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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (24)

Search Parameters:
Keywords = Helmholtz decomposition

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 1422 KiB  
Article
Limitations and Performance Analysis of Spherical Sector Harmonics for Sound Field Processing
by Hanwen Bi, Shaoheng Xu, Fei Ma, Thushara D. Abhayapala and Prasanga N. Samarasinghe
Appl. Sci. 2024, 14(22), 10633; https://doi.org/10.3390/app142210633 - 18 Nov 2024
Viewed by 496
Abstract
Developing spherical sector harmonics (SSHs) benefits sound field decomposition and analysis over spherical sector regions. Although SSHs demonstrate potential in the field of spatial audio, a comprehensive investigation into their properties and performance is absent. This paper seeks to close this gap by [...] Read more.
Developing spherical sector harmonics (SSHs) benefits sound field decomposition and analysis over spherical sector regions. Although SSHs demonstrate potential in the field of spatial audio, a comprehensive investigation into their properties and performance is absent. This paper seeks to close this gap by revealing three key limitations of SSHs and exploring their performance in two aspects: sector sound field radial extrapolation and sector sound field decomposition and reconstruction. First, SSHs are not solutions to the Helmholtz equation, which is their main limitation. Then, due to the violation of the Helmholtz equation, SSHs lack the ability to conduct sound field radial extrapolation, especially for interior cases. Third, when using SSHs to decompose and reconstruct a sound field, the shifted associated Legendre polynomials and scaled exponential function in SSHs result in severe distortion around the edge of the sector region. In light of these three limitations, the future implementation of SSHs should focus on processing and analyzing the measurement sector region without any extrapolation process, and the measurement region should be larger than the target sector region. Full article
(This article belongs to the Special Issue Spatial Audio and Sound Design)
Show Figures

Figure 1

Figure 1
<p>Illustration of the coordinate system.</p>
Full article ">Figure 2
<p>Spherical sector sound field radial extrapolation setups for (<b>a</b>) exterior case and (<b>b</b>) interior case. Illustration of the sound source region <math display="inline"><semantics> <mi mathvariant="double-struck">S</mi> </semantics></math>, the measurement region <math display="inline"><semantics> <mi mathvariant="double-struck">M</mi> </semantics></math>, and the extrapolation region <math display="inline"><semantics> <mi mathvariant="double-struck">T</mi> </semantics></math>.</p>
Full article ">Figure 3
<p>Spherical sector sound field radial extrapolation for the exterior case, with <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>M</mi> </msub> <mo>=</mo> <msup> <mn>120</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>M</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>300</mn> <mspace width="3.33333pt"/> <mi>Hz</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math>: (<b>a</b>) sector sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>b</b>) estimated sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>c</b>) sector sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>5</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>d</b>) estimated sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>5</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Extrapolation error <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> for the exterior case with different settings of frequencies, sizes of <math display="inline"><semantics> <mi mathvariant="double-struck">T</mi> </semantics></math>, and distances between <math display="inline"><semantics> <msub> <mi>R</mi> <mi>T</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>R</mi> <mi>M</mi> </msub> </semantics></math>: (<b>a</b>) extrapolation error with 300 Hz; (<b>b</b>) extrapolation error with 500 Hz; (<b>c</b>) extrapolation error with 700 Hz; (<b>d</b>) extrapolation error with 900 Hz.</p>
Full article ">Figure 4 Cont.
<p>Extrapolation error <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> for the exterior case with different settings of frequencies, sizes of <math display="inline"><semantics> <mi mathvariant="double-struck">T</mi> </semantics></math>, and distances between <math display="inline"><semantics> <msub> <mi>R</mi> <mi>T</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>R</mi> <mi>M</mi> </msub> </semantics></math>: (<b>a</b>) extrapolation error with 300 Hz; (<b>b</b>) extrapolation error with 500 Hz; (<b>c</b>) extrapolation error with 700 Hz; (<b>d</b>) extrapolation error with 900 Hz.</p>
Full article ">Figure 5
<p>Spherical sector sound field radial extrapolation for the interior case with a plane wave, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>M</mi> </msub> <mo>=</mo> <msup> <mn>120</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>M</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>300</mn> <mspace width="3.33333pt"/> <mi>Hz</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math>: (<b>a</b>) true sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>b</b>) estimated sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>c</b>) true sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>0.6</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>d</b>) estimated sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>0.6</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Extrapolation error <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> for the interior case with a plane wave, different settings of frequencies, sizes of <math display="inline"><semantics> <mi mathvariant="double-struck">T</mi> </semantics></math>, and distances between <math display="inline"><semantics> <msub> <mi>R</mi> <mi>T</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>R</mi> <mi>M</mi> </msub> </semantics></math>: (<b>a</b>) extrapolation error with 300 Hz; (<b>b</b>) extrapolation error with 500 Hz; (<b>c</b>) extrapolation error with 700 Hz; (<b>d</b>) extrapolation error with 900 Hz.</p>
Full article ">Figure 7
<p>Extrapolation error <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> with different arriving directions of the plane wave: (<b>a</b>) extrapolation error with <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>M</mi> </msub> <mo>=</mo> <msup> <mn>45</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>; (<b>b</b>) extrapolation error with <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>M</mi> </msub> <mo>=</mo> <msup> <mn>75</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>; (<b>c</b>) extrapolation error with <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>M</mi> </msub> <mo>=</mo> <msup> <mn>105</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>; (<b>d</b>) extrapolation error with <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>M</mi> </msub> <mo>=</mo> <msup> <mn>135</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Spherical sector sound field radial extrapolation for the interior case with a point source, where <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>M</mi> </msub> <mo>=</mo> <msup> <mn>120</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>M</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>300</mn> <mspace width="3.33333pt"/> <mi>Hz</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math>: (<b>a</b>) true sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>b</b>) estimated sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>1</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>c</b>) true sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>0.6</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>; (<b>d</b>) estimated sound field with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>0.6</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Extrapolation error <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> for the interior case with a point source, different settings of frequencies, sizes of <math display="inline"><semantics> <mi mathvariant="double-struck">T</mi> </semantics></math>, and distances between <math display="inline"><semantics> <msub> <mi>R</mi> <mi>T</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>R</mi> <mi>M</mi> </msub> </semantics></math>: (<b>a</b>) extrapolation error with 300 Hz; (<b>b</b>) extrapolation error with 500 Hz; (<b>c</b>) extrapolation error with 700 Hz; (<b>d</b>) extrapolation error with 900 Hz.</p>
Full article ">Figure 9 Cont.
<p>Extrapolation error <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> for the interior case with a point source, different settings of frequencies, sizes of <math display="inline"><semantics> <mi mathvariant="double-struck">T</mi> </semantics></math>, and distances between <math display="inline"><semantics> <msub> <mi>R</mi> <mi>T</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>R</mi> <mi>M</mi> </msub> </semantics></math>: (<b>a</b>) extrapolation error with 300 Hz; (<b>b</b>) extrapolation error with 500 Hz; (<b>c</b>) extrapolation error with 700 Hz; (<b>d</b>) extrapolation error with 900 Hz.</p>
Full article ">Figure 10
<p>Illustration of the mapping process for both elevation angle and azimuth angle perspective: (<b>a</b>) map the target spherical sector region <math display="inline"><semantics> <mi mathvariant="double-struck">T</mi> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>∈</mo> <mrow> <mo stretchy="false">[</mo> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo stretchy="false">]</mo> </mrow> <mo>,</mo> <mi>ϕ</mi> <mo>∈</mo> <mrow> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math>) to a whole sphere <math display="inline"><semantics> <msup> <mi mathvariant="double-struck">S</mi> <mn>2</mn> </msup> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>∈</mo> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mi>π</mi> <mo stretchy="false">]</mo> <mo>,</mo> <mi>ϕ</mi> <mo>∈</mo> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>); (<b>b</b>) map the target spherical sector region <math display="inline"><semantics> <mi mathvariant="double-struck">T</mi> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>∈</mo> <mrow> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mi>π</mi> <mo stretchy="false">]</mo> </mrow> <mo>,</mo> <mi>ϕ</mi> <mo>∈</mo> <mrow> <mo stretchy="false">[</mo> <msub> <mi>ϕ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>ϕ</mi> <mn>2</mn> </msub> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math>) to a whole sphere <math display="inline"><semantics> <msup> <mi mathvariant="double-struck">S</mi> <mn>2</mn> </msup> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>∈</mo> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mi>π</mi> <mo stretchy="false">]</mo> <mo>,</mo> <mi>ϕ</mi> <mo>∈</mo> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>).</p>
Full article ">Figure 11
<p>Reconstruction error in the elevation direction <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>e</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>θ</mi> <mo>,</mo> <mi>k</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> with different <math display="inline"><semantics> <mi>θ</mi> </semantics></math> values.</p>
Full article ">Figure 12
<p>Reconstruction error in the azimuth direction <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>a</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>ϕ</mi> <mo>,</mo> <mi>k</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> with different <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math> values.</p>
Full article ">
15 pages, 744 KiB  
Article
Causal Hierarchy in the Financial Market Network—Uncovered by the Helmholtz–Hodge–Kodaira Decomposition
by Tobias Wand, Oliver Kamps and Hiroshi Iyetomi
Entropy 2024, 26(10), 858; https://doi.org/10.3390/e26100858 - 11 Oct 2024
Cited by 1 | Viewed by 867
Abstract
Granger causality can uncover the cause-and-effect relationships in financial networks. However, such networks can be convoluted and difficult to interpret, but the Helmholtz–Hodge–Kodaira decomposition can split them into rotational and gradient components which reveal the hierarchy of the Granger causality flow. Using Kenneth [...] Read more.
Granger causality can uncover the cause-and-effect relationships in financial networks. However, such networks can be convoluted and difficult to interpret, but the Helmholtz–Hodge–Kodaira decomposition can split them into rotational and gradient components which reveal the hierarchy of the Granger causality flow. Using Kenneth French’s business sector return time series, it is revealed that during the COVID crisis, precious metals and pharmaceutical products were causal drivers of the financial network. Moreover, the estimated Granger causality network shows a high connectivity during the crisis, which means that the research presented here can be especially useful for understanding crises in the market better by revealing the dominant drivers of crisis dynamics. Full article
(This article belongs to the Special Issue Complexity in Financial Networks)
Show Figures

