA) and Bayesian approach (B.A.) at different orders, respectively. The blue rectangles together with the error bars are for B.A., the green error bars are brought by different types of PAAs, and the exact values of the <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>n</mi> </msub> <mrow> <mo>(</mo> <msub> <mi>M</mi> <mi>H</mi> </msub> <mo>)</mo> </mrow> <msub> <mrow> <mo>|</mo> </mrow> <mi>PMC</mi> </msub> </mrow> </semantics></math> at different orders, respectively.</p>
Full article ">Figure 2
<p>The fiducial cross section <math display="inline"><semantics> <mrow> <msub> <mi>σ</mi> <mi>fid</mi> </msub> <mrow> <mo>(</mo> <mi>p</mi> <mi>p</mi> <mo>→</mo> <mi>H</mi> <mo>→</mo> <mi>γ</mi> <mi>γ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> using <math display="inline"><semantics> <mrow> <mo>Γ</mo> <mo>(</mo> <mi>H</mi> <mo>→</mo> <mi>γ</mi> <mi>γ</mi> <mo>)</mo> </mrow> </semantics></math> up to the N<sup>4</sup>LO level. The LHC−XS prediction, the ATLAS measurements [<a href="#B62-symmetry-16-00173" class="html-bibr">62</a>,<a href="#B65-symmetry-16-00173" class="html-bibr">65</a>,<a href="#B66-symmetry-16-00173" class="html-bibr">66</a>,<a href="#B67-symmetry-16-00173" class="html-bibr">67</a>], and the CMS measurement [<a href="#B68-symmetry-16-00173" class="html-bibr">68</a>,<a href="#B69-symmetry-16-00173" class="html-bibr">69</a>,<a href="#B70-symmetry-16-00173" class="html-bibr">70</a>] are presented as a comparison.</p>
Full article ">
22 pages, 4289 KiB  
Article
A Method for Precise Tracking Control of Pneumatic Artificial-Muscle-Driven Exoskeletal Robot
by Gaoke Ma, Hongyun Jia, Jichun Xiao and Lina Hao
Appl. Sci. 2023, 13(21), 12038; https://doi.org/10.3390/app132112038 - 4 Nov 2023
Cited by 2 | Viewed by 1059
Abstract
Exoskeletal robots are of critical importance in the domain of mechanical boosting. The pneumatic artificial muscle (PAM) is commonly used as a flexible actuator in exoskeletal robots designed for upper limbs due to its high power-to-weight ratio, conformability, and safety. This study establishes [...] Read more.
Exoskeletal robots are of critical importance in the domain of mechanical boosting. The pneumatic artificial muscle (PAM) is commonly used as a flexible actuator in exoskeletal robots designed for upper limbs due to its high power-to-weight ratio, conformability, and safety. This study establishes a new model based on the existing model to improve its control precision by implementing elastic and frictional forces and empirical coefficients, battling against the time-variant hysteresis that PAM’s output force exhibits. In the meantime, a BP neural network is employed in reverse modeling, followed by the adoption of the least-square-based particle swarm optimization algorithm in order to determine the optimized parameter values. PAM provides the Upper Limb Exoskeletal Robot with appropriate auxiliary power, which can be adjusted to accommodate variations in posture change during the lifting process. PAM is also capable of handling variable loads based on the principle of torque balance, constructing a control system according to the inverse dynamics of exoskeletal robots accompanied by an inverse model of PAM’s output force, and finally, rendering tracking control of the elbow angle during the auxiliary process possible. Finally, the tracking error results are calculated and shown; the maximum angular error in the tracking process is 0.0175 rad, MAE value is 0.0038 rad, RMSE value is 0.0048 rad, and IEAT value is 4.6426 rad. This control method is able to improve the precision of tracking control of the elbow angle of the upper limb–exoskeleton coupled system during the process of lifting goods. Full article
Show Figures

Figure 1

Figure 1
<p>Self-assembled McKibben-style pneumatic muscle.</p>
Full article ">Figure 2
<p>Structure of pneumatic muscle.</p>
Full article ">Figure 3
<p>Analysis of force on strand mesh.</p>
Full article ">Figure 4
<p>PAM data collection platform.</p>
Full article ">Figure 5
<p>Schematic drawing of the experiment.</p>
Full article ">Figure 6
<p>The structure of the flexible Upper Limb Exoskeletal robot. (<b>a</b>) Structural Diagram. (<b>b</b>) Actual Picture.</p>
Full article ">Figure 7
<p>Analysis of force during lifting process.</p>
Full article ">Figure 8
<p>Establishment of Upper Limb Exoskeleton coordinate axis. (<b>a</b>) Physical map. (<b>b</b>) Structure diagram.</p>
Full article ">Figure 9
<p>Target trajectory planning.</p>
Full article ">Figure 10
<p>Control system flow chart of Upper Limb Exoskeleton handling process.</p>
Full article ">Figure 11
<p>Schematic diagram of the Upper Limb Exoskeleton assisting process.</p>
Full article ">Figure 12
<p>Load change diagrams of left and right arms during handling. (<b>a</b>) Load torque process of elbow joint. (<b>b</b>) Pneumatic muscle loading pressure process.</p>
Full article ">Figure 13
<p>Changes in the pneumatic muscles of both arms. (<b>a</b>) Process of pneumatic muscle deformation. (<b>b</b>) Process of pneumatic muscle length variation. (<b>c</b>) Process of pneumatic muscle length variation velocity. (<b>d</b>) Process of pneumatic muscle length variation acceleration.</p>
Full article ">Figure 14
<p>Control torque and air pressure changes in left and right arms. (<b>a</b>) Process of Elbow joint control torque process. (<b>b</b>) Process of Pneumatic muscle control air pressure.</p>
Full article ">Figure 15
<p>Left and right elbow joint tracking results. (<b>a</b>) Comparison of elbow angle tracking. (<b>b</b>) Real-time image of elbow angle tracking error. (<b>c</b>) Comparison of elbow angular velocity tracking. (<b>d</b>) Real-time image of angular velocity tracking error of elbow joint. (<b>e</b>) Comparison of elbow angular acceleration tracking. (<b>f</b>) Real-time map of elbow angular acceleration tracking error.</p>
Full article ">Figure 15 Cont.
<p>Left and right elbow joint tracking results. (<b>a</b>) Comparison of elbow angle tracking. (<b>b</b>) Real-time image of elbow angle tracking error. (<b>c</b>) Comparison of elbow angular velocity tracking. (<b>d</b>) Real-time image of angular velocity tracking error of elbow joint. (<b>e</b>) Comparison of elbow angular acceleration tracking. (<b>f</b>) Real-time map of elbow angular acceleration tracking error.</p>
Full article ">
14 pages, 312 KiB  
Article
Maximum Principle for Variable-Order Fractional Conformable Differential Equation with a Generalized Tempered Fractional Laplace Operator
by Tingting Guan and Lihong Zhang
Fractal Fract. 2023, 7(11), 798; https://doi.org/10.3390/fractalfract7110798 - 1 Nov 2023
Cited by 1 | Viewed by 1486
Abstract
In this paper, we investigate properties of solutions to a space-time fractional variable-order conformable nonlinear differential equation with a generalized tempered fractional Laplace operatorby using the maximum principle. We first establish some new important fractional various-order conformable inequalities. With these inequalities, we prove [...] Read more.
In this paper, we investigate properties of solutions to a space-time fractional variable-order conformable nonlinear differential equation with a generalized tempered fractional Laplace operatorby using the maximum principle. We first establish some new important fractional various-order conformable inequalities. With these inequalities, we prove a new maximum principle with space-time fractional variable-order conformable derivatives and a generalized tempered fractional Laplace operator. Moreover, we discuss some results about comparison principles and properties of solutions for a family of space-time fractional variable-order conformable nonlinear differential equations with a generalized tempered fractional Laplace operator by maximum principle. Full article
19 pages, 561 KiB  
Review
Elimination of QCD Renormalization Scale and Scheme Ambiguities
by Sheng-Quan Wang, Stanley J. Brodsky, Xing-Gang Wu, Jian-Ming Shen and Leonardo Di Giustino
Universe 2023, 9(4), 193; https://doi.org/10.3390/universe9040193 - 17 Apr 2023
Cited by 4 | Viewed by 2018
Abstract
The setting of the renormalization scale (μr) in the perturbative QCD (pQCD) is one of the crucial problems for achieving precise fixed-order pQCD predictions. The conventional prescription is to take its value as the typical momentum transfer Q in a [...] Read more.
The setting of the renormalization scale (μr) in the perturbative QCD (pQCD) is one of the crucial problems for achieving precise fixed-order pQCD predictions. The conventional prescription is to take its value as the typical momentum transfer Q in a given process, and theoretical uncertainties are then evaluated by varying it over an arbitrary range. The conventional scale-setting procedure introduces arbitrary scheme-and-scale ambiguities in fixed-order pQCD predictions. The principle of maximum conformality (PMC) provides a systematic way to eliminate the renormalization scheme-and-scale ambiguities. The PMC method has rigorous theoretical foundations; it satisfies the renormalization group invariance (RGI) and all of the self-consistency conditions derived from the renormalization group. The PMC has now been successfully applied to many physical processes. In this paper, we summarize recent PMC applications, including event shape observables and heavy quark pair production near the threshold region in e+e annihilation and top-quark decay at hadronic colliders. In addition, estimating the contributions related to the uncalculated higher-order terms is also summarized. These results show that the major theoretical uncertainties caused by different choices of μr are eliminated, and the improved pQCD predictions are thus obtained, demonstrating the generality and applicability of the PMC. Full article
(This article belongs to the Special Issue The Quantum Chromodynamics: 50th Anniversary of the Discovery)
Show Figures