Figure 1

Figure 1
<p>Example of the Helmholtz–Hodge–Kodaira decomposition for a single graph into a gradient-based graph (g) and a circular graph (c). Note that direction of the flux between A and C is different in (g) and (c), which is the same as changing the sign <math display="inline"><semantics> <mrow> <msubsup> <mi>J</mi> <mrow> <mi>A</mi> <mi>C</mi> </mrow> <mrow> <mo>(</mo> <mi>g</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <mo>−</mo> <msubsup> <mi>J</mi> <mrow> <mi>C</mi> <mi>A</mi> </mrow> <mrow> <mo>(</mo> <mi>g</mi> <mo>)</mo> </mrow> </msubsup> </mrow> </semantics></math>, and hence, their sum is given by <math display="inline"><semantics> <mrow> <msubsup> <mi>J</mi> <mrow> <mi>C</mi> <mi>A</mi> </mrow> <mrow> <mo>(</mo> <mi>g</mi> <mo>)</mo> </mrow> </msubsup> <mo>+</mo> <msubsup> <mi>J</mi> <mrow> <mi>C</mi> <mi>A</mi> </mrow> <mrow> <mo>(</mo> <mi>c</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <mo>−</mo> <mn>0.6</mn> <mo>+</mo> <mn>0.7</mn> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, and the original flux <math display="inline"><semantics> <msub> <mi>J</mi> <mrow> <mi>C</mi> <mi>A</mi> </mrow> </msub> </semantics></math> is reconstructed. Also, note that the total flux between two nodes is path-independent for (g) as <math display="inline"><semantics> <mrow> <msubsup> <mi>J</mi> <mrow> <mi>A</mi> <mi>C</mi> </mrow> <mrow> <mo>(</mo> <mi>g</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <msubsup> <mi>J</mi> <mrow> <mi>A</mi> <mi>B</mi> </mrow> <mrow> <mo>(</mo> <mi>g</mi> <mo>)</mo> </mrow> </msubsup> <mo>+</mo> <msubsup> <mi>J</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> <mrow> <mo>(</mo> <mi>g</mi> <mo>)</mo> </mrow> </msubsup> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>The network structure used for the vector autoregression which generates synthetic time series. One node is at the top of the hierarchy without any causal parent, whereas eight nodes are in the second layer and forty are in the final layer. Each node in the second layer is the parent node of 5 nodes in the final layer and has the node in the first layer as their parent node. Sketched via the software [<a href="#B35-entropy-26-00858" class="html-bibr">35</a>].</p>
Full article ">Figure 3
<p>Results of the RCGCI-HHKD analysis for annual data from [<a href="#B22-entropy-26-00858" class="html-bibr">22</a>]. The gray shaded area is the CI for the network connectivity of random data without any causal coupling. Note that the lines that connect the dots are only a visual aid, and no linear interpolation between the periods is assumed.</p>
Full article ">Figure 4
<p>For the analysis of annual data from 2004 to 2023, KDE of the sum of all influx and outflux of Granger causality and the total number of years with at least one inward or outward link in the RCGCI network. Values on the x-axis have been normalized to the same scale.</p>
Full article ">Figure 5
<p>For time periods of 12 months, two network measures are depicted here: the network connectivity and the gradient contribution <math display="inline"><semantics> <mi>γ</mi> </semantics></math>. The network connectivity is the percentage of sectors connected to the network and is displayed here against the random connectivity expected for independent time series. If the network is complete and has a connectivity of 1, the gradient contribution <math display="inline"><semantics> <mi>γ</mi> </semantics></math> is also calculated according to Equation (<a href="#FD10-entropy-26-00858" class="html-disp-formula">10</a>). Note that the time on the x-axis is the midpoint of the 12-month intervals of data.</p>
Full article ">Figure 6
<p>For the same time intervals as in <a href="#entropy-26-00858-f005" class="html-fig">Figure 5</a>, the potentials <math display="inline"><semantics> <msub> <mi mathvariant="normal">Φ</mi> <mi>i</mi> </msub> </semantics></math> of each sector are shown as dots. Note that for each time interval, the potentials have been centered via <math display="inline"><semantics> <mrow> <msub> <mo>∑</mo> <mi>i</mi> </msub> <msub> <mi mathvariant="normal">Φ</mi> <mi>i</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. Some selected sectors are shown in color, and the gray area shows the spread between the <math display="inline"><semantics> <mrow> <mn>25</mn> <mo>%</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>75</mn> <mo>%</mo> </mrow> </semantics></math> quantiles for each time period.</p>
Full article ">Figure 7
<p>The estimated Granger causality influence network ordered by the HHKD potentials for the periods from January 2007 to December 2007 (the sector FabPr is not shown because it has no link to any other sector) and from October 2019 to September 2020. The width of the arrows reflects the strength of the Granger causality, and selected sectors are highlighted with the same color coding as in <a href="#entropy-26-00858-f006" class="html-fig">Figure 6</a> whereas all other sectors are shown in blue.</p>
Full article ">
17 pages, 4862 KiB  
Article
Modelling and Characterisation of Orthotropic Damage in Aluminium Alloy 2024
by Nenad Djordjevic, Ravindran Sundararajah, Rade Vignjevic, James Campbell and Kevin Hughes
Materials 2024, 17(17), 4281; https://doi.org/10.3390/ma17174281 - 29 Aug 2024
Viewed by 585
Abstract
The aim of the work presented in this paper was development of a thermodynamically consistent constitutive model for orthotopic metals and determination of its parameters based on standard characterisation methods used in the aerospace industry. The model was derived with additive decomposition of [...] Read more.
The aim of the work presented in this paper was development of a thermodynamically consistent constitutive model for orthotopic metals and determination of its parameters based on standard characterisation methods used in the aerospace industry. The model was derived with additive decomposition of the strain tensor and consisted of an elastic part, derived from Helmholtz free energy, Hill’s thermodynamic potential, which controls evolution of plastic deformation, and damage orthotopic potential, which controls evolution of damage in material. Damage effects were incorporated using the continuum damage mechanics approach, with the effective stress and energy equivalence principle. Material characterisation and derivation of model parameters was conducted with standard specimens with a uniform cross-section, although a number of tests with non-uniform cross-sections were also conducted here. The tests were designed to assess the extent of damage in material over a range of plastic deformation values, where displacement was measured locally using digital image correlation. The new model was implemented as a user material subroutine in Abaqus and verified and validated against the experimental results for aerospace-grade aluminium alloy 2024-T3. Verification was conducted in a series of single element tests, designed to separately validate elasticity, plasticity and damage-related parts of the model. Validation at this stage of the development was based on comparison of the numerical results with experimental data obtained in the quasistatic characterisation tests, which illustrated the ability of the modelling approach to predict experimentally observed behaviour. A validated user material subroutine allows for efficient simulation-led design improvements of aluminium components, such as stiffened panels and the other thin-wall structures used in the aerospace industry. Full article
Show Figures

Figure 1

Figure 1
<p>Instron 8032 Servo hydraulic test machine with tensile test specimen and 3D Dantec digital image correlation system Q400.</p>
Full article ">Figure 2
<p>AA2024-T3 specimens used for quasistatic testing: (<b>a</b>) standard specimen with uniform size of the gauge cross-section, denoted UCS; (<b>b</b>) specimen with non-uniform size of the gauge cross-section, where plastic deformation was localised within a small zone in the middle of the gauge section, denoted VCS.</p>
Full article ">Figure 3
<p>(<b>a</b>) UCS specimen sample images of optical measurement using Dantec 3D DIC system Q400 images and (<b>b</b>) longitudinal strain surface distribution just before failure (range from 0 to 240).</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) UCS specimen sample images of optical measurement using Dantec 3D DIC system Q400 images and (<b>b</b>) longitudinal strain surface distribution just before failure (range from 0 to 240).</p>
Full article ">Figure 4
<p>True stress true strain curves obtained with two specimens: standard uniform cross-section (UCS) and specimen with varying cross-section (VSC).</p>
Full article ">Figure 5
<p>Instron 8032 Servo hydraulic machine cyclic test input (cross-head displacement versus time) used in Dantec Dynamics Q-400 DIC non-contact measurements.</p>
Full article ">Figure 6
<p>AA2024-T3 cyclic test results in rolling 0° direction with the unloading/reloading slopes that determine elastic modulus degradation due to damage; black lines represent the Young’s moduli of material at a certain level of plastic deformation.</p>
Full article ">Figure 7
<p>AA2024-T3 uniaxial cyclic test data from coupons for damage characterisation of AA-2024-T3 material on elastic modulus degradation ratio versus plastic strain.</p>
Full article ">Figure 8
<p>B vs. β fit from uniaxial experimental results of AA-2024-T3 material cyclic tests, R<sup>2</sup> = 1.</p>
Full article ">Figure 9
<p>True stress—true strain curves calculated from the uniaxial experimental data. R<sup>2</sup> = 0.983.</p>
Full article ">Figure 10
<p>Constitutive model implementation flow chart for the damage model.</p>
Full article ">Figure 11
<p>FEM models with considered mesh size, boundary conditions and damage contour output.</p>
Full article ">Figure 12
<p>Uniaxial tensile stress test: experimental results versus simulation results.</p>
Full article ">Figure 13
<p>Evolution of damage variables versus true strain: rolling direction, <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mn>1</mn> </msub> </mrow> </semantics></math> (0 degree to rolling direction), transverse direction <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mn>2</mn> </msub> </mrow> </semantics></math> (90 degree to rolling direction or transverse direction (TD)) and <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mn>3</mn> </msub> </mrow> </semantics></math> (through-thickness direction TTD).</p>
Full article ">
22 pages, 5524 KiB  
Article
Evaluation of Film Cooling Adiabatic Effectiveness and Net Heat Flux Reduction on a Flat Plate Using Scale-Adaptive Simulation and Stress-Blended Eddy Simulation Approaches
by Rosario Nastasi, Nicola Rosafio, Simone Salvadori and Daniela Anna Misul
Energies 2024, 17(11), 2782; https://doi.org/10.3390/en17112782 - 6 Jun 2024
Viewed by 953
Abstract
The use of film cooling is crucial to avoid high metal temperatures in gas turbine applications, thus ensuring a high lifetime for vanes and blades. The complex turbulent mixing process between the coolant and the main flow requires an accurate numerical prediction to [...] Read more.
The use of film cooling is crucial to avoid high metal temperatures in gas turbine applications, thus ensuring a high lifetime for vanes and blades. The complex turbulent mixing process between the coolant and the main flow requires an accurate numerical prediction to correctly estimate the impact of ejection conditions on the cooling performance. Recent developments in numerical models aim at using hybrid approaches that combine high precision with low computational cost. This paper is focused on the numerical simulation of a cylindrical film cooling hole that operates at a unitary blowing ratio, with a hot gas Mach number of Mam = 0.6, while the coolant is characterized by plenum conditions (Mac = 0). The adopted numerical approach is the Stress-Blended Eddy Simulation model (SBES), which is a blend between a Reynolds-Averaged Navier–Stokes approach and a modeled Large Eddy Simulation based on the local flow and mesh characteristics. The purpose of this paper is to investigate the ability of the hybrid model to capture the complex mixing between the coolant and the main flow. The cooling performance of the hole is quantified through the film cooling effectiveness, the Net Heat Flux Reduction (NHFR), and the discharge coefficient CD calculation. Numerical results are compared both with the experimental data obtained by the University of Karlsruhe during the EU-funded TATEF2 project and with a Scale Adaptive Simulation (SAS) run on the same computational grid. The use of λ2 profiles extracted from the flow field allows for isolating the main vortical structures such as horseshoe vortices, counter-rotating vortex pairs (e.g., kidney vortices), Kelvin–Helmholtz instabilities, and hairpin vortices. Eventually, the contribution of the unsteady phenomena occurring at the hole exit section is quantified through Proper Orthogonal Decomposition (POD) and Spectral Proper Orthogonal Decomposition methods (SPOD). Full article
Show Figures

Figure 1

Figure 1
<p>Numerical domain.</p>
Full article ">Figure 2
<p>SBES shielding function: (<b>a</b>) central plane, (<b>b</b>) mixing region, and (<b>c</b>) hole exit region.</p>
Full article ">Figure 3
<p>Time−averaged adiabatic effectiveness: (<b>a</b>) experimental, (<b>b</b>) SAS, and (<b>c</b>) SBES.</p>
Full article ">Figure 4
<p>Time−averaged laterally averaged adiabatic effectiveness, SAS/SBES comparison with experiments.</p>
Full article ">Figure 5
<p>Time−averaged NHFR maps: (<b>a</b>) SAS and (<b>b</b>) SBES.</p>
Full article ">Figure 6
<p>Time−averaged laterally averaged NHFR, SAS/SBES comparison with experiments.</p>
Full article ">Figure 7
<p>Instantaneous low <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>2</mn> </msub> </semantics></math> modulus iso−surfaces: (<b>a</b>) SAS and (<b>b</b>) SBES.</p>
Full article ">Figure 8
<p>Instantaneous high <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>2</mn> </msub> </semantics></math> modulus iso−surfaces: (<b>a</b>) SAS and (<b>b</b>) SBES.</p>
Full article ">Figure 9
<p>Energy spectrum of <math display="inline"><semantics> <mi mathvariant="italic">tke</mi> </semantics></math> (<b>a</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="italic">X</mi> <mo>/</mo> <mi mathvariant="italic">D</mi> </mrow> </semantics></math> = −1 and <math display="inline"><semantics> <mrow> <mi mathvariant="italic">Y</mi> <mo>/</mo> <mi mathvariant="italic">D</mi> </mrow> </semantics></math> = 0.2, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="italic">X</mi> <mo>/</mo> <mi mathvariant="italic">D</mi> </mrow> </semantics></math> = 0.5 and <math display="inline"><semantics> <mrow> <mi mathvariant="italic">Y</mi> <mo>/</mo> <mi mathvariant="italic">D</mi> </mrow> </semantics></math> = 0.2, and (<b>c</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="italic">X</mi> <mo>/</mo> <mi mathvariant="italic">D</mi> </mrow> </semantics></math> = 1 and <math display="inline"><semantics> <mrow> <mi mathvariant="italic">Y</mi> <mo>/</mo> <mi mathvariant="italic">D</mi> </mrow> </semantics></math> = 0.2. (<b>d</b>) Points on the computational domain.</p>
Full article ">Figure 10
<p>Instantaneous low <math display="inline"><semantics> <msub> <mi>λ</mi> <mn>2</mn> </msub> </semantics></math> modulus iso−surfaces with temperature and x vorticity: (<b>a</b>) SAS and (<b>b</b>) SBES.</p>
Full article ">Figure 11
<p>Central plane comparison of lateral vorticity: (<b>a</b>) SAS and (<b>b</b>) SBES.</p>
Full article ">Figure 12
<p>Fast Fourier Transform of coolant mass flow rate for SAS and SBES simulations.</p>
Full article ">Figure 13
<p>POD energy fraction per mode in the central plane: (<b>a</b>) SAS, (<b>b)</b> SBES, (<b>c</b>) time coefficients of modes #1 and #2 (SAS), and (<b>d</b>) time coefficients of modes #1 and #2 (SBES).</p>
Full article ">Figure 14
<p>POD modes in the central plane: (<b>a</b>) SAS mode #1, (<b>b</b>) SBES mode #1, (<b>c</b>) SAS mode #2, and (<b>d</b>) SBES mode #2.</p>
Full article ">Figure 15
<p>POD energy fraction per mode in the cooled region: (<b>a</b>) SAS, (<b>b</b>) SBES, (<b>c</b>) time coefficients of modes #1 and #2 (SAS), and (<b>d</b>) time coefficients of modes #1 and #2 (SBES).</p>
Full article ">Figure 16
<p>POD modes in the cooled region: (<b>a</b>) SAS mode #1, (<b>b</b>) SBES mode #1, (<b>c</b>) SAS mode #2, and (<b>d</b>) SBES mode #2.</p>
Full article ">Figure 17
<p>SPOD energy spectrum in the cooled region: (<b>a</b>) SAS and (<b>b</b>) SBES.</p>
Full article ">Figure 18
<p>SPOD mode <math display="inline"><semantics> <mrow> <mo>#</mo> <mn>1</mn> </mrow> </semantics></math> magnitude in the cooled region: (<b>a</b>) SAS <span class="html-italic">St</span> = 0.4, (<b>b</b>) SBES <span class="html-italic">St</span> = 0.2, (<b>c</b>) SAS <span class="html-italic">St</span> = 0.74, and (<b>d</b>) SBES <span class="html-italic">St</span> = 0.33.</p>
Full article ">Figure 19
<p>SPOD mode #1 phase in the cooled region: (<b>a</b>) SAS <span class="html-italic">St</span> = 0.4, (<b>b</b>) SBES <span class="html-italic">St</span> = 0.2, (<b>c</b>) SAS <span class="html-italic">St</span> = 0.74, and (<b>d</b>) SBES <span class="html-italic">St</span> = 0.33.</p>
Full article ">Figure 20
<p>SPOD energy spectrum in the central plane: (<b>a</b>) SAS and (<b>b</b>) SBES.</p>
Full article ">Figure 21
<p>SPOD mode <math display="inline"><semantics> <mrow> <mo>#</mo> <mn>1</mn> </mrow> </semantics></math> magnitude in the central plane: (<b>a</b>), SAS <span class="html-italic">St</span> = 0.4, (<b>b</b>) SBES <span class="html-italic">St</span> = 0.33, (<b>c</b>) SAS <span class="html-italic">St</span> = 0.74, and (<b>d</b>) SBES <span class="html-italic">St</span> = 0.74.</p>
Full article ">Figure 22
<p>SPOD mode <math display="inline"><semantics> <mrow> <mo>#</mo> <mn>1</mn> </mrow> </semantics></math> phase in the central plane: (<b>a</b>) SAS <span class="html-italic">St</span> = 0.4, (<b>b</b>) SBES <span class="html-italic">St</span> = 0.33, (<b>c</b>) SAS <span class="html-italic">St</span> = 0.74, (<b>d</b>) SBES <span class="html-italic">St</span> = 0.74, (<b>e</b>) SAS Instantaneous Temperature field, and (<b>f</b>) SBES Instantaneous Temperature field.</p>
Full article ">
16 pages, 4075 KiB  
Article
Impact of K-H Instability on NOx Emissions in N2O Thermal Decomposition Using Premixed CH4 Co-Flow Flames and Electric Furnace
by Juwon Park, Suhyeon Kim, Siyeong Yu, Dae Geun Park, Dong Hyun Kim, Jae-Hyuk Choi and Sung Hwan Yoon
Energies 2024, 17(1), 96; https://doi.org/10.3390/en17010096 - 23 Dec 2023
Viewed by 1176
Abstract
This study systematically investigates the formation of NOx in the thermal decomposition of N2O, focusing on the impact of Kelvin–Helmholtz (K-H) instability in combustion environments. Using premixed CH4 co-flow flames and an electric furnace as distinct heat sources, we [...] Read more.
This study systematically investigates the formation of NOx in the thermal decomposition of N2O, focusing on the impact of Kelvin–Helmholtz (K-H) instability in combustion environments. Using premixed CH4 co-flow flames and an electric furnace as distinct heat sources, we explored NOx emission dynamics under varying conditions, including reaction temperature, residence time, and N2O dilution rates (XN2O). Our findings demonstrate that diluting N2O around a premixed flame increases flame length and decreases flame propagation velocity, inducing K-H instability. This instability was quantitatively characterized using Richardson and Strouhal numbers, highlighting N2O’s role in augmenting oxygen supply within the flame and significantly altering flame dynamics. The study reveals that higher XN2O consistently led to increased NO formation independently of nozzle exit velocity (ujet) or co-flow rate, emphasizing the influence of N2O concentration on NO production. In scenarios without K-H instability, particularly at lower ujet, an exponential rise in NO2 formation rates was observed, due to the reduced residence time of N2O near the flame surface, limiting pyrolysis effectiveness. Conversely, at higher ujet where K-H instability occurs, the formation rate of NO2 drastically decreased. This suggests that K-H instability is crucial in optimizing N2O decomposition for minimal NOx production. Full article
(This article belongs to the Section J2: Thermodynamics)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the experimental setup: premixed CH<sub>4</sub> co-flow flames.</p>
Full article ">Figure 2
<p>Schematic diagram of the experimental setup: electric furnace.</p>
Full article ">Figure 3
<p>(<b>a</b>) Captured flame images and (<b>b</b>) flame propagation velocity in Bunsen premixed flame with varied nozzle exit velocities, <span class="html-italic">u</span><sub>jet</sub>, under a constant value of <span class="html-italic">Q</span><sub>co</sub> = 0.88 L/min.</p>
Full article ">Figure 4
<p>Effects of <span class="html-italic">u</span><sub>jet</sub> on (<b>a</b>) NO and (<b>b</b>) NO<sub>2</sub> formation rate.</p>
Full article ">Figure 5
<p>Effects of <span class="html-italic">Q</span><sub>co</sub> on (<b>a</b>) NO and (<b>b</b>) NO<sub>2</sub> formation rates at <span class="html-italic">u</span><sub>jet</sub> = 55 cm/s, and (<b>c</b>) NO<sub>2</sub> formation rate at <span class="html-italic">u</span><sub>jet</sub> = 50 cm/s.</p>
Full article ">Figure 6
<p>Relative comparison of NO and NO<sub>2</sub> emissions in N<sub>2</sub>O reduction process. The dashed line represents the variation in NOx emission trends according to <span class="html-italic">u</span><sub>jet</sub>.</p>
Full article ">Figure 7
<p>FFT result of K-H instability at <span class="html-italic">u</span><sub>jet</sub> = 55 cm/s, <span class="html-italic">Q</span><sub>co</sub> = 0.88 L/min and <span class="html-italic">X</span><sub>N2O</sub> = 0.25%.</p>
Full article ">Figure 8
<p>(<b>a</b>) Frequency variation with <span class="html-italic">Q</span><sub>co</sub> and (<b>b</b>) correlation between Richardson number and Strouhal number in K-H instability.</p>
Full article ">Figure 9
<p>(<b>a</b>) Schematic of computational domain and (<b>b</b>) top view of the co-flow burner.</p>
Full article ">Figure 10
<p>Calculated reaction temperature-dependent formation rates of (<b>a</b>) NO and (<b>b</b>) NO<sub>2</sub>.</p>
Full article ">Figure 11
<p>Effects of reaction temperature and residence time on the concentrations of (<b>a</b>) NO and (<b>b</b>) NO<sub>2</sub> at electric furnace.</p>
Full article ">Figure 12
<p>Characterization of (<b>a</b>) NO and (<b>b</b>) NO<sub>2</sub> formation rates with effective thermal conductivity in N<sub>2</sub>O decomposition process.</p>
Full article ">
13 pages, 5697 KiB  
Communication
A 3D Anisotropic Thermomechanical Model for Thermally Induced Woven-Fabric-Reinforced Shape Memory Polymer Composites
by Yingyu Wang, Zhiyi Wang, Jia Ma, Chao Luo, Guangqiang Fang and Xiongqi Peng
Sensors 2023, 23(14), 6455; https://doi.org/10.3390/s23146455 - 17 Jul 2023
Viewed by 1156
Abstract
Soft robotic grippers offer great advantages over traditional rigid grippers with respect to grabbing objects with irregular or fragile shapes. Shape memory polymer composites are widely used as actuators and holding elements in soft robotic grippers owing to their finite strain, high specific [...] Read more.
Soft robotic grippers offer great advantages over traditional rigid grippers with respect to grabbing objects with irregular or fragile shapes. Shape memory polymer composites are widely used as actuators and holding elements in soft robotic grippers owing to their finite strain, high specific strength, and high driving force. In this paper, a general 3D anisotropic thermomechanical model for woven fabric-reinforced shape memory polymer composites (SMPCs) is proposed based on Helmholtz free energy decomposition and the second law of thermodynamics. Furthermore, the rule of mixtures is modified to describe the stress distribution in the SMPCs, and stress concentration factors are introduced to account for the shearing interaction between the fabric and matrix and warp yarns and weft yarns. The developed model is implemented with a user material subroutine (UMAT) to simulate the shape memory behaivors of SMPCs. The good consistency between the simulation results and experimental validated the proposed model. Furthermore, a numerical investigation of the effects of yarn orientation on the shape memory behavior of the SMPC soft gripper was also performed. Full article
(This article belongs to the Special Issue Recent Trends and Advances on Space Robot)
Show Figures