Figure 1

Figure 1
<p>The thrust (<span class="html-italic">T</span>) and <span class="html-italic">C</span>-parameter (<span class="html-italic">C</span>) distributions using the conventional scale-setting method at <math display="inline"><semantics> <mrow> <msqrt> <mi>s</mi> </msqrt> <mo>=</mo> <mn>91.2</mn> </mrow> </semantics></math> GeV, where the dashed, dot-dashed, and dotted lines are the conventional results at LO, NLO, and NNLO [<a href="#B45-universe-09-00193" class="html-bibr">45</a>,<a href="#B48-universe-09-00193" class="html-bibr">48</a>], respectively. The bands are obtained by varying the scale <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi>r</mi> </msub> <mo>∈</mo> <mrow> <mo>[</mo> <msqrt> <mi>s</mi> </msqrt> <mo>/</mo> <mn>2</mn> <mo>,</mo> <mn>2</mn> <msqrt> <mi>s</mi> </msqrt> <mo>]</mo> </mrow> </mrow> </semantics></math>. The experimental data are taken from the ALEPH Collaboration [<a href="#B39-universe-09-00193" class="html-bibr">39</a>].</p>
Full article ">Figure 2
<p>The PMC scales for the event shape observables thrust (<span class="html-italic">T</span>) and <span class="html-italic">C</span>-parameter (<span class="html-italic">C</span>) at <math display="inline"><semantics> <mrow> <msqrt> <mi>s</mi> </msqrt> <mo>=</mo> <mn>91.2</mn> </mrow> </semantics></math> GeV.</p>
Full article ">Figure 3
<p>The thrust (<span class="html-italic">T</span>) and <span class="html-italic">C</span>-parameter (<span class="html-italic">C</span>) distributions using PMC scale setting for <math display="inline"><semantics> <mrow> <msqrt> <mi>s</mi> </msqrt> <mo>=</mo> <mn>91.2</mn> </mrow> </semantics></math> GeV. The experimental data are taken from the ALEPH Collaboration [<a href="#B39-universe-09-00193" class="html-bibr">39</a>].</p>
Full article ">Figure 4
<p>The extracted running coupling <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mi>s</mi> </msub> <mrow> <mo>(</mo> <msup> <mi>Q</mi> <mn>2</mn> </msup> <mo>)</mo> </mrow> </mrow> </semantics></math> from the thrust (<span class="html-italic">T</span>) and <span class="html-italic">C</span>-parameter (<span class="html-italic">C</span>) distributions by comparing the PMC predictions with the ALEPH data [<a href="#B39-universe-09-00193" class="html-bibr">39</a>] measured at a single energy of <math display="inline"><semantics> <mrow> <msqrt> <mi>s</mi> </msqrt> <mo>=</mo> <mn>91.2</mn> </mrow> </semantics></math> GeV. As a comparison, the three lines represent the world average [<a href="#B38-universe-09-00193" class="html-bibr">38</a>].</p>
Full article ">Figure 5
<p>The extracted QCD coupling <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mi>s</mi> </msub> <mrow> <mo>(</mo> <msup> <mi>Q</mi> <mn>2</mn> </msup> <mo>)</mo> </mrow> </mrow> </semantics></math> from the thrust and <span class="html-italic">C</span>-parameter mean values by comparing PMC predictions with the JADE and OPAL data [<a href="#B41-universe-09-00193" class="html-bibr">41</a>,<a href="#B58-universe-09-00193" class="html-bibr">58</a>]. The error bars are the squared averages of the experimental and theoretical errors. The three lines are the world average [<a href="#B38-universe-09-00193" class="html-bibr">38</a>].</p>
Full article ">Figure 6
<p>The behavior of the Coulomb terms in the V-scheme for the <span class="html-italic">b</span> quark pair production, where <math display="inline"><semantics> <mrow> <msubsup> <mi>δ</mi> <mi>v</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mrow> <mo>=</mo> <mo>(</mo> </mrow> <msubsup> <mi>δ</mi> <mrow> <mi>v</mi> <mo>,</mo> <mi>i</mi> <mi>n</mi> </mrow> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <msub> <mrow> <mo>|</mo> </mrow> <mi>V</mi> </msub> <mo>+</mo> <msubsup> <mi>δ</mi> <mrow> <mi>v</mi> <mo>,</mo> <msub> <mi>n</mi> <mi>f</mi> </msub> </mrow> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mrow> <msub> <mo>|</mo> <mi>V</mi> </msub> <mspace width="0.166667em"/> <msub> <mi>n</mi> <mi>f</mi> </msub> <mo>)</mo> </mrow> </mrow> </semantics></math> for conventional scale setting, and for PMC scale setting, <math display="inline"><semantics> <mrow> <msubsup> <mi>δ</mi> <mi>v</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <msubsup> <mi>δ</mi> <mrow> <mi>v</mi> <mo>,</mo> <mi>sc</mi> </mrow> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <msub> <mrow> <mo>|</mo> </mrow> <mi>V</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>The behavior of the Coulomb terms for the <math display="inline"><semantics> <mi>τ</mi> </semantics></math> lepton pair production, where <math display="inline"><semantics> <mrow> <msubsup> <mi>δ</mi> <mi>v</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <mrow> <mo>(</mo> <msubsup> <mi>δ</mi> <mrow> <mi>v</mi> <mo>,</mo> <mi>i</mi> <mi>n</mi> </mrow> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>+</mo> <msubsup> <mi>δ</mi> <mrow> <mi>v</mi> <mo>,</mo> <msub> <mi>n</mi> <mi>f</mi> </msub> </mrow> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mspace width="0.166667em"/> <msub> <mi>n</mi> <mi>f</mi> </msub> <mo>)</mo> </mrow> </mrow> </semantics></math> for conventional scale setting, and for PMC scale setting, <math display="inline"><semantics> <mrow> <msubsup> <mi>δ</mi> <mi>v</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <msubsup> <mi>δ</mi> <mrow> <mi>v</mi> <mo>,</mo> <mi>i</mi> <mi>n</mi> </mrow> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>The top-quark total decay width <math display="inline"><semantics> <msubsup> <mi mathvariant="sans-serif">Γ</mi> <mi>t</mi> <mi>tot</mi> </msubsup> </semantics></math> versus the top-quark mass <math display="inline"><semantics> <msub> <mi>m</mi> <mi>t</mi> </msub> </semantics></math>. The solid line is the PMC prediction, and the dashed line stands for the conventional prediction. As a comparison, the CMS measured value [<a href="#B83-universe-09-00193" class="html-bibr">83</a>] is also presented.</p>
Full article ">Figure 9
<p>The probability density distributions of <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>(</mo> <mi>Q</mi> <mo>=</mo> <mn>31.6</mn> <mspace width="0.277778em"/> <mi>GeV</mi> <mo>)</mo> </mrow> </semantics></math> with different states of knowledge predicted by PMCs and BA. The blue dotted, the black dash-dotted, the green solid, and the red dashed curves represent the results with given LO, NLO, N<math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math>LO, and N<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math>LO series, respectively.</p>
Full article ">
14 pages, 8513 KiB  
Article
Bending and Torsional Stress Factors in Hypotrochoidal H-Profiled Shafts Standardised According to DIN 3689-1
by Masoud Ziaei
Eng 2023, 4(1), 829-842; https://doi.org/10.3390/eng4010050 - 6 Mar 2023
Cited by 4 | Viewed by 2334
Abstract
Hypotrochoidal profile contours have been produced in industrial applications in recent years using two-spindle processes, and they are considered effective high-quality solutions for form-fit shaft and hub connections. This study mainly concerns analytical approaches to determine the stresses and deformations in hypotrochoidal profile [...] Read more.
Hypotrochoidal profile contours have been produced in industrial applications in recent years using two-spindle processes, and they are considered effective high-quality solutions for form-fit shaft and hub connections. This study mainly concerns analytical approaches to determine the stresses and deformations in hypotrochoidal profile shafts due to pure bending loads. The formulation was developed according to bending principles using the mathematical theory of elasticity and conformal mappings. The loading was further used to investigate the rotating bending behaviour. The stress factors for the classical calculation of maximum bending stresses were also determined for all those profiles presented and compiled in the German standard DIN3689-1 for practical applications. The results were also compared with the corresponding numerical and experimental results, and very good agreement was observed. Additionally, based on previous work, the stress factor was determined for the case of torsional loading to calculate the maximum torsional stresses in the standardised profiles, and the results are listed in a table. This study contributes to the further refinement of the current DIN3689 standard. Full article
(This article belongs to the Special Issue Feature Papers in Eng 2022)
Show Figures