Figure 1

Figure 1
<p>A rheological representation of the anisotropic thermomechanical model.</p>
Full article ">Figure 2
<p>The yarn orientation unit vectors.</p>
Full article ">Figure 3
<p>The fitting results regarding <math display="inline"><semantics><mrow><msub><mi>k</mi><mn>1</mn></msub><mo>,</mo><mtext> </mtext><msub><mi>k</mi><mn>2</mn></msub><mo>,</mo><mtext> </mtext><msub><mi>k</mi><mn>3</mn></msub></mrow></semantics></math>.</p>
Full article ">Figure 4
<p>The fitting results regarding <math display="inline"><semantics><mrow><msub><mi>k</mi><mn>4</mn></msub><mo>,</mo><msub><mi>k</mi><mn>5</mn></msub><mo>,</mo><msub><mi>k</mi><mn>6</mn></msub></mrow></semantics></math>.</p>
Full article ">Figure 5
<p>The fitting results regarding <math display="inline"><semantics><mrow><msub><mi>v</mi><mrow><mi>f</mi><mo>_</mo><mi>r</mi><mi>e</mi><mi>f</mi></mrow></msub></mrow></semantics></math>.</p>
Full article ">Figure 6
<p>The fitting results regarding <math display="inline"><semantics><mrow><msub><mi>γ</mi><mrow><mi>i</mi><mi>n</mi><mi>t</mi><mi>e</mi><mi>r</mi></mrow></msub></mrow></semantics></math>.</p>
Full article ">Figure 7
<p>The temperature histories in the cooling and reheating steps.</p>
Full article ">Figure 8
<p>Comparison between experiment and simulation results regarding shape recovery for SMPs.</p>
Full article ">Figure 9
<p>Comparison between experimental and simulation results regarding shape recovery for SMPCs.</p>
Full article ">Figure 10
<p>The soft robotic gripper’s work process.</p>
Full article ">Figure 11
<p>The finite element model of individual SMPC’s soft gripper part.</p>
Full article ">Figure 12
<p>The shape recovery of individual SMPC’s soft gripper part.</p>
Full article ">Figure 13
<p>The stored stress of individual SMPC’s soft gripper part at the beginning of the reheating step.</p>
Full article ">Figure 14
<p>The change in the angle radian values of the 0/90° and ±45° SMPCs in the pre-deformed stage and the unloading step.</p>
Full article ">
24 pages, 5225 KiB  
Article
Reconstruction of the Instantaneous Images Distorted by Surface Waves via Helmholtz–Hodge Decomposition
by Bijian Jian, Chunbo Ma, Yixiao Sun, Dejian Zhu, Xu Tian and Jun Ao
J. Mar. Sci. Eng. 2023, 11(1), 164; https://doi.org/10.3390/jmse11010164 - 9 Jan 2023
Cited by 1 | Viewed by 2048
Abstract
Imaging through water waves will cause complex geometric distortions and motion blur, which seriously affect the correct identification of an airborne scene. The current methods main rely on high-resolution video streams or a template image, which limits their applicability in real-time observation scenarios. [...] Read more.
Imaging through water waves will cause complex geometric distortions and motion blur, which seriously affect the correct identification of an airborne scene. The current methods main rely on high-resolution video streams or a template image, which limits their applicability in real-time observation scenarios. In this paper, a novel recovery method for the instantaneous images distorted by surface waves is proposed. The method first actively projects an adaptive and adjustable structured light pattern onto the water surface for which random fluctuation will cause the image to degrade. Then, the displacement field of the feature points in the structured light image is used to estimate the motion vector field of the corresponding sampling points in the scene image. Finally, from the perspective of fluid mechanics, the distortion-free scene image is reconstructed based on the Helmholtz-Hodge Decomposition (HHD) theory. Experimental results show that our method not only effectively reduces the distortion to the image, but also significantly outperforms state-of-the-art methods in terms of computational efficiency. Moreover, we tested the real-scene sequences of a certain length to verify the stability of the algorithm. Full article
(This article belongs to the Special Issue Underwater Engineering and Image Processing)
Show Figures

Figure 1

Figure 1
<p>Example of Snell’s window as the WAI is calm.</p>
Full article ">Figure 2
<p>Contour distribution of irradiance on the image plane. (<b>a</b>) Contour distribution of normalized illuminance for downwelling radiation from the sky on the image plane. (<b>b</b>) Contour distribution of normalized illuminance (the red dashed line is the boundary of Snell’s window in flat water; the blue solid line is the boundary of Snell’s window in wavy water).</p>
Full article ">Figure 3
<p>Illustration of refractive distortion.</p>
Full article ">Figure 4
<p>Geometry of the image-restoration model via structured light projection, comprising a structured light projection system <math display="inline"><semantics> <mi mathvariant="normal">S</mi> </semantics></math> and a viewing camera <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math>. Component <math display="inline"><semantics> <mi mathvariant="normal">S</mi> </semantics></math> consists of a projector, diffuser plane, and a camera s.</p>
Full article ">Figure 5
<p>Examples of WAI sampling via structured light projection. (<b>a</b>) A preset structured light pattern in the projector. (<b>b</b>) Distribution of WAI sampling points. (<b>c</b>) Control point distribution on the image plane, where the points correspond to WAI sampled points.</p>
Full article ">Figure 6
<p>Processing of structured light images in the experiment. (<b>a</b>) Reference structured light image for flat water surface. (<b>b</b>) Distorted structured light image in the presence of surface waves. (<b>c</b>) Feature extraction for the reference frame. (<b>d</b>) Feature extraction for the distorted frame.</p>
Full article ">Figure 7
<p>Framework of the feature extraction method in this paper.</p>
Full article ">Figure 8
<p>Illustration of the distortion estimation of the scene image. <math display="inline"><semantics> <mrow> <mstyle mathvariant="bold" mathsize="normal"> <mi>δ</mi> </mstyle> <mo stretchy="false">(</mo> <msub> <mstyle mathvariant="bold" mathsize="normal"> <mi>p</mi> </mstyle> <mi>k</mi> </msub> <mo stretchy="false">)</mo> </mrow> </semantics></math> is the amount of displacement of the reflected ray corresponding to the incident light <math display="inline"><semantics> <mrow> <msup> <mstyle mathvariant="bold" mathsize="normal"> <mover accent="true"> <mi>v</mi> <mo>^</mo> </mover> </mstyle> <mi mathvariant="normal">p</mi> </msup> <msub> <mrow/> <mi>k</mi> </msub> </mrow> </semantics></math> on the diffuser plane; <math display="inline"><semantics> <mrow> <msub> <mstyle mathvariant="bold" mathsize="normal"> <mi>d</mi> </mstyle> <mi mathvariant="normal">s</mi> </msub> <mo stretchy="false">(</mo> <msubsup> <mstyle mathvariant="bold" mathsize="normal"> <mi>x</mi> </mstyle> <mi>i</mi> <mi>k</mi> </msubsup> <mo stretchy="false">)</mo> </mrow> </semantics></math> is the amount of displacement of the airborne viewing ray <math display="inline"><semantics> <mrow> <msubsup> <mstyle mathvariant="bold" mathsize="normal"> <mover accent="true"> <mi>v</mi> <mo>^</mo> </mover> </mstyle> <mi mathvariant="normal">a</mi> <mi>k</mi> </msubsup> </mrow> </semantics></math> corresponding to the back-projected ray <math display="inline"><semantics> <mrow> <msubsup> <mstyle mathvariant="bold" mathsize="normal"> <mover accent="true"> <mi>v</mi> <mo>^</mo> </mover> </mstyle> <mi mathvariant="normal">w</mi> <mi>k</mi> </msubsup> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <mstyle mathvariant="bold" mathsize="normal"> <mi>δ</mi> </mstyle> <mo stretchy="false">(</mo> <msub> <mstyle mathvariant="bold" mathsize="normal"> <mi>p</mi> </mstyle> <mi>k</mi> </msub> <mo stretchy="false">)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mstyle mathvariant="bold" mathsize="normal"> <mi>d</mi> </mstyle> <mi mathvariant="normal">s</mi> </msub> <mo stretchy="false">(</mo> <msubsup> <mstyle mathvariant="bold" mathsize="normal"> <mi>x</mi> </mstyle> <mi>i</mi> <mi>k</mi> </msubsup> <mo stretchy="false">)</mo> </mrow> </semantics></math> are a pair of corresponding displacement vectors. <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="bold">w</mi> <mi>k</mi> <mrow> <mi>str</mi> </mrow> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi mathvariant="bold">w</mi> <mo>(</mo> <msubsup> <mi mathvariant="bold">x</mi> <mi>i</mi> <mi>k</mi> </msubsup> <mo>)</mo> </mrow> </semantics></math> are a pair of corresponding pixel displacement vectors.</p>
Full article ">Figure 9
<p>Flow chart of image recovery algorithm based on HHD.</p>
Full article ">Figure 10
<p>Experiment setup. (<b>a</b>) Our real laboratory scene. (<b>b</b>) Scheme of the experiment, comprising a projector, diffuser plane, a viewing camera, and a water tank.</p>
Full article ">Figure 11
<p>A real scene image captured by the camera.</p>
Full article ">Figure 12
<p>Processing image. (<b>a</b>) WAI sampling. (<b>b</b>) Reference structured light image as the WAI was flat. (<b>c</b>) Distorted structured light image. (<b>d</b>) Ground-truth image. (<b>e</b>) Distorted image. (<b>f</b>) Recovered image using our method.</p>
Full article ">Figure 13
<p>Recovery results of different methods. Left to right: the ground-truth image, the distorted structured light image, the sampled image at the corresponding moment, the results of the LWM method [<a href="#B51-jmse-11-00164" class="html-bibr">51</a>], and the results of our method. Up to down: recovery results of sampled images at different times using the above two methods.</p>
Full article ">Figure 14
<p>Recovery results of different methods. Left to right: the ground-truth image, the sampled image, the results of Oreifej’s method using 10 frames, the results of Oreifej’s method using 20 frames, the results of T. Sun’s method using 10 frames, the results of T. Sun’s method using 20 frames, the results of James’s method using 10 frames, the results of James’s method using 20 frames, and the results of our method using a single image.</p>
Full article ">Figure 15
<p>An example of recovery results using the different corner-detection algorithms [<a href="#B58-jmse-11-00164" class="html-bibr">58</a>].</p>
Full article ">Figure 16
<p>Recovery result of a sampled frame using the proposed method.</p>
Full article ">
19 pages, 4762 KiB  
Article
The Role of Inertia in the Onset of Turbulence in a Vortex Filament
by Jean-Paul Caltagirone
Fluids 2023, 8(1), 16; https://doi.org/10.3390/fluids8010016 - 2 Jan 2023
Cited by 2 | Viewed by 1907
Abstract
The decay of the kinetic energy of a turbulent flow with time is not necessarily monotonic. This is revealed by simulations performed in the framework of discrete mechanics, where the kinetic energy can be transformed into pressure energy or vice versa; this persistent [...] Read more.
The decay of the kinetic energy of a turbulent flow with time is not necessarily monotonic. This is revealed by simulations performed in the framework of discrete mechanics, where the kinetic energy can be transformed into pressure energy or vice versa; this persistent phenomenon is also observed for inviscid fluids. Different types of viscous vortex filaments generated by initial velocity conditions show that vortex stretching phenomena precede an abrupt onset of vortex bursting in high-shear regions. In all cases, the kinetic energy starts to grow by borrowing energy from the pressure before the transfer phase to the small turbulent structures. The result observed on the vortex filament is also found for the Taylor–Green vortex, which significantly differs from the previous results on this same case simulated from the Navier–Stokes equations. This disagreement is attributed to the physical model used, that of discrete mechanics, where the formulation is based on the conservation of acceleration. The reasons for this divergence are analyzed in depth; however, a spectral analysis allows finding the established laws on the decay of kinetic energy as a function of the wave number. Full article
(This article belongs to the Special Issue Recent Advances in Fluid Mechanics: Feature Papers, 2022)
Show Figures

Figure 1

Figure 1
<p>Local frame of reference of discrete mechanics; the rectilinear segment <math display="inline"><semantics> <mo>Γ</mo> </semantics></math> oriented by <math display="inline"><semantics> <mi mathvariant="bold">t</mi> </semantics></math> is the constitutive element of the primal structure and the support of the intrinsic acceleration <math display="inline"><semantics> <mi mathvariant="bold">γ</mi> </semantics></math> or imposed by the exterior <math display="inline"><semantics> <mi mathvariant="bold-italic">h</mi> </semantics></math>. This segment, limited by its extremities <span class="html-italic">a</span> and <span class="html-italic">b</span>, is of length <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>h</mi> </mrow> </semantics></math> and called the discrete horizon. The dual structure is schematized by the contour <math display="inline"><semantics> <mo mathvariant="sans-serif">Δ</mo> </semantics></math>, which defines the induced actions then projected on <math display="inline"><semantics> <mo>Γ</mo> </semantics></math>.</p>
Full article ">Figure 2
<p>Geometric structures of discrete mechanics: the primal structure (in blue) is composed of a collection of segments <math display="inline"><semantics> <mo>Γ</mo> </semantics></math> oriented by a unit vector <math display="inline"><semantics> <mi mathvariant="bold">t</mi> </semantics></math> bounded by the ends <span class="html-italic">a</span> or <span class="html-italic">b</span>; these segments form a polygonal planar surface <math display="inline"><semantics> <mi mathvariant="script">S</mi> </semantics></math>, whose barycenter, denoted <span class="html-italic">c</span>, defines the normal <math display="inline"><semantics> <mi mathvariant="bold">n</mi> </semantics></math> that is positively oriented according to Maxwell’s rule. The dual structure (in red) has a flat polygonal surface <math display="inline"><semantics> <mi mathvariant="script">D</mi> </semantics></math> bounded by segments <math display="inline"><semantics> <mo mathvariant="sans-serif">Δ</mo> </semantics></math>. The unit vectors are orthogonal by construction, <math display="inline"><semantics> <mrow> <mi mathvariant="bold">n</mi> <mo>·</mo> <mi mathvariant="bold">t</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Inviscid vortex filament initiated by the velocity field (<a href="#FD16-fluids-08-00016" class="html-disp-formula">16</a>); initial potential field <math display="inline"><semantics> <msup> <mi>ϕ</mi> <mi>o</mi> </msup> </semantics></math> (<b>left</b>), potential field at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> decorated by the local kinetic energy (<b>center</b>), and a snapshot of streamlines (<b>right</b>).</p>
Full article ">Figure 4
<p>Evolutions of the mean kinetic energy <math display="inline"><semantics> <msub> <mi>E</mi> <mi>k</mi> </msub> </semantics></math> and the scalar potential <math display="inline"><semantics> <msup> <mi>ϕ</mi> <mi>o</mi> </msup> </semantics></math> where <math display="inline"><semantics> <msub> <mi>E</mi> <mi>k</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msup> <mi>ϕ</mi> <mi>o</mi> </msup> </semantics></math> are given in <math display="inline"><semantics> <mrow> <msup> <mi>m</mi> <mn>2</mn> </msup> <mspace width="0.222222em"/> <msup> <mi>s</mi> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math> and <span class="html-italic">t</span> in <span class="html-italic">s</span>.</p>
Full article ">Figure 5
<p>Snapshots of potential fields decorated by kinetic energy for three vortex filaments with circular, hexagonal, and square bases.</p>
Full article ">Figure 6
<p>Evolution of the kinetic energy with time in the case of a cylinder (<b>left</b>) and for the geometry with square base (<b>righ</b>).</p>
Full article ">Figure 7
<p>Global kinetic energy <math display="inline"><semantics> <msub> <mi>E</mi> <mi>k</mi> </msub> </semantics></math> and compression energy <math display="inline"><semantics> <msub> <mi>E</mi> <mi>c</mi> </msub> </semantics></math> at Reynolds number <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1600</mn> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>c</mi> </msub> <mo>=</mo> <msup> <mn>256</mn> <mn>3</mn> </msup> </mrow> </semantics></math> cells and <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>e</mi> </msub> <mo>≈</mo> <mn>50</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math> unknowns.</p>
Full article ">Figure 8
<p>Snapshot of the potential field decorated by the kinetic energy for <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>c</mi> </msub> <mo>=</mo> <msup> <mn>256</mn> <mn>3</mn> </msup> </mrow> </semantics></math> et <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>e</mi> </msub> <mo>≈</mo> <mn>50</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math> unknowns.</p>
Full article ">Figure 9
<p>Energy spectrum as a function of the wave number <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mi>k</mi> </msub> <mrow> <mo>(</mo> <mi>k</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math> on one-quarter of the domain.</p>
Full article ">Figure 10
<p>The evolution of the kinetic energy <math display="inline"><semantics> <msub> <mi>E</mi> <mi>k</mi> </msub> </semantics></math> corresponding to the Taylor–Green vortex case shows a similar behavior at Reynolds numbers of <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>500</mn> <mo>,</mo> <mn>1600</mn> </mrow> </semantics></math> and for an inviscid fluid.</p>
Full article ">Figure 11
<p>Taylor–Green vortex at <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1600</mn> </mrow> </semantics></math>; the spectral simulation provides the evolution of <math display="inline"><semantics> <msub> <mi>E</mi> <mi>k</mi> </msub> </semantics></math> and dissipation rate <math display="inline"><semantics> <mrow> <mi>ε</mi> <mo>=</mo> <mo>−</mo> <mi>d</mi> <msub> <mi>E</mi> <mi>k</mi> </msub> <mo>/</mo> <mi>d</mi> <mi>t</mi> </mrow> </semantics></math> for the Navier–Stokes model after W. van Rees [<a href="#B11-fluids-08-00016" class="html-bibr">11</a>], file Re-1600-512.gdiag.</p>
Full article ">Figure 12
<p>Scheme of the inertia vector <math display="inline"><semantics> <msub> <mi mathvariant="sans-serif-bold-italic">κ</mi> <mi>i</mi> </msub> </semantics></math> on the oriented segment <math display="inline"><semantics> <mo>Γ</mo> </semantics></math> and the respective representations of its two components such that <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif-bold-italic">κ</mi> <mi>i</mi> </msub> <mo>=</mo> <mo>∇</mo> <msub> <mi>ϕ</mi> <mi>i</mi> </msub> <mo>−</mo> <msup> <mo>∇</mo> <mi>d</mi> </msup> <mo>×</mo> <msub> <mi mathvariant="bold">ψ</mi> <mi>i</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">
21 pages, 7178 KiB  
Article
Assessment of a Hybrid Eulerian–Lagrangian CFD Solver for Wind Turbine Applications and Comparison with the New MEXICO Experiment
by Nikos Spyropoulos, George Papadakis, John M. Prospathopoulos and Vasilis A. Riziotis
Fluids 2022, 7(9), 296; https://doi.org/10.3390/fluids7090296 - 8 Sep 2022
Cited by 1 | Viewed by 2021
Abstract
In this paper, the hybrid Lagrangian–Eulerian solver HoPFlow is presented and evaluated against wind tunnel measurements from the New MEXICO experiment. In the paper, the distinct solvers that assemble the HoPFlow solver are presented, alongside with details on their mutual coupling and interaction. [...] Read more.
In this paper, the hybrid Lagrangian–Eulerian solver HoPFlow is presented and evaluated against wind tunnel measurements from the New MEXICO experiment. In the paper, the distinct solvers that assemble the HoPFlow solver are presented, alongside with details on their mutual coupling and interaction. The Eulerian solver, MaPFlow, solves the compressible Navier–Stokes equations under a cell-centered finite-volume discretization scheme, while the Lagrangian solver uses numerical particles that carry mass, pressure, dilatation and vorticity as flow markers in order to represent the flow-field by following their trajectories. The velocity field is calculated with the use of the decomposition theorem introduced by Helmholtz. Computational performance is enhanced by utilizing the particle mesh (PM) methodology in order to solve the Poisson equations for the scalar potential ϕ and the stream function ψ. The hybrid solver is tested in 3-D unsteady simulations concerning the axial flow around the wind turbine (WT) model rotor tested in the New MEXICO experimental campaign. Simulation results are presented as integrated rotor loads, radial distribution of aerodynamic forces and moments and pressure distributions at various span-wise positions along the rotor blades. Comparison is made against experimental data and computational results produced by the pure Eulerian solver. A total of 5 PM nodes per chord length of the blade section at 75% have been found to be sufficient to predict the loading at the tip region of the blade with great accuracy. Discrepancies with respect to measurements, observed at the root and middle sections of the blade, are attributed to the omission of the spinner geometry in the simulations. Full article
Show Figures