Figure 1

Figure 1
<p>Description of exemplary hypotrochoid (H-profile) with four concave sides. A detailed explanation of the parameters is given below in <a href="#sec2-eng-04-00050" class="html-sec">Section 2</a>.</p>
Full article ">Figure 2
<p>Some H-profiles manufactured by two-spindle process, Iprotec GmbH, © Guido Kochsiek, <a href="http://www.iprotec.de" target="_blank">www.iprotec.de</a>, Zwiesel, Germany [<a href="#B5-eng-04-00050" class="html-bibr">5</a>].</p>
Full article ">Figure 3
<p>Roller milling manufacturing for H-profile [<a href="#B7-eng-04-00050" class="html-bibr">7</a>].</p>
Full article ">Figure 4
<p>Examples of H-profiles with different numbers of sides (<span class="html-italic">n</span>) and eccentricities.</p>
Full article ">Figure 5
<p>The bending coordinate system for a loaded profile shaft.</p>
Full article ">Figure 6
<p>FE mesh and boundary conditions for the H-profile with <span class="html-italic">n</span> = 3 according to DIN 3689-1.</p>
Full article ">Figure 7
<p>Circumferential distribution of the bending stress on the lateral surface of a standardised H3 profile.</p>
Full article ">Figure 8
<p>Bending loads test bench (Machine Elements Laboratory at West Saxon University of Zwickau).</p>
Full article ">Figure 9
<p>Comparison of the experimental results with the analytical solutions.</p>
Full article ">Figure 10
<p>Stress factors for the bending stress at the profile head (Equation (18)) with varying relative eccentricity and number of sides.</p>
Full article ">Figure 11
<p>Rotated coordinate system for determining the bending moment of inertia.</p>
Full article ">Figure 12
<p>Distributions of the bending stresses on the profile contour for different angles of rotation <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math>, with <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>18.18</mn> </mrow> </semantics></math> mm, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>e</mi> <mo>=</mo> <mn>1.818</mn> </mrow> </semantics></math> mm, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>M</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> = 500 Nm.</p>
Full article ">Figure 13
<p>Deflection in a DIN3689-H3-40 × 32.73 profile shaft.</p>
Full article ">
21 pages, 14300 KiB  
Article
Wearable Polarization Conversion Metasurface MIMO Antenna for Biomedical Applications in 5 GHz WBAN
by Rigeng Wu, Jian Dong and Meng Wang
Biosensors 2023, 13(1), 73; https://doi.org/10.3390/bios13010073 - 1 Jan 2023
Cited by 29 | Viewed by 3304
Abstract
This paper presents a wearable metasurface multiple-input multiple-output (MIMO) antenna for biomedical applications in a 5 GHz wireless body area network (WBAN) with broadband, circular polarization (CP), and high gain. The physical properties of the MIMO antenna element and the principles of polarization [...] Read more.
This paper presents a wearable metasurface multiple-input multiple-output (MIMO) antenna for biomedical applications in a 5 GHz wireless body area network (WBAN) with broadband, circular polarization (CP), and high gain. The physical properties of the MIMO antenna element and the principles of polarization conversion are analyzed in-depth using characteristic mode analysis. For the proposed MIMO antenna, the measured −10 dB impedance bandwidth is 34.87% (4.76–6.77 GHz), and the 3 dB axial ratio bandwidth is 22.94% (4.9–6.17 GHz). By adding an isolation strip, the measured isolation of the two antenna elements is greater than 19.85 dB. The overall size of the MIMO antenna is 1.67λ0 × 0.81λ0 × 0.07λ0 at 5.6 GHz, and the maximum gain is 7.95 dBic. The envelope correlation coefficient (ECC) is less than 0.007, with the maximum diversity gain greater than 9.98 dB, and the channel capacity loss is less than 0.29 b/s/Hz. The specific absorption rate (SAR) of the wearable MIMO antenna is simulated by the human tissue model, which proves that the proposed antenna conforms to international standards and is harmless to humans. The proposed wearable metasurface MIMO antenna has CP, broadband, high gain, low ECC, and low SAR, which can be used in wearable devices for biomedical applications. Full article
(This article belongs to the Special Issue Devices and Wearable Devices toward Innovative Applications)
Show Figures