Figure 1

Figure 1
<p>Decomposition of Eulerian <math display="inline"><semantics> <mfenced separators="" open="(" close=")"> <msub> <mi>D</mi> <mi>E</mi> </msub> </mfenced> </semantics></math> and Lagrangian <math display="inline"><semantics> <mfenced separators="" open="(" close=")"> <msub> <mi>D</mi> <mi>P</mi> </msub> </mfenced> </semantics></math> computational domains. <math display="inline"><semantics> <msub> <mi>S</mi> <mi>B</mi> </msub> </semantics></math> denotes the solid-wall boundaries, and <math display="inline"><semantics> <msub> <mi>S</mi> <mi>E</mi> </msub> </semantics></math> the far-field of the Eulerian domain.</p>
Full article ">Figure 2
<p>The Lagrangian particle solution is used to define the far-field boundary conditions of the Eulerian solver. Lagrangian particles are depicted as solid blue circles. Solid and dotted red circles denote the centers of the Eulerian cells and ghost cells, respectively.</p>
Full article ">Figure 3
<p>The Eulerian solution is used in order to correct the Lagrangian particles that are inside the Eulerian domain. Lagrangian particles are depicted as solid blue circles. Dotted red circles denote the centers of the Eulerian cells. Small solid red circles illustrate the Eulerian particles that correct the Lagrangian ones.</p>
Full article ">Figure 4
<p>Flowchart of the hybrid solver.</p>
Full article ">Figure 5
<p>Lateral view of the purely Eulerian grid.</p>
Full article ">Figure 6
<p>Eulerian sub-domain used in hybrid solver simulations.</p>
Full article ">Figure 7
<p>Visualization of the Lagrangian and the Eulerian sub-domains in the hybrid solver simulations.</p>
Full article ">Figure 8
<p>Thrust (N) and Torque (Nm) estimation with respect to the number of PM nodes per chord length.</p>
Full article ">Figure 9
<p>Radial distribution of normal and tangential forces. Comparison between experimental measurements and computational predictions.</p>
Full article ">Figure 10
<p>Pressure distribution. Comparison between experimental measurements and predictions by different computational tools.</p>
Full article ">Figure 10 Cont.
<p>Pressure distribution. Comparison between experimental measurements and predictions by different computational tools.</p>
Full article ">Figure 11
<p>Radial distribution of normal forces. Comparison between experimental measurements and computational predictions.</p>
Full article ">Figure 12
<p>Pressure distribution at <math display="inline"><semantics> <mrow> <mn>60</mn> <mo>%</mo> </mrow> </semantics></math> of the blade. Comparison between experimental measurements and computational predictions.</p>
Full article ">
18 pages, 3473 KiB  
Article
Helmholtz–Galerkin Regularizing Technique for the Analysis of the THz-Range Surface-Plasmon-Mode Resonances of a Graphene Microdisk Stack
by Mario Lucido
Micro 2022, 2(2), 295-312; https://doi.org/10.3390/micro2020019 - 16 May 2022
Cited by 4 | Viewed by 1699
Abstract
The aim of this paper is the accurate and efficient analysis of the surface-plasmon-mode resonances of a graphene microdisk stack in the terahertz range. By means of suitable generalized boundary conditions and Fourier series expansion, the problem is formulated in terms of sets [...] Read more.
The aim of this paper is the accurate and efficient analysis of the surface-plasmon-mode resonances of a graphene microdisk stack in the terahertz range. By means of suitable generalized boundary conditions and Fourier series expansion, the problem is formulated in terms of sets of one-dimensional integral equations in the vector Hankel transform domain for the harmonics of the surface current densities. In virtue of the Helmholtz decomposition, the unknowns are replaced by the corresponding surface curl-free and divergence-free contributions. An approximate solution is achieved by means of the Galerkin method. The proper selection of expansion functions reconstructing the physical behavior of the surface current densities leads to a fast-converging Fredholm second-kind matrix equation, whose elements are accurately and efficiently evaluated by means of a suitable analytical procedure in the complex plane. It is shown that the surface-plasmon-mode resonance frequencies upshift by increasing the number of disks and by decreasing the distance between the disks, and that new resonances can arise for small with respect to the radius distances between the disks, resembling the dipole-mode resonances of the dielectric disk, while, for larger distances, the surface-plasmon-mode resonances can split. Full article
Show Figures

Figure 1

Figure 1
<p>Geometry of the problem.</p>
Full article ">Figure 2
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the disk’s plane of the graphene disk with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, when a plane wave normally impinges onto the disk: (<b>a</b>) TSCS and ACS; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.4008974</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>13.6008776</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the top disk’s plane of the graphene disk stack with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mo> </mo> <mn>0.1</mn> <mo>,</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math> when a plane wave normally impinges onto the disk stack: (<b>a</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>15.4504705</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>18.3603764</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>d</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>; (<b>e</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>13.5905306</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>f</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>15.5404975</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>g</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>; (<b>h</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.4006314</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>i</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>13.5605316</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 3 Cont.
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the top disk’s plane of the graphene disk stack with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mo> </mo> <mn>0.1</mn> <mo>,</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math> when a plane wave normally impinges onto the disk stack: (<b>a</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>15.4504705</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>18.3603764</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>d</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>; (<b>e</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>13.5905306</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>f</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>15.5404975</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>g</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>; (<b>h</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.4006314</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>i</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>13.5605316</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 3 Cont.
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the top disk’s plane of the graphene disk stack with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mo> </mo> <mn>0.1</mn> <mo>,</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math> when a plane wave normally impinges onto the disk stack: (<b>a</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>15.4504705</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>18.3603764</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>d</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>; (<b>e</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>13.5905306</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>f</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>15.5404975</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>g</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>; (<b>h</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.4006314</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>i</b>) near-electric field for <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>13.5605316</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the top disk’s plane of the graphene disk stack with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> <mo>,</mo> <mn>5</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, when a plane wave normally impinges onto the disk stack: (<b>a</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>2.8537002</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>18.2283268</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>d</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>; (<b>e</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>3.4242523</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>f</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>20.3768854</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>g</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math>; (<b>h</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>4.0848799</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>i</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>22.0854271</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 4 Cont.
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the top disk’s plane of the graphene disk stack with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> <mo>,</mo> <mn>5</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, when a plane wave normally impinges onto the disk stack: (<b>a</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>2.8537002</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>18.2283268</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>d</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>; (<b>e</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>3.4242523</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>f</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>20.3768854</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>g</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math>; (<b>h</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>4.0848799</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>i</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>22.0854271</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the top disk’s plane of the graphene disk stack with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mfrac> <mn>1</mn> <mn>3</mn> </mfrac> </mrow> </semantics></math>, when a plane wave normally impinges onto the disk stack: (<b>a</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>10.9806450</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.8206178</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>d</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>; (<b>e</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.3517838</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>f</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>12.0124114</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>g</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>; (<b>h</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.6220305</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>i</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>12.1325255</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 5 Cont.
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the top disk’s plane of the graphene disk stack with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mfrac> <mn>1</mn> <mn>3</mn> </mfrac> </mrow> </semantics></math>, when a plane wave normally impinges onto the disk stack: (<b>a</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>10.9806450</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.8206178</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>d</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>; (<b>e</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.3517838</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>f</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>12.0124114</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>g</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>; (<b>h</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.6220305</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>i</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>12.1325255</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 5 Cont.
<p>TSCS and ACS for varying values of the frequency, and near-electric field behavior in the top disk’s plane of the graphene disk stack with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>10</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mrow> <mi>relax</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>ps</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi mathvariant="normal">c</mi> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi>eV</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mfrac> <mi>d</mi> <mi>a</mi> </mfrac> <mo>=</mo> <mfrac> <mn>1</mn> <mn>3</mn> </mfrac> </mrow> </semantics></math>, when a plane wave normally impinges onto the disk stack: (<b>a</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>; (<b>b</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>10.9806450</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>c</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.8206178</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>d</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>; (<b>e</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.3517838</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>f</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>12.0124114</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>g</b>) TSCS and ACS for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>; (<b>h</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>11.6220305</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>; (<b>i</b>) near-electric field for <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>12.1325255</mn> <mrow> <mtext> </mtext> <mi>THz</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">
23 pages, 3275 KiB  
Article
Can DtN and GenEO Coarse Spaces Be Sufficiently Robust for Heterogeneous Helmholtz Problems?
by Niall Bootland and Victorita Dolean
Math. Comput. Appl. 2022, 27(3), 35; https://doi.org/10.3390/mca27030035 - 21 Apr 2022
Viewed by 2722
Abstract
Numerical solutions of heterogeneous Helmholtz problems present various computational challenges, with descriptive theory remaining out of reach for many popular approaches. Robustness and scalability are key for practical and reliable solvers in large-scale applications, especially for large wave number problems. In this work, [...] Read more.
Numerical solutions of heterogeneous Helmholtz problems present various computational challenges, with descriptive theory remaining out of reach for many popular approaches. Robustness and scalability are key for practical and reliable solvers in large-scale applications, especially for large wave number problems. In this work, we explore the use of a GenEO-type coarse space to build a two-level additive Schwarz method applicable to highly indefinite Helmholtz problems. Through a range of numerical tests on a 2D model problem, discretised by finite elements on pollution-free meshes, we observe robust convergence, iteration counts that do not increase with the wave number, and good scalability of our approach. We further provide results showing a favourable comparison with the DtN coarse space. Our numerical study shows promise that our solver methodology can be effective for challenging heterogeneous applications. Full article
(This article belongs to the Special Issue Domain Decomposition Methods)
Show Figures

Figure 1

Figure 1
<p>Local eigenfunctions for <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>46.5</mn> </mrow> </semantics></math>. <b>Top row:</b> Examples using DtN (14). <b>Middle row:</b> Equivalent examples using H-GenEO (19). <b>Bottom row:</b> Examples using H-GenEO which are not found amongst the DtN eigenfunctions.</p>
Full article ">Figure 2
<p>Schematic of the 2D wave guide model problem with example triangular mesh.</p>
Full article ">Figure 3
<p>The size of coarse space utilised for the homogeneous problem when using ORAS with the DtN and H-GenEO coarse spaces. A uniform decomposition into <math display="inline"><semantics> <mrow> <msqrt> <mi>N</mi> </msqrt> <mo>×</mo> <msqrt> <mi>N</mi> </msqrt> </mrow> </semantics></math> square subdomains is used. (<b>a</b>) Varying the wave number <span class="html-italic">k</span> for <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>25</mn> </mrow> </semantics></math>, (<b>b</b>) Varying the number of subdomains <span class="html-italic">N</span> for <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>73.8</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Piecewise constant layer profiles for the wave speed <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>)</mo> </mrow> </semantics></math>. For the darkest shade <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>(</mo> <mi mathvariant="bold-italic">x</mi> <mo>)</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, while for the lightest shade <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>(</mo> <mi mathvariant="bold-italic">x</mi> <mo>)</mo> <mo>=</mo> <mi>ρ</mi> </mrow> </semantics></math>, with <math display="inline"><semantics> <mi>ρ</mi> </semantics></math> being the contrast factor.</p>
Full article ">Figure 5
<p>Schematic of the growing 2D wave guide model problem used in a weak scaling test on <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>25</mn> <mi>L</mi> </mrow> </semantics></math> fixed size subdomains, with the underlying non-overlapping subdomains shown in grey.</p>
Full article ">Figure 6
<p>Timings for the homogeneous problem when using ORAS with H-GenEO (<math display="inline"><semantics> <mfrac> <mn>1</mn> <mn>2</mn> </mfrac> </semantics></math>) and a varying number of subdomains for <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>186.0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi>h</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <mo>=</mo> <mn>3200</mn> </mrow> </semantics></math>, giving a total of <math display="inline"><semantics> <mrow> <mn>10</mn> <mo>,</mo> <mspace width="-0.166667em"/> <mn>246</mn> <mo>,</mo> <mspace width="-0.166667em"/> <mn>401</mn> </mrow> </semantics></math> dofs. A non-uniform decomposition into <span class="html-italic">N</span> subdomains is used, given by METIS.</p>
Full article ">
16 pages, 1980 KiB  
Article
Depolarization of Vector Light Beams on Propagation in Free Space
by Nikolai Petrov
Photonics 2022, 9(3), 162; https://doi.org/10.3390/photonics9030162 - 6 Mar 2022
Cited by 11 | Viewed by 2758
Abstract
Nonparaxial propagation of the vector vortex light beams in free space was investigated theoretically. Propagation-induced polarization changes in vector light beams with different spatial intensity distributions were analyzed. It is shown that the hybrid vector Bessel modes with polarization-OAM (orbital angular momentum) entanglement [...] Read more.
Nonparaxial propagation of the vector vortex light beams in free space was investigated theoretically. Propagation-induced polarization changes in vector light beams with different spatial intensity distributions were analyzed. It is shown that the hybrid vector Bessel modes with polarization-OAM (orbital angular momentum) entanglement are the exact solutions of the vector Helmholtz equation. Decomposition of arbitrary vector beams in the initial plane z = 0 into these polarization-invariant beams with phase and polarization singularities was used to analyze the evolution of the polarization of light within the framework of the 2 × 2 coherency matrix formalism. It is shown that the 2D degree of polarization decreases with distance if the incident vector beam is not the modal solution. The close relationship of the degree of polarization with the quantum-mechanical purity parameter is emphasized. Full article
(This article belongs to the Special Issue Polarized Light and Optical Systems)
Show Figures