Figure 1

Figure 1
<p>Structure of the MIMO antenna element: (<b>a</b>) Front view; (<b>b</b>) Back view; (<b>c</b>) Side view.</p>
Full article ">Figure 2
<p>Slot layer structure and CMA results: (<b>a</b>) Slot layer structure; (<b>b</b>) MS.</p>
Full article ">Figure 3
<p>The characteristic current of the slot layer: (<b>a</b>) The characteristic current of Mode 1s at 4.8 GHz; (<b>b</b>) The characteristic current of Mode 2s at 4.7 GHz.</p>
Full article ">Figure 4
<p>CMA results of metasurface layer: (<b>a</b>) MS; (<b>b</b>) CV; (<b>c</b>) CA; (<b>d</b>) Phase difference.</p>
Full article ">Figure 5
<p>Far-field results for each characteristic mode at 6 GHz: (<b>a</b>) Mode 1; (<b>b</b>) Mode 2; (<b>c</b>) Mode 3; (<b>d</b>) Mode 4; (<b>e</b>) Mode 5; (<b>f</b>) Mode 6.</p>
Full article ">Figure 6
<p>CMA results of the overall structure of the MIMO antenna element: (<b>a</b>) MS; (<b>b</b>) CV; (<b>c</b>) CA; (<b>d</b>) Phase Difference.</p>
Full article ">Figure 7
<p>MWC results of the overall MIMO antenna element: (<b>a</b>) MWC; (<b>b</b>) Phase difference of MWC.</p>
Full article ">Figure 8
<p>Fabricated MIMO antenna element and measurement environment: (<b>a</b>) Fabricated MIMO antenna element; (<b>b</b>) Anechoic chamber measurement environment.</p>
Full article ">Figure 9
<p>Simulated and measured S<sub>11</sub> and ARBW results of the proposed MIMO antenna element: (<b>a</b>) S<sub>11</sub>; (<b>b</b>) ARBW.</p>
Full article ">Figure 10
<p>Simulated and measured gain and efficiency results of the proposed MIMO antenna element.</p>
Full article ">Figure 11
<p>Comparing the simulated and measured radiation patterns of the proposed MIMO antenna element at 6 GHz: (<b>a</b>) Phi = 0°; (<b>b</b>) Phi = 90°.</p>
Full article ">Figure 12
<p>Structure of the two-port MIMO antenna: (<b>a</b>) Front view; (<b>b</b>) Back view; (<b>c</b>) Side view.</p>
Full article ">Figure 13
<p>MIMO antenna without and with isolation strip: (<b>a</b>) Without isolation strip; (<b>b</b>) With isolation strip.</p>
Full article ">Figure 14
<p>Simulated S-parameters of MIMO antenna without and with isolation strip: (<b>a</b>) Without isolation strip; (<b>b</b>) With isolation strip.</p>
Full article ">Figure 15
<p>Human phantom model: (<b>a</b>) Perspective view; (<b>b</b>) Side view.</p>
Full article ">Figure 16
<p>SAR results for simulated 1 g tissue at different frequencies: (<b>a</b>) 5 GHz; (<b>b</b>) 5.5 GHz; (<b>c</b>) 6 GHz.</p>
Full article ">Figure 17
<p>SAR results for simulated 10 g tissue at different frequencies: (<b>a</b>) 5 GHz; (<b>b</b>) 5.5 GHz; (<b>c</b>) 6 GHz.</p>
Full article ">Figure 18
<p>Antenna object and measurement environment: (<b>a</b>) Fabricated MIMO antenna; (<b>b</b>) Anechoic chamber measurement environment.</p>
Full article ">Figure 19
<p>Comparison of simulated and measured results of S-parameters and ARBW: (<b>a</b>) S-parameters; (<b>b</b>) ARBW.</p>
Full article ">Figure 20
<p>Comparison of simulated and measured results of the MIMO antenna gain and efficiency.</p>
Full article ">Figure 21
<p>Simulated and measured radiation pattern of MIMO antenna at 5.6 GHz: (<b>a</b>) Port 1 at phi = 0°; (<b>b</b>) Port 2 at phi = 0°; (<b>c</b>) Port 1 at phi = 90°; (<b>d</b>) Port 2 at phi = 90°.</p>
Full article ">Figure 21 Cont.
<p>Simulated and measured radiation pattern of MIMO antenna at 5.6 GHz: (<b>a</b>) Port 1 at phi = 0°; (<b>b</b>) Port 2 at phi = 0°; (<b>c</b>) Port 1 at phi = 90°; (<b>d</b>) Port 2 at phi = 90°.</p>
Full article ">Figure 22
<p>Simulated and measured results of ECC and DG.</p>
Full article ">Figure 23
<p>Simulated and measured results of ME and TARC.</p>
Full article ">Figure 24
<p>Simulated and measured results of MEG and CCL: (<b>a</b>) MEG; (<b>b</b>) CCL.</p>
Full article ">Figure 25
<p>Measured S-parameters for placing the MIMO antenna at different parts of the human body: (<b>a</b>) S<sub>11</sub> and S<sub>22</sub>; (<b>b</b>) S<sub>21</sub> and S<sub>12</sub>.</p>
Full article ">Figure 26
<p>Measured ARBW of MIMO antenna placed at different parts of the human body.</p>
Full article ">
15 pages, 3200 KiB  
Article
Parameter Optimization of the 3PG Model Based on Sensitivity Analysis and a Bayesian Method
by Chenjian Liu, Xiaoman Zheng and Yin Ren
Forests 2020, 11(12), 1369; https://doi.org/10.3390/f11121369 - 21 Dec 2020
Cited by 8 | Viewed by 2614
Abstract
Sensitivity analysis and parameter optimization of stand models can improve their efficiency and accuracy, and increase their applicability. In this study, the sensitivity analysis, screening, and optimization of 63 model parameters of the Physiological Principles in Predicting Growth (3PG) model were performed by [...] Read more.
Sensitivity analysis and parameter optimization of stand models can improve their efficiency and accuracy, and increase their applicability. In this study, the sensitivity analysis, screening, and optimization of 63 model parameters of the Physiological Principles in Predicting Growth (3PG) model were performed by combining a sensitivity analysis method and the Markov chain Monte Carlo (MCMC) method of Bayesian posterior estimation theory. Additionally, a nine-year observational dataset of Chinese fir trees felled in the Shunchang Forest Farm, Nanping, was used to analyze, screen, and optimize the 63 model parameters of the 3PG model. The results showed the following: (1) The parameters that are most sensitive to stand stocking and diameter at breast height (DBH) are nWs(power in stem mass vs. diameter relationship), aWs(constant in stem mass vs. diameter relationship), alphaCx(maximum canopy quantum efficiency), k(extinction coefficient for PAR absorption by canopy), pRx(maximum fraction of NPP to roots), pRn(minimum fraction of NPP to roots), and CoeffCond(defines stomatal response to VPD); (2) MCMC can be used to optimize the parameters of the 3PG model, in which the posterior probability distributions of nWs, aWs, alphaCx, pRx, pRn, and CoeffCond conform to approximately normal or skewed distributions, and the peak value is prominent; and (3) compared with the accuracy before sensitivity analysis and a Bayesian method, the biomass simulation accuracy of the stand model was increased by 13.92%, and all indicators show that the accuracy of the improved model is superior. This method can be used to calibrate the parameters and analyze the uncertainty of multi-parameter complex stand growth models, which are important for the improvement of parameter estimation and simulation accuracy. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Map of the study area.</p>
Full article ">Figure 2
<p>Flowchart of the methodology followed in this study. 3PG: Physiological Principles in Predicting Growth model; MCMC: Markov chain Monte Carlo method; stemNo: stand stocking; WF: foliage biomass; WR: root biomass; WS: stem biomass including branches and bark; LAI: canopy leaf area index (LAI); standvol: stand volume excluding branches and bark; MAI: mean annual volume increment; avDBH: stand-based mean diameter at breast height (DBH).</p>
Full article ">Figure 3
<p>The effects of the adjustment of model parameters on their sensitivity to stand stocking (stemNo) and diameter at breast height (DBH).</p>
Full article ">Figure 4
<p>Sensitivity grades of model parameters to stand stocking and DBH.</p>
Full article ">Figure 5
<p>The posterior distributions of seven parameters. Note: The x-axis represents the value of the parameter and the y-axis represents the probability density.</p>
Full article ">Figure 6
<p>Comparison of stand biomass before and after adjustment. Note: the red dots represent the observational data of <span class="html-italic">Cunninghamia lanceolata</span> trees from the Shunchang Forest Farm, the thin line represents the predicted stand biomass value (according to the observational data, the 3PG model can automatically generate a prediction curve), and the thick line represents the simulated stand biomass value.</p>
Full article ">
15 pages, 2462 KiB  
Article
Using Inverse Reinforcement Learning with Real Trajectories to Get More Trustworthy Pedestrian Simulations
by Francisco Martinez-Gil, Miguel Lozano, Ignacio García-Fernández, Pau Romero, Dolors Serra and Rafael Sebastián
Mathematics 2020, 8(9), 1479; https://doi.org/10.3390/math8091479 - 2 Sep 2020
Cited by 15 | Viewed by 2960
Abstract
Reinforcement learning is one of the most promising machine learning techniques to get intelligent behaviors for embodied agents in simulations. The output of the classic Temporal Difference family of Reinforcement Learning algorithms adopts the form of a value function expressed as a numeric [...] Read more.
Reinforcement learning is one of the most promising machine learning techniques to get intelligent behaviors for embodied agents in simulations. The output of the classic Temporal Difference family of Reinforcement Learning algorithms adopts the form of a value function expressed as a numeric table or a function approximator. The learned behavior is then derived using a greedy policy with respect to this value function. Nevertheless, sometimes the learned policy does not meet expectations, and the task of authoring is difficult and unsafe because the modification of one value or parameter in the learned value function has unpredictable consequences in the space of the policies it represents. This invalidates direct manipulation of the learned value function as a method to modify the derived behaviors. In this paper, we propose the use of Inverse Reinforcement Learning to incorporate real behavior traces in the learning process to shape the learned behaviors, thus increasing their trustworthiness (in terms of conformance to reality). To do so, we adapt the Inverse Reinforcement Learning framework to the navigation problem domain. Specifically, we use Soft Q-learning, an algorithm based on the maximum causal entropy principle, with MARL-Ped (a Reinforcement Learning-based pedestrian simulator) to include information from trajectories of real pedestrians in the process of learning how to navigate inside a virtual 3D space that represents the real environment. A comparison with the behaviors learned using a Reinforcement Learning classic algorithm (Sarsa(λ)) shows that the Inverse Reinforcement Learning behaviors adjust significantly better to the real trajectories. Full article
Show Figures

Figure 1

Figure 1
<p>State space features.</p>
Full article ">Figure 2
<p>Prototype extraction using K-means. The set of prototypes (in black) constitutes a generalization of the observed states (in blue). The features of the states are mainly distances and velocities which represent a dynamic situation.</p>
Full article ">Figure 3
<p>(<b>Left</b>) Real environment. The pole marks the position of the goal. The pedestrian is placed initially opposite to the pole after the farthest wall in the picture. (<b>Right</b>) Several sampled real trajectories displayed in the 3D environment.</p>
Full article ">Figure 4
<p>Sequence of images of the learned trajectory using a standard RL process with Sarsa(<math display="inline"><semantics> <mi>λ</mi> </semantics></math>) algorithm. Sarsa is a standard RL algorithm. The thick (red) line shows the learned trajectory.</p>
Full article ">Figure 5
<p>Learning curve for the last iteration of the schema.</p>
Full article ">Figure 6
<p>Sequence of images of the learned trajectory using the IRL schema. The thick (red) line is the learned trajectory.</p>
Full article ">Figure 7
<p>Curves for vectors <math display="inline"><semantics> <mover accent="true"> <msub> <mi>f</mi> <mi>ς</mi> </msub> <mo>→</mo> </mover> </semantics></math> and the last iteration <math display="inline"><semantics> <mover accent="true"> <msub> <mi>f</mi> <mi>π</mi> </msub> <mo>→</mo> </mover> </semantics></math> after being pre-processed by Procrustes analysis. Note the similarity of both curves.</p>
Full article ">
20 pages, 1677 KiB  
Article
Magnetized Particle Motion around Black Holes in Conformal Gravity: Can Magnetic Interaction Mimic Spin of Black Holes?
by Kamoliddin Haydarov, Ahmadjon Abdujabbarov, Javlon Rayimbaev and Bobomurat Ahmedov
Universe 2020, 6(3), 44; https://doi.org/10.3390/universe6030044 - 17 Mar 2020
Cited by 28 | Viewed by 2764
Abstract
Magnetized particle motion around black holes in conformal gravity immersed in asymptotically uniform magnetic field has been studied. We have also analyzed the behavior of magnetic fields near the horizon of the black hole in conformal gravity and shown that with the increase [...] Read more.
Magnetized particle motion around black holes in conformal gravity immersed in asymptotically uniform magnetic field has been studied. We have also analyzed the behavior of magnetic fields near the horizon of the black hole in conformal gravity and shown that with the increase of conformal parameters L and N the value of angular component of magnetic field at the stellar surface decreases. The maximum value of the effective potential corresponding to circular motion of the magnetized particle increases with the increase of conformal parameters. It is shown that in all cases of neutral, charged and magnetized particle collisions in the black hole environment the center-of-mass energy decreases with the increase of conformal parameters L and N. In the case of the magnetized and negatively charged particle collisions, the innermost collision point with the maximum center-of-mass energy comes closer to the central object due to the effects of the parameters of the conformal gravity. We have applied the results to the real astrophysical scenario when a pulsar treated as a magnetized particle is orbiting the super massive black hole (SMBH) Sgr A* in the center of our galaxy in order to obtain the estimation of magnetized compact object’s orbital parameter. The possible detection of pulsar in Sgr A* close environment can provide constraints on black hole parameters. Here we have shown that there is degeneracy between spin of SMBH and ambient magnetic field and consequently the interaction of magnetic field 10 2 Gauss with magnetic moment of magnetized neutron star can in principle mimic spin of Kerr black holes up to 0.6 . Full article
(This article belongs to the Special Issue Relativistic Astrophysics)
Show Figures

Figure 1

Figure 1
<p>Radial dependence of normalized angular component of magnetic field for different angles with different values of conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. The <b>top left panel</b> is for <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math>, the <b>top right one</b> is for <math display="inline"><semantics> <mrow> <mi>π</mi> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math> and the <b>bottom one</b> is for <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>6</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Profiles of the magnetic field lines for the different values of the conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span> on <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> <mi>x</mi> </mrow> </semantics></math> plane (where new coordinates are defined as <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mi>r</mi> <mo form="prefix">cos</mo> <mi>θ</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mi>r</mi> <mo form="prefix">sin</mo> <mi>θ</mi> </mrow> </semantics></math>).</p>
Full article ">Figure 3
<p>Radial dependence of magnetic coupling function for the different values of conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. The <b>top-left panel</b> corresponds to the Schwarzschild black hole (<math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>) and in the <b>top-right</b> and <b>bottom panels</b> correspond to the cases of effects of conformal gravity for the different values of angular momentum of magnetized particles when the energy is fixed as <math display="inline"><semantics> <mrow> <mi mathvariant="script">E</mi> <mo>=</mo> <msqrt> <mrow> <mn>0.9</mn> </mrow> </msqrt> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Radial dependence of the minimum energy of the particle for the different values of conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span> with the specific angular momentum <math display="inline"><semantics> <mrow> <msup> <mi>l</mi> <mn>2</mn> </msup> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math>. The value of minimum energy decreases with the increase of conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span> near the horizon.</p>
Full article ">Figure 5
<p>The radial dependence of the minimum value of parameter <math display="inline"><semantics> <mi>β</mi> </semantics></math> for the different values of the conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. The colored region represents the range for the values of <math display="inline"><semantics> <mi>β</mi> </semantics></math> in which stable circular orbits are allowed for magnetized particles for <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>&gt;</mo> <mn>3</mn> <mi>M</mi> </mrow> </semantics></math>. The values of upper and lower limits of the parameter <math display="inline"><semantics> <msub> <mi>β</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </semantics></math> decrease with the increase of conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. In <b>top-left panel</b> the blue and the light-red colored areas correspond to the case of Schwarzschild black hole and the values of the conformal parameters <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>0.4</mn> <mi>M</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, respectively. In <b>top-right</b> and <b>bottom panels</b> the blue and the light-red colored areas correspond to the values of the conformal parameter <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>0.4</mn> <mi>M</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>0.9</mn> <mi>M</mi> </mrow> </semantics></math> for the fixed values of the parameter <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, respectively.</p>
Full article ">Figure 6
<p>Radial dependence of center-of-mass energy of two colliding magnetized particles near the black hole in conformal gravity. The plots have been produced for the values of <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mo>−</mo> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">E</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi mathvariant="script">E</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>β</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>. Conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span> cause to decrease of the value of the center-of-mass energy of the colliding particles.</p>
Full article ">Figure 7
<p>Radial dependence of center-of-mass energy of collision of charged and magnetized particles for the different values of conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. The plots have been taken for the values of <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mo>−</mo> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">E</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi mathvariant="script">E</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>ω</mi> <mo>=</mo> <mo>±</mo> <mn>0.5</mn> </mrow> </semantics></math>. In the case of collision of the negative charged and magnetized particles value of the center-of-mass energy decreases with increasing of the values of the parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. The <b>top</b> and <b>bottom</b> panels show the collision of magnetized particle with negatively and positively charged particles, respectively.</p>
Full article ">Figure 8
<p>Radial dependence of center-of-mass energy of colliding neutral and magnetized particles for the different values of conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. The plots are taken for the values of <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mo>−</mo> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">E</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi mathvariant="script">E</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>. The center-of-mass energy decreases with the increase of the conformal parameters.</p>
Full article ">Figure 9
<p>Radial dependence of center-of-mass energy of two colliding charged particles for the different values of conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. The plots are taken for the values of <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>=</mo> <mo>−</mo> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">E</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi mathvariant="script">E</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>|</mo> <mi>ω</mi> <mo>|</mo> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>. The energy decreases with increasing the values of both conformal parameters <span class="html-italic">L</span> and <span class="html-italic">N</span>. <b>Left-top panel</b> shows the both positive and negative charged particles around Schwarzschild black hole (<math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>). <b>Top-right</b> and <b>bottom panels</b> correspond to the cases when the value of the parameter <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> at the fixed values of the parameter <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mi>M</mi> </mrow> </semantics></math>, respectively.</p>
Full article ">Figure 10
<p>Comparisons of ISCO radius of the neutral particles around rotating Kerr black holes with one for magnetized particles around Schwarzschild black holes immersed in external magnetic field. Red dashed line corresponds to the dependence of ISCO radius of magnetized particle on magnetic parameter <math display="inline"><semantics> <mi>β</mi> </semantics></math> while blue solid line corresponds to the dependence of test particles on rotation parameter <span class="html-italic">a</span>.</p>
Full article ">Figure 11
<p>The degeneracy plot showing the dependence of rotation parameter <span class="html-italic">a</span> on <math display="inline"><semantics> <mi>β</mi> </semantics></math>. The line corresponds to the matching values of parameters for the same values of ISCO.</p>
Full article ">
27 pages, 336 KiB  
Article
Dynamic Maximum Entropy Reduction
by Václav Klika, Michal Pavelka, Petr Vágner and Miroslav Grmela
Entropy 2019, 21(7), 715; https://doi.org/10.3390/e21070715 - 22 Jul 2019
Cited by 24 | Viewed by 5302
Abstract
Any physical system can be regarded on different levels of description varying by how detailed the description is. We propose a method called Dynamic MaxEnt (DynMaxEnt) that provides a passage from the more detailed evolution equations to equations for the less detailed state [...] Read more.
Any physical system can be regarded on different levels of description varying by how detailed the description is. We propose a method called Dynamic MaxEnt (DynMaxEnt) that provides a passage from the more detailed evolution equations to equations for the less detailed state variables. The method is based on explicit recognition of the state and conjugate variables, which can relax towards the respective quasi-equilibria in different ways. Detailed state variables are reduced using the usual principle of maximum entropy (MaxEnt), whereas relaxation of conjugate variables guarantees that the reduced equations are closed. Moreover, an infinite chain of consecutive DynMaxEnt approximations can be constructed. The method is demonstrated on a particle with friction, complex fluids (equipped with conformation and Reynolds stress tensors), hyperbolic heat conduction and magnetohydrodynamics. Full article
(This article belongs to the Special Issue Entropy and Non-Equilibrium Statistical Mechanics)
Show Figures