Figure 1

Figure 1
<p>The purity (<b>a</b>) and impurity (<b>b</b>) as function of distance. 1 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 30 μm; 2 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 50 μm; 3 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 100 μm. <span class="html-italic">λ</span> = 0.63 μm.</p>
Full article ">Figure 2
<p>(<b>a</b>) Impurity as function of distance. (<b>b</b>) Entropy as function of distance. <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>15</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Degree of polarization as function of distance. 1 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 30 μm; 2 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 50 μm; 3 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 100 μm. (<b>b</b>) Degree of polarization as function of distance. <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>15</mn> <mo> </mo> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Degree of polarization as function of distance. Gaussian beam (black line), <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 30 μm; BG beam (red line), <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 200 μm, <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mi>B</mi> </msub> </mrow> </semantics></math> = 40 μm.</p>
Full article ">Figure 5
<p>Intensity profiles of Gaussian (black line) and BG (red line) beams. <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 30 μm, <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 200 μm, <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mi>B</mi> </msub> </mrow> </semantics></math> = 40 μm.</p>
Full article ">Figure 6
<p>(<b>a</b>) Degree of polarization as function of distance. 1 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 50 μm; 2 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 70 μm; 3 − <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 100 μm. λ = 0.63 μm. (<b>b</b>) Degree of polarization as function of distance. Gaussian beam (black line), <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 45 μm; BG beam (red line), <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mi>B</mi> </msub> </mrow> </semantics></math> = 70 μm; <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 200 μm. λ = 0.63 μm.</p>
Full article ">Figure 7
<p>Intensity profiles of Gaussian (black line) and BG (red line) beams. <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 45 μm; <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> </mrow> </semantics></math> = 200 μm, <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mi>B</mi> </msub> </mrow> </semantics></math> = 70 μm.</p>
Full article ">
29 pages, 2024 KiB  
Article
A Parallel Dissipation-Free and Dispersion-Optimized Explicit Time-Domain FEM for Large-Scale Room Acoustics Simulation
by Takumi Yoshida, Takeshi Okuzono and Kimihiro Sakagami
Buildings 2022, 12(2), 105; https://doi.org/10.3390/buildings12020105 - 23 Jan 2022
Cited by 15 | Viewed by 3890
Abstract
Wave-based acoustics simulation methods such as finite element method (FEM) are reliable computer simulation tools for predicting acoustics in architectural spaces. Nevertheless, their application to practical room acoustics design is difficult because of their high computational costs. Therefore, we propose herein a parallel [...] Read more.
Wave-based acoustics simulation methods such as finite element method (FEM) are reliable computer simulation tools for predicting acoustics in architectural spaces. Nevertheless, their application to practical room acoustics design is difficult because of their high computational costs. Therefore, we propose herein a parallel wave-based acoustics simulation method using dissipation-free and dispersion-optimized explicit time-domain FEM (TD-FEM) for simulating room acoustics at large-scale scenes. It can model sound absorbers with locally reacting frequency-dependent impedance boundary conditions (BCs). The method can use domain decomposition method (DDM)-based parallel computing to compute acoustics in large rooms at kilohertz frequencies. After validation studies of the proposed method via impedance tube and small cubic room problems including frequency-dependent impedance BCs of two porous type sound absorbers and a Helmholtz type sound absorber, the efficiency of the method against two implicit TD-FEMs was assessed. Faster computations and equivalent accuracy were achieved. Finally, acoustics simulation of an auditorium of 2271 m3 presenting a problem size of about 150,000,000 degrees of freedom demonstrated the practicality of the DDM-based parallel solver. Using 512 CPU cores on a parallel computer system, the proposed parallel solver can compute impulse responses with 3 s time length, including frequency components up to 3 kHz within 9000 s. Full article
Show Figures

Figure 1

Figure 1
<p>Dispersion errors of 4th-E and Opt-E TD-FEMs compared to 4th-I and 2nd-I TD-FEMs. The comparison is made as a function of the spatial resolution. Opt-E TD-FEMs use four optimized conditions in which the dispersion error is minimized at different spatial resolutions.</p>
Full article ">Figure 2
<p>Impedance tube model with a frequency-dependent impedance boundary.</p>
Full article ">Figure 3
<p>Modulated Gaussian pulse used as a source signal: (<b>a</b>) waveform of volume acceleration and (<b>b</b>) spectrum characteristic.</p>
Full article ">Figure 4
<p>Comparisons of normal incidence absorption coefficient (<b>Left column</b>) and surface impedance (<b>Right column</b>) among the theory, the fitted rational model and the 4th-E TD-FEM for (<b>a</b>) GW, (<b>b</b>) NF, and (<b>c</b>) MPP-GW.</p>
Full article ">Figure 5
<p>Analyzed cubic room of 1 m<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math>, where the colored surfaces are, respectively, assigned frequency-dependent impedance of GW (red), NF (blue), and MPP-GW (green).</p>
Full article ">Figure 6
<p>Comparisons of mean SPLs between the FD-FEM and the 4th-E TD-FEM with convergence tolerance of 10<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </semantics></math> for (<b>a</b>) Case 1 and (<b>b</b>) Case 2, where the values in parentheses following TD-FEM represent the convergence tolerance for CR method.</p>
Full article ">Figure 7
<p>Comparisons of SPL spectra at receiver (<span class="html-italic">x</span>, <span class="html-italic">y</span>, <span class="html-italic">z</span>) = (0.6, 0.5, 0.5): (<b>a</b>) the reference solution vs. the Opt-E TD-FEM (optimized at <span class="html-italic">R</span> = 6.25), (<b>b</b>) the reference solution vs. the 4th-E TD-FEM, (<b>c</b>) the reference solution vs. the 4th-I TD-FEM, and (<b>d</b>) the reference solution vs. the 2nd-I TD-FEM.</p>
Full article ">Figure 8
<p>Comparisons of SPL spectra above 1.8 kHz at receiver (<span class="html-italic">x</span>, <span class="html-italic">y</span>, <span class="html-italic">z</span>) = (0.6, 0.5, 0.5): (<b>a</b>) the reference solution vs. Opt-E TD-FEM (optimized at <span class="html-italic">R</span> = 6.87) and (<b>b</b>) the reference solution vs. Opt-E TD-FEM (optimized at <span class="html-italic">R</span> = 6.25), (<b>c</b>) the reference solution vs. Opt-E TD-FEM (optimized at <span class="html-italic">R</span> = 5.5) and (<b>d</b>) the reference solution vs. Opt-E TD-FEM (optimized at <span class="html-italic">R</span> = 4.91).</p>
Full article ">Figure 9
<p>Analyzed auditorium model of 2271 <math display="inline"><semantics> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> </semantics></math>: (<b>Left</b>) dimension and (<b>Right</b>) setting of boundary conditions where non-colored surfaces are reflective with frequency-independent impedance and where colored surfaces are absorptive with frequency-dependent impedance of GW (red, yellow), NF (blue), and MPP-GW (green), in which yellow surfaces are assigned absorption properties only in Cond. 2. The yellow surface size is 1.74 m × 1.2 m (<span class="html-italic">x</span>×<span class="html-italic">z</span>).</p>
Full article ">Figure 10
<p>Decomposed auditorium model with 256 subdomains.</p>
Full article ">Figure 11
<p>Distribution of workload on each subdomain. Blue line: DOF; Pink line: number of communication data per communication.</p>
Full article ">Figure 12
<p>Visualized <math display="inline"><semantics> <mrow> <mi>z</mi> <mi>x</mi> </mrow> </semantics></math> plane (<span class="html-italic">y</span> = 0) sound fields at (<b>a</b>) 5 ms, (<b>b</b>) 12.5 ms, (<b>c</b>) 20 ms, and (<b>d</b>) 27.5 ms.</p>
Full article ">Figure 13
<p>Calculated results at receiver R13 in Cond. 1 (<b>Left</b>) and Cond. 2 (<b>Right</b>): (<b>a</b>) IRs and (<b>b</b>) EDCs.</p>
Full article ">Figure 14
<p>Comparisons of room acoustic parameters for Cond. 1 and Cond. 2: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>T</mi> <mn>30</mn> </msub> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>E</mi> <mi>D</mi> <mi>T</mi> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <msub> <mi>C</mi> <mn>50</mn> </msub> </semantics></math>, and (<b>d</b>) <span class="html-italic">G</span>.</p>
Full article ">Figure A1
<p>Comparisons of mean SPLs among the 4th-E TD-FEM with various tolerance values for (<b>a</b>) Case 1 and (<b>b</b>) Case 2, where the values in parentheses following TD-FEM represent the convergence tolerance for CR method.</p>
Full article ">Figure A2
<p>Absolute differences with respect to mean SPL in each tolerance condition from the sufficiently accurate condition of 10<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </semantics></math> in (<b>a</b>) Case 1 and (<b>b</b>) Case 2, where the values in parentheses following TD-FEM represent the convergence tolerance for CR method.</p>
Full article ">
33 pages, 4557 KiB  
Article
Stochastic Chaos and Markov Blankets
by Karl Friston, Conor Heins, Kai Ueltzhöffer, Lancelot Da Costa and Thomas Parr
Entropy 2021, 23(9), 1220; https://doi.org/10.3390/e23091220 - 17 Sep 2021
Cited by 69 | Viewed by 8018
Abstract
In this treatment of random dynamical systems, we consider the existence—and identification—of conditional independencies at nonequilibrium steady-state. These independencies underwrite a particular partition of states, in which internal states are statistically secluded from external states by blanket states. The existence of such partitions [...] Read more.
In this treatment of random dynamical systems, we consider the existence—and identification—of conditional independencies at nonequilibrium steady-state. These independencies underwrite a particular partition of states, in which internal states are statistically secluded from external states by blanket states. The existence of such partitions has interesting implications for the information geometry of internal states. In brief, this geometry can be read as a physics of sentience, where internal states look as if they are inferring external states. However, the existence of such partitions—and the functional form of the underlying densities—have yet to be established. Here, using the Lorenz system as the basis of stochastic chaos, we leverage the Helmholtz decomposition—and polynomial expansions—to parameterise the steady-state density in terms of surprisal or self-information. We then show how Markov blankets can be identified—using the accompanying Hessian—to characterise the coupling between internal and external states in terms of a generalised synchrony or synchronisation of chaos. We conclude by suggesting that this kind of synchronisation may provide a mathematical basis for an elemental form of (autonomous or active) sentience in biology. Full article
Show Figures

Figure 1

Figure 1
<p>The Lorenz system and stochastic chaos. (<b>Upper panel</b>) this illustrates the solution to the Lorenz system of equations with (lines of asterisks) and without (solid lines) random fluctuations, with a variance of 16. (<b>Middle panels</b>) the left panel shows the corresponding solutions in the three-dimensional state-space, illustrating the butterfly shape of the limit set (deterministic solution: solid line) and random attractor (stochastic solution: line of asterisks). The trajectory in the right panel is the deterministic solution to the Laplacian form of the Lorenz system based upon a Helmholtz decomposition parameterised with a second-order polynomial. (<b>Lower left panel</b>) this plots the fluctuations in the potential evaluated using the Laplacian form, which expresses the self-information (i.e., potential) as an analytic (second-order polynomial) function of the states. (<b>Lower right panel</b>) this potential function (of the first two states) is shown as an image, with the trajectory superimposed.</p>
Full article ">Figure 2
<p>The Helmholtz decomposition of flows. (<b>Upper panel</b>) this illustrates the flow of the Lorenz system using the Helmholtz decomposition into solenoidal flow (red) gradient flow (blue) and correction flow (gold). The flow is shown as a quiver plot at equally spaced points in state-space. (<b>Lower panel</b>) this uses the same format but for a Laplacian system based upon the Lorenz system in the upper panel. The key difference here is that the dissipative part of the flow operator and Hessian are positive definite, which means the gradient flows converge to the maximum of the nonequilibrium steady-state density. This is reflected in the blue arrows that point to the centre of this state-space.</p>
Full article ">Figure 3
<p>A chaotic Laplacian system. (<b>Upper panels</b>) these show the solution or trajectory of three states comprising a Laplacian approximation to a stochastic Lorenz system. The red dots mark a deterministic solution to the equations of motion, while the purple dots illustrate a stochastic solution with random fluctuations. These trajectories are superimposed on an image representation of the nonequilibrium steady-state density that—by construction in this Laplacian system—is multivariate Gaussian. (<b>Middle panels</b>) the left panel shows the deterministic (solid lines) and stochastic (dotted lines) solutions as a function of time, while the right panel plots the same trajectories in state-space. The shape of the attractor retains a butterfly-like form but is clearly different from the Lorenz attractor. (<b>Lower left panel</b>) this plots the potential or self-information as a function of time based upon the analytic form for the equations of motion and the deterministic trajectory of the previous panel. In the absence of the correction term, the gradient flow would ensure that this potential decreased over time, because solenoidal flow is divergence free (i.e., is conservative). However, there are slight fluctuations around the minimum potential induced by the correction term. (<b>Lower right panel</b>) this plots the flow of the Laplacian system (approximate flow) against the flow of the Lorenz system (true flow) evaluated at 64 equally spaced sample points. The different colours correspond to the components of flow or motion in the three dimensions. It can be seen that although there is a high correlation between the flows of the Laplacian and Lorenz systems, they are not identical.</p>
Full article ">Figure 4
<p>Nonequilibrium steady-state density. (<b>Upper panel</b>) these images report the constraints on coupling entailed by the Jacobian (left) and manifest in terms of the Hessian (right). The inverse of the Hessian matrix can be read (under the Laplace assumption) as the covariance matrix of the three states. In this example, the third state is independent of the first pair, where this independence rests on the directed coupling from the third to the first state. The matrices correspond to the log of the absolute values of the matrix elements—to disclose their sparsity structure. (<b>Middle panel</b>) these show slices through the ensuing steady-state density over two states, at increasing values of the remaining state. They illustrate the fact that the only correlation in play is between the first and second states. (<b>Lower panel</b>) this correlation is illustrated in terms of the conditional density over the first state, given the second. The shaded areas correspond to the probability density and the white line is the conditional expectation. The red line is the realised trajectory of the first state that is largely confined to the 90% credible intervals. This characterisation uses the trajectory from the stochastic solution shown in <a href="#entropy-23-01220-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>A high-order approximation. This figure uses the same format as <a href="#entropy-23-01220-f003" class="html-fig">Figure 3</a> to illustrate the dynamics of a three-dimensional system that is indistinguishable from a Lorentz system. However, in this instance, the equations of motion can be decomposed into a solenoidal and gradient flow in which the dissipative part of the flow operator and Hessian are positive definite. In other words, this system is apt to describe stochastic chaos driven by random fluctuations to a proper nonequilibrium steady-state density. In this example, the solenoidal flow was parameterised up to second-order and the potential up to fourth-order, with constraints to ensure the Hessian was positive definite everywhere. The high order terms in the Hessian mean that the steady-state density in the upper panels is no longer Gaussian (although univariate and bivariate conditional densities remain Gaussian).</p>
Full article ">Figure 6
<p>Density dynamics. These images show snapshots of a time-dependent probability density for the chaotic Laplacian system in <a href="#entropy-23-01220-f003" class="html-fig">Figure 3</a>. They report the marginal density over the first two states, averaged over successive epochs of two seconds (assuming an integration time of 1/64 s). The system was prepared in an initial state with a relatively precise density, centred around [4, 4, 8]. This density converges to the steady-state density (see <a href="#entropy-23-01220-f003" class="html-fig">Figure 3</a>) after about 16 s; however, it takes a rather circuitous route from this particular set of initial states. Note that the average density over short periods of time can be highly non-Gaussian, even though the density at any point in time is, by construction, Gaussian.</p>
Full article ">Figure 7
<p>Generalised synchrony in a Laplace system. (<b>Upper panels</b>) This figure uses the same format as <a href="#entropy-23-01220-f003" class="html-fig">Figure 3</a> but, in this instance, reporting the Laplacian approximation to coupled Lorenz systems evincing generalised synchrony. (<b>Middle panels</b>) The deterministic (solid lines) and stochastic (dotted lines) solutions of this six-dimensional system are shown in a three-dimensional state-space by plotting the three states of the coupled systems on the same axes (in blue and cyan, respectively). This illustrates the degree of synchronisation, which is particularly marked for the deterministic solutions (corresponding to identical synchronisation). (<b>Lower panel</b>) As in <a href="#entropy-23-01220-f003" class="html-fig">Figure 3</a>, the flow of the Lorenz (true) and Laplace (approximate) systems are not identical but highly correlated.</p>
Full article ">Figure 8
<p>Variational inference. (<b>Upper panel</b>) these images use the same format as <a href="#entropy-23-01220-f004" class="html-fig">Figure 4</a> to illustrate the sparsity of the coupling (in the Jacobian: left) and ensuing conditional independencies (in the Hessian: right). This sparsity structure now supports a particular partition into internal (states five and six), active (fourth state), sensory (first state) and external (second and third) states. This partition is illustrated with boxes over the Hessian: blue—internal states, red—active states, magenta—sensory states and cyan—external states. The remarkable thing here is that despite their conditional independence there are correlations between internal and external states and here, between the second and fifth states. (<b>Second panel</b>) this plots a stochastic solution of all six states as a function of time. (<b>Lower panels</b>) the correlations between the fourth (internal) and second (external) states imply one can be predicted from the other. This is illustrated by plotting the conditional expectation of the internal state, given the sensory (first) state, in the third panel, and the associated conditional density over the external state in the fourth panel. The conditional density is shown in terms of the conditional expectation (blue line) and 90% credible intervals (shaded area). The red line corresponds to the realised trajectory of the external state that lies largely within the 90% credible intervals.</p>
Full article ">Figure 9
<p>A numerical analysis of conditional independence. (<b>Upper panel</b>) The partial correlation <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>X</mi> <mi>Y</mi> <mo>·</mo> <mi>Z</mi> </mrow> </msub> </mrow> </semantics></math> (right) between each pair of states, regressing out the effect of the remaining states, approximates the Hessian of the (Laplace-approximated) coupled Lorenz system (left). In this figure, the Hessian and the partial correlation are displayed in terms of their norms (i.e., the elements squared). The partial correlation matrix was based on 128 stochastic solutions, each lasting 500 s. The upper right panel shows the average partial correlation matrix based on the entire timeseries and averaged over realisations. (<b>Lower panel</b>) partial correlations are shown for two pairs of states—the 3rd and 6th states (the second dimensions of the respective “internal states” of the first and second Lorenz systems, left side) and the 4th and 5th states (the “sensory state” and first dimension of the “internal state” of the second Lorenz system, right side). The <span class="html-italic">x</span>-axis denotes the increasing length of the timeseries used to evaluate the partial correlations. Note that the 4th and 5th states are not conditionally independent, explaining why the average partial correlation converges to a value around −0.33, whereas the 3rd and 6th states are conditionally independent, given the other states, meaning that the partial correlation converges to 0.</p>
Full article ">Figure 10
<p>Markov blankets. These influence diagrams illustrate a particular partition of states into internal states (blue) and hidden or external states (cyan) that are separated by a Markov blanket comprising sensory (magenta) and active states (red). The upper panel shows this partition as it would be applied to a single-cell organism, where internal states are associated with the intracellular states, the sensory states become the surface states or cell membrane overlying active states (e.g., the actin filaments of the cytoskeleton). The dotted lines indicate directed influences from external (respectively internal) to active (respectively sensory) states. Particular states constitute a particle; namely, autonomous and sensory states—or blanket and internal states. The lower panel illustrates how this partition applies to the six states of the coupled system considered in the main text.</p>
Full article ">
18 pages, 355 KiB  
Article
On the General Solutions of Some Non-Homogeneous Div-Curl Systems with Riemann–Liouville and Caputo Fractional Derivatives
by Briceyda B. Delgado and Jorge E. Macías-Díaz
Fractal Fract. 2021, 5(3), 117; https://doi.org/10.3390/fractalfract5030117 - 10 Sep 2021
Cited by 23 | Viewed by 1949
Abstract
In this work, we investigate analytically the solutions of a nonlinear div-curl system with fractional derivatives of the Riemann–Liouville or Caputo types. To this end, the fractional-order vector operators of divergence, curl and gradient are identified as components of the fractional Dirac operator [...] Read more.
In this work, we investigate analytically the solutions of a nonlinear div-curl system with fractional derivatives of the Riemann–Liouville or Caputo types. To this end, the fractional-order vector operators of divergence, curl and gradient are identified as components of the fractional Dirac operator in quaternionic form. As one of the most important results of this manuscript, we derive general solutions of some non-homogeneous div-curl systems that consider the presence of fractional-order derivatives of the Riemann–Liouville or Caputo types. A fractional analogous to the Teodorescu transform is presented in this work, and we employ some properties of its component operators, developed in this work to establish a generalization of the Helmholtz decomposition theorem in fractional space. Additionally, right inverses of the fractional-order curl, divergence and gradient vector operators are obtained using Riemann–Liouville and Caputo fractional operators. Finally, some consequences of these results are provided as applications at the end of this work. Full article
Back to TopTop