Figure 1

Figure 1
<p>A summary of static MaxEnt highlighting relations between state variables on the higher level and the lower level of description and their conjugates. MaxEnt provides lower entropy <math display="inline"><semantics> <mrow> <mrow> <msup> <mrow/> <mo>↓</mo> </msup> <mi>S</mi> </mrow> <mrow> <mo>(</mo> <mi mathvariant="bold">y</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> and a relation <math display="inline"><semantics> <mrow> <mi mathvariant="bold">x</mi> <mo>=</mo> <msup> <mi>π</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <mrow> <mo>(</mo> <mi mathvariant="bold">y</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> from composition of <math display="inline"><semantics> <mrow> <mover accent="true"> <mi mathvariant="bold">x</mi> <mo stretchy="false">˜</mo> </mover> <mrow> <mo>(</mo> <msup> <mi mathvariant="bold">y</mi> <mo>*</mo> </msup> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mover accent="true"> <msup> <mi mathvariant="bold">y</mi> <mo>*</mo> </msup> <mo stretchy="false">˜</mo> </mover> <mrow> <mo>(</mo> <mi mathvariant="bold">y</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>. LT denotes a relation via Legendre transformation, <math display="inline"><semantics> <mi>π</mi> </semantics></math> stands for a projection from the microscale to the macroscale variables and by an arrow we depict a mapping (written above or below the arrow) that relates the variables in the connected nodes.</p>
Full article ">
11 pages, 3066 KiB  
Article
Maximum Entropy Method for Solving the Turbulent Channel Flow Problem
by T.-W. Lee
Entropy 2019, 21(7), 675; https://doi.org/10.3390/e21070675 - 11 Jul 2019
Cited by 5 | Viewed by 3038
Abstract
There are two components in this work that allow for solutions of the turbulent channel flow problem: One is the Galilean-transformed Navier-Stokes equation which gives a theoretical expression for the Reynolds stress (u′v′); and the second the maximum entropy principle which [...] Read more.
There are two components in this work that allow for solutions of the turbulent channel flow problem: One is the Galilean-transformed Navier-Stokes equation which gives a theoretical expression for the Reynolds stress (u′v′); and the second the maximum entropy principle which provides the spatial distribution of turbulent kinetic energy. The first concept transforms the momentum balance for a control volume moving at the local mean velocity, breaking the momentum exchange down to its basic components, u′v′, u′2, pressure and viscous forces. The Reynolds stress gradient budget confirms this alternative interpretation of the turbulence momentum balance, as validated with DNS data. The second concept of maximum entropy principle states that turbulent kinetic energy in fully-developed flows will distribute itself until the maximum entropy is attained while conforming to the physical constraints. By equating the maximum entropy state with maximum allowable (viscous) dissipation at a given Reynolds number, along with other constraints, we arrive at function forms (inner and outer) for the turbulent kinetic energy. This allows us to compute the Reynolds stress, then integrate it to obtain the velocity profiles in channel flows. The results agree well with direct numerical simulation (DNS) data at Reτ = 400 and 1000. Full article
(This article belongs to the Section Multidisciplinary Applications)
Show Figures

Figure 1

Figure 1
<p>Reynolds stress gradient budget. DNS channel flow data (circle symbol) for Re<sub>τ</sub> = 1000 [<a href="#B9-entropy-21-00675" class="html-bibr">9</a>] are used. Bold line is the RHS side of Equation (2), with <span class="html-italic">u</span><sup>2</sup>-transport, pressure and the viscous terms combined.</p>
Full article ">Figure 2
<p><span class="html-italic">u</span>′<sup>2</sup> profile obtained from Equation (3). Reynolds stress gradient budget. DNS channel flow data (circle symbol) for Re<sub>τ</sub> = 1000 [<a href="#B9-entropy-21-00675" class="html-bibr">9</a>] are used. Bold line is the RHS side of Equation (3), with <span class="html-italic">u</span>′<span class="html-italic">v</span>′-transport, pressure and the viscous terms combined.</p>
Full article ">Figure 3
<p>Total integrated <span class="html-italic">u</span>′<sup>2</sup>(<span class="html-italic">E</span>) and dissipation (<span class="html-italic">ε</span>) as a function of the Reynolds numbers, from the DNS data [<a href="#B9-entropy-21-00675" class="html-bibr">9</a>,<a href="#B11-entropy-21-00675" class="html-bibr">11</a>].</p>
Full article ">Figure 4
<p>The location of the <span class="html-italic">u</span>′<sup>2</sup> peak as a function of the Reynolds numbers, from the DNS data [<a href="#B9-entropy-21-00675" class="html-bibr">9</a>,<a href="#B11-entropy-21-00675" class="html-bibr">11</a>].</p>
Full article ">Figure 5
<p><span class="html-italic">u</span>′<sup>2</sup> profile as a combination of lognormal (inner) and beta (outer) functions. Symbols are the DNS data [<a href="#B9-entropy-21-00675" class="html-bibr">9</a>,<a href="#B11-entropy-21-00675" class="html-bibr">11</a>].</p>
Full article ">Figure 6
<p><span class="html-italic">v</span>′<sup>2</sup> and <span class="html-italic">w</span>′<sup>2</sup> profiles and lognormal functions. Symbols are the DNS data [<a href="#B9-entropy-21-00675" class="html-bibr">9</a>].</p>
Full article ">Figure 7
<p>The Reynolds stress profiles computed using Equation (2), compared with DNS data [<a href="#B9-entropy-21-00675" class="html-bibr">9</a>,<a href="#B11-entropy-21-00675" class="html-bibr">11</a>].</p>
Full article ">Figure 8
<p>The mean velocity inner and outer solutions, compared with DNS data [<a href="#B9-entropy-21-00675" class="html-bibr">9</a>,<a href="#B11-entropy-21-00675" class="html-bibr">11</a>].</p>
Full article ">Figure A1
<p>Illustration of the displacement effect leading to d/dx → C<sub>1</sub>Ud/dy transform.</p>
Full article ">Figure A2
<p>Off-set line of motion for the control volume, for “probing” the <span class="html-italic">d</span>/<span class="html-italic">dy</span> gradient.</p>
Full article ">
16 pages, 2934 KiB  
Article
The Conformal Design of an Island-Bridge Structure on a Non-Developable Surface for Stretchable Electronics
by Lin Xiao, Chen Zhu, Wennan Xiong, YongAn Huang and Zhouping Yin
Micromachines 2018, 9(8), 392; https://doi.org/10.3390/mi9080392 - 7 Aug 2018
Cited by 34 | Viewed by 7448
Abstract
Conformal design of the island-bridge structure is the key to construct high-performance inorganic stretchable electronics that can be conformally transferred to non-developable surfaces. Former studies in conformal problems of epidermal electronics are mainly focused on soft surfaces that can adapt to the deformation [...] Read more.
Conformal design of the island-bridge structure is the key to construct high-performance inorganic stretchable electronics that can be conformally transferred to non-developable surfaces. Former studies in conformal problems of epidermal electronics are mainly focused on soft surfaces that can adapt to the deformation of the electronics, which are not suitable for applications in hard, non-developable surfaces because of their loose surface constraints. In this paper, the conformal design problem for the island-bridge structure on a hard, non-developable surface was studied, including the critical size for island and stiffness and the demand for stretchability for the bridge. Firstly, the conformal model for an island on a part of torus surface was established to determine the relationship between the maximum size of the island and the curvatures of the surface. By combining the principle of energy minimization and the limit of material failure, a critical non-dimensional width for conformability was given for the island as a function of its thickness and interfacial adhesion energy, and the ratio of two principal curvatures of the surface. Then, the dependency of the tensile stiffness of the bridge on its geometric parameters was studied by finite element analysis (FEA) to guide the deterministic assembly of the islands on the surface. Finally, the location-dependent demands for the stretchability of the bridges were given by geometric mapping. This work will provide a design rule for stretchable electronics that fully conforms to the non-developable surface. Full article
(This article belongs to the Special Issue Flexible Electronics: Fabrication and Ubiquitous Integration)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) An island-bridge structure array on a non-developable surface; (<b>b</b>) theory model of island on a torus surface under control by two principal curvatures, <span class="html-italic">κ</span><sub>1</sub> and <span class="html-italic">κ</span><sub>2</sub>; (<b>c</b>) schematic of geometric parameters for a serpentine bridge with <span class="html-italic">m</span> unit cells; (<b>d</b>) a numbered island-bridge structure array with <span class="html-italic">m</span> rows and <span class="html-italic">n</span> columns of islands.</p>
Full article ">Figure 2
<p>Strain and strain energy in the island during conformal contact: (<b>a</b>) maximum strain in island with non-dimensional width <span class="html-italic">κ</span><sub>2</sub><span class="html-italic">w</span><sub>island</sub> at <span class="html-italic">κ</span><sub>2</sub><span class="html-italic">t</span><sub>island</sub> = 10<sup>−6</sup>; (<b>b</b>) the ratio of stretching strain energy to bending strain energy with non-dimensional parameter <span class="html-italic">η</span>.</p>
Full article ">Figure 3
<p>The non-dimensional critical conformal width <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>2</mn> </msub> <msub> <mi>w</mi> <mrow> <mi>critical</mi> </mrow> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mi>ξ</mi> </semantics></math> for <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mrow> <mi>critical</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>ν</mi> <mrow> <mi>island</mi> </mrow> </msub> <mo>=</mo> <mn>0.32</mn> </mrow> </semantics></math>. Two regions, weak adhesion and strong adhesion, are divided by <math display="inline"><semantics> <mrow> <msub> <mi>ξ</mi> <mrow> <mi>critical</mi> </mrow> </msub> <mo>=</mo> <mn>0.56</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>The conformal behaviors between sphere and PVC islands with different width: (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>2</mn> </msub> <msub> <mi>w</mi> <mrow> <mi>critical</mi> </mrow> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>2</mn> </msub> <msub> <mi>w</mi> <mrow> <mi>critical</mi> </mrow> </msub> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>2</mn> </msub> <msub> <mi>w</mi> <mrow> <mi>critical</mi> </mrow> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>2</mn> </msub> <msub> <mi>w</mi> <mrow> <mi>critical</mi> </mrow> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>2</mn> </msub> <msub> <mi>w</mi> <mrow> <mi>critical</mi> </mrow> </msub> <mo>=</mo> <mn>0.6</mn> </mrow> </semantics></math>, and (<b>f</b>) enlarge view of wrinkle in (<b>c</b>).</p>
Full article ">Figure 5
<p>The axial force of a serpentine interconnect under stretching, obtained from the finite element analysis with different parameters: (<b>a</b>) applying strain, (<b>b</b>) wave numbers of bridge, (<b>c</b>) thickness of bridge, (<b>d</b>) width of bridge.</p>
Full article ">Figure 6
<p>The max principal strain in the bridge versus the applied strain, and the corresponding deformation configurations in xy and yz viewport: (<b>a</b>) <span class="html-italic">ε</span><sub>appl</sub> = 21%, (<b>b</b>) <span class="html-italic">ε</span><sub>appl</sub> = 22%, (<b>c</b>) <span class="html-italic">ε</span><sub>appl</sub> = 40%, (<b>d</b>) <span class="html-italic">ε</span><sub>appl</sub> = 60%, (<b>e</b>) <span class="html-italic">ε</span><sub>appl</sub> = 80%, and (<b>f</b>) <span class="html-italic">ε</span><sub>appl</sub> = 100%.</p>
Full article ">Figure 7
<p>Demands for stretchability of the bridges given by geometric method: location-dependent property of demands for stretchability in the array (<b>a</b>) and at the first row (<b>b</b>) for the horizontal bridges; maximum demand for stretchability of the device with the number of islands (<b>c</b>) and area coverage (<b>d</b>).</p>
Full article ">Figure A1
<p>The schematic for island on surface with initial angle <span class="html-italic">θ</span>.</p>
Full article ">Figure A2
<p>The conformal strain energy per unit area in island with initial angle <span class="html-italic">θ</span> at different width-length ratios (<b>a</b>) and different curvature ratios (<b>b</b>).</p>
Full article ">Figure A3
<p>The relative error between approximate solution and accurate solution with <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>2</mn> </msub> <msub> <mi>w</mi> <mrow> <mi>island</mi> </mrow> </msub> </mrow> </semantics></math>: (<b>a</b>) for strain and (<b>b</b>) for conformal strain energy.</p>
Full article ">Figure A4
<p>Experimental apparatus and experimental data for material parameter test: (<b>a</b>) an universal mechanical tester (INSTRON 5944); (<b>b</b>) a home-made peel platform with an angle-adjustable jig, in which the X or Z-motion of the translation stage is able to be driven by two independent linear/electric actuators, and its Y-motion depends on a manually single axis table; (<b>c</b>) stress-strain curve for PVC sticker in tension test; and (<b>d</b>) peel force for PVC sticker in peel test with peel angle <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <msup> <mrow> <mn>135</mn> </mrow> <mo>∘</mo> </msup> </mrow> </semantics></math> and peel rate <math display="inline"><semantics> <mrow> <msub> <mi>v</mi> <mrow> <mi>peel</mi> </mrow> </msub> <mo>=</mo> <mrow> <mrow> <mn>1</mn> <mo> </mo> <mi>mm</mi> </mrow> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">
14 pages, 8596 KiB  
Article
Kurchatovite and Clinokurchatovite, Ideally CaMgB2O5: An Example of Modular Polymorphism
by Yulia A. Pankova, Sergey V. Krivovichev, Igor V. Pekov, Edward S. Grew and Vasiliy O. Yapaskurt
Minerals 2018, 8(8), 332; https://doi.org/10.3390/min8080332 - 2 Aug 2018
Cited by 5 | Viewed by 3503
Abstract
Kurchatovite and clinokurchatovite, both of ideal composition CaMgB2O5, from the type localities (Solongo, Buryatia, Russia, and Sayak-IV, Kazakhstan, respectively) have been studied using electron microprobe and single-crystal X-ray diffraction methods. The empirical formulae of the samples are Ca1.01 [...] Read more.
Kurchatovite and clinokurchatovite, both of ideal composition CaMgB2O5, from the type localities (Solongo, Buryatia, Russia, and Sayak-IV, Kazakhstan, respectively) have been studied using electron microprobe and single-crystal X-ray diffraction methods. The empirical formulae of the samples are Ca1.01Mg0.87Mn0.11Fe2+0.02B1.99O5 and Ca0.94Mg0.91Fe2+0.10Mn0.04B2.01O5 for kurchatovite and clinokurchatovite, respectively. The crystal structures of the two minerals are similar and based upon two-dimensional blocks arranged parallel to the c axis in kurchatovite and parallel to the a axis in clinokurchatovite. The blocks are built up from diborate B2O5 groups, and Ca2+ and Mg2+ cations in seven- and six-fold coordination, respectively. Detailed analysis of geometrical parameters of the adjacent blocks reveals that symmetrically different diborate groups have different degrees of conformation in terms of the δ angles between the planes of two BO3 triangles sharing a common O atom, featuring two discrete sets of the δ values of ca. 55° (B’ blocks) and 34° (B” blocks). The stacking of the blocks in clinokurchatovite can be presented as …(+B’)(+B”)(+B’)(+B”)… or [(+B’)(+B”)], whereas in kurchatovite it is more complex and corresponds to the sequence …(+B’)(+B”)(+B’)(−B’)(−B”)(−B’)(+B’)(+B”)(+B’)(−B’)(−B”)(−B’)… or [(+B’)(+B”)(+B’)(−B’)(−B”)(−B’)]. The B’:B” ratios for clinokurchatovite and kurchatovite are 1:1 and 2:1, respectively. According to this description, the two minerals cannot be considered as polytypes and their mutual relationship corresponds to the term modular polymorphs. From the viewpoint of information-based measures of structural complexity, clinokurchatovite (IG = 4.170 bits/atom and IG,total = 300.235 bits/cell) is structurally simpler than kurchatovite (IG = 4.755 bits/atom and IG,total = 1027.056 bits/cell). The high structural complexity of kurchatovite can be inferred from the modular character of its structure. The analysis of structural combinatorics in terms of the modular approach allows to construct the whole family of theoretically possible “kurchatovite”-type structures that bear the same structural features common for kurchatovite and clinokurchatovite. However, the crystal structures of the latter minerals are the simplest and are the only ones that have been observed in nature. The absence of other possible structures is remarkable and can be explained by either the maximum-entropy of the least-action fundamental principles. Full article
(This article belongs to the Special Issue New Mineral Species and Their Crystal Structures)
Show Figures

Figure 1

Figure 1
<p>The crystal structure of kurchatovite projected along the <span class="html-italic">b</span> axis (<b>a</b>) and the arrangement of B<sub>2</sub>O<sub>5</sub> groups (<b>b</b>). The curved brackets in (<b>a</b>) denote the basic two-dimensional blocks, whereas the dashed lines in (<b>b</b>) indicate directions of division of the structure into single slices consisting of diborate groups.</p>
Full article ">Figure 2
<p>The crystal structure of clinokurchatovite projected along the <span class="html-italic">c</span> axis (<b>a</b>) and the arrangement of B<sub>2</sub>O<sub>5</sub> groups (<b>b</b>). The dashed lines in (<b>b</b>) indicate directions of division of the structure into single slices consisting of diborate groups.</p>
Full article ">Figure 3
<p>The structure topology of the 6 Å-block in the crystal structures of kurchatovite and clinokurchatovite (<b>a</b>; the picture corresponds to one of the symmetrically independent blocks in kurchatovite) and the coordination figures of Ca (<b>b</b>) and Mg (<b>c</b>) atoms.</p>
Full article ">Figure 4
<p>The slices of the crystal structures of clinokurchatovite (<b>a</b>,<b>b</b>) and kurchatovite (<b>c</b>,<b>d</b>) and consisting of diborate groups (divided along the dashed lines as shown in <a href="#minerals-08-00332-f001" class="html-fig">Figure 1</a> and <a href="#minerals-08-00332-f002" class="html-fig">Figure 2</a>). The dot-and-dash lines in (<b>a</b>,<b>c</b>) indicate the relative positions of the <b>B’</b> and <b>B”</b> of different orientations (distinguished by the “<b>+</b>” and “<b>−</b>” signs).</p>
Full article ">Figure 5
<p>The arrangements of diborate groups in the crystal structures of kurchatovite (<b>a</b>) and clinokurchatovite (<b>b</b>), featuring their angular characteristics (the B–O–B angles are shown in black; the values in blue boxes correspond to the δ angles between the planes of the adjacent BO<sub>3</sub> triangles).</p>
Full article ">Figure 6
<p>The illustration of the rule of overlapping triplets using the example of the <b>I–V</b> pair (<b>a</b>), the diagram of relations between the triplets (<b>b</b>), and the parts of diagram that result in the generation of clinokurchatovite (<b>c</b>) and kurchatovite (<b>d</b>) structures. See text for details.</p>
Full article ">
258 KiB  
Article
Conjugate Representations and Characterizing Escort Expectations in Information Geometry
by Tatsuaki Wada and Hiroshi Matsuzoe
Entropy 2017, 19(7), 309; https://doi.org/10.3390/e19070309 - 28 Jun 2017
Cited by 4 | Viewed by 3417
Abstract
Based on the maximum entropy (MaxEnt) principle for a generalized entropy functional and the conjugate representations introduced by Zhang, we have reformulated the method of information geometry. For a set of conjugate representations, the associated escort expectation is naturally introduced and characterized by [...] Read more.
Based on the maximum entropy (MaxEnt) principle for a generalized entropy functional and the conjugate representations introduced by Zhang, we have reformulated the method of information geometry. For a set of conjugate representations, the associated escort expectation is naturally introduced and characterized by the generalized score function which has zero-escort expectation. Furthermore, we show that the escort expectation induces a conformal divergence. Full article
(This article belongs to the Special Issue Information Geometry II)
286 KiB  
Article
Molecular Orbital and Density Functional Study of the Formation, Charge Transfer, Bonding and the Conformational Isomerism of the Boron Trifluoride (BF3) and Ammonia (NH3) Donor-Acceptor Complex
by Dulal C. Ghosh and Soma Bhattacharyya
Int. J. Mol. Sci. 2004, 5(8), 239-264; https://doi.org/10.3390/i5050239 - 30 Sep 2004
Cited by 46 | Viewed by 11570
Abstract
The formation of the F3B–NH3 supermolecule by chemical interaction of its fragment parts, BF3 and NH3, and the dynamics of internal rotation about the ‘B–N’ bond have been studied in terms of parameters provided by the molecular [...] Read more.
The formation of the F3B–NH3 supermolecule by chemical interaction of its fragment parts, BF3 and NH3, and the dynamics of internal rotation about the ‘B–N’ bond have been studied in terms of parameters provided by the molecular orbital and density functional theories. It is found that the pairs of frontier orbitals of the interacting fragments have matching symmetry and are involved in the charge transfer interaction. The donation process stems from the HOMO of the donor into the LUMO of the acceptor and simultaneously, back donation stems from the HOMO of acceptor into the LUMO of the donor. The density functional computation of chemical activation in the donor and acceptor fragments, associated with the physical process of structural reorganization just prior to the event of chemical reaction, indicates that BF3 becomes more acidic and NH3 becomes more basic, compared to their separate equilibrium states. Theoretically it is observed that the chemical reaction event of the formation of the supermolecule from its fragment parts is in accordance with the chemical potential equalization principle of the density functional theory and the electronegativity equalization principle of Sanderson. The energetics of the chemical reaction, the magnitude of the net charge transfer and the energy of the newly formed bond are quite consistent, both internally and with the principle of maximum hardness, PMH. The dynamics of the internal rotation of one part with respect to the other part of the supermolecule about the ‘B–N’ bond mimics the pattern of the conformational isomerism of the isostructural ethane molecule. It is also observed that the dynamics and evolution of molecular conformations as a function of dihedral angles is also in accordance with the principle of maximum hardness, PMH. Quite consistent with spectroscopic predictions, the height of the molecule’s barrier to internal rotation is very small. A rationale for the low height of the barrier has been put forward in terms of the energy partitioning analysis. On the question of origin of the barrier to internal rotation, we conclude that the conformational barrier to internal rotation does not originate from a particular region of the molecule, but rather it is a result of the subtle conjoint interplay of a number of opposing effects of one- and two-center bonded and nonbonded energy terms involving the entire skeleton of the molecule. Full article
Show Figures

Figure 1

Figure 1
<p>The intuitive structure and dynamics of the structural reorganization prior to the event of chemical reaction of BX<sub>3</sub> with the donor (L) molecules.</p>
Full article ">Figure 2
<p>Plot of total energy of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angles</p>
Full article ">Figure 3
<p>Plot of gross atomic charge densities on N and H atoms in F3B-NH3 system as a function of torsional angles</p>
Full article ">Figure 4
<p>Plot of gross atomic charge densities on B and F atoms of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angles</p>
Full article ">Figure 5
<p>Plot of Global hardness of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angles (E-Eclipsed, S-Staggered)</p>
Full article ">Figure 6
<p>Plot of 'B-N' bonded interaction energy of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angle.</p>
Full article ">Figure 7
<p>Plot of 'N-H' and 'B-F' two center bonded interaction energies of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angle (E-Eclipsed, S-Staggered)</p>
Full article ">Figure 8
<p>Plot of 'H.···F' non-bonded interaction of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angle (E-Eclipsed, S-Staggered)</p>
Full article ">Figure 9
<p>Plot of two-center non-bonded interaction energy of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angle.</p>
Full article ">Figure 10
<p>Plot of two-center 'F···F' non-bonded interaction energy of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angles (E-Eclipsed, S-Staggered)</p>
Full article ">Figure 11
<p>Plot of one-center energy (on N and F) of F<sub>3</sub>B-NH<sub>3</sub> system as a function torsional angle (E-Eclipsed, S-Staggered)</p>
Full article ">Figure 12
<p>Plot of one-center energy on 'H' of F<sub>3</sub>B-NH<sub>3</sub> system as a function of torsional angle.</p>
Full article ">
Back to TopTop