[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 (22)

Search Parameters:
Keywords = cylindrical symmetry condition

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
39 pages, 562 KiB  
Article
Elastic Wave Scattering off a Single and Double Array of Periodic Defects
by Omer Haq and Sergei V. Shabanov
Mathematics 2024, 12(21), 3425; https://doi.org/10.3390/math12213425 - 31 Oct 2024
Viewed by 582
Abstract
The scattering problem of elastic waves impinging on periodic single and double arrays of parallel cylindrical defects is considered for isotropic materials. An analytic expression for the scattering matrix is obtained by means of the Lippmann–Schwinger formalism and analyzed in the long-wavelength approximation. [...] Read more.
The scattering problem of elastic waves impinging on periodic single and double arrays of parallel cylindrical defects is considered for isotropic materials. An analytic expression for the scattering matrix is obtained by means of the Lippmann–Schwinger formalism and analyzed in the long-wavelength approximation. It is proved that, for a specific curve in the space of physical and geometrical parameters, the scattering is dominated by resonances. The shear mode polarized parallel to the cylinders is decoupled from the other two polarization modes due to the translational symmetry along the cylinders. It is found that a relative mass density and relative Lamé coefficients of the scatterers give opposite contributions to the width of resonances in this mode. A relation between the Bloch phase and material parameters is found to obtain a global minimum of the width. The minimal width is shown to vanish in the leading order of the long wavelength limit for the single array. This new effect is not present in similar acoustic and photonic systems. The shear and compression modes in a plane perpendicular to the cylinders are coupled due to the normal traction boundary condition and have different group velocities. For the double array, it is proved that, under certain conditions on physical and geometrical parameters, there exist resonances with the vanishing width, known as Bound States in the Continuum (BSC). Necessary and sufficient conditions for the existence of BSC are found for any number of open diffraction channels. Analytic BSC solutions are obtained. Spectral parameters of BSC are given in terms of the Bloch phase and group velocities of the shear and compression modes. Full article
(This article belongs to the Section Mathematical Physics)
Show Figures

Figure 1

Figure 1
<p>The cross section of the double array of cylinders in a plane perpendicular to the cylinders. The cylinders are infinite and parallel to the <span class="html-italic">z</span> axis. The figure displays the cross section of the cylinders in the <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </semantics></math> plane. The <span class="html-italic">x</span> axis is horizontal, and the <span class="html-italic">y</span> axis is vertical (the arrays are periodic in this direction). The length is measured in units of the period of the array. The origin is set so that the system is symmetric under the reflection <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>→</mo> <mo>−</mo> <mi>x</mi> </mrow> </semantics></math>. The axis of each cylinder is parallel to the <span class="html-italic">z</span> axis and passes through the points <math display="inline"><semantics> <mrow> <msubsup> <mi>r</mi> <mi>n</mi> <mo>±</mo> </msubsup> <mo>=</mo> <mi>n</mi> <mover accent="true"> <mi>y</mi> <mo>^</mo> </mover> <mo>±</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>d</mi> <mn>2</mn> </mfrac> </mstyle> <mover accent="true"> <mi>x</mi> <mo>^</mo> </mover> </mrow> </semantics></math> in the <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </semantics></math> plane, where <span class="html-italic">d</span> is the distance between the arrays, <span class="html-italic">n</span> ranges over all integers, and <math display="inline"><semantics> <mover accent="true"> <mi>x</mi> <mo>^</mo> </mover> </semantics></math> and <math display="inline"><semantics> <mover accent="true"> <mi>y</mi> <mo>^</mo> </mover> </semantics></math> are unit vectors parallel to the coordinate axes.</p>
Full article ">Figure 2
<p>Plot of <math display="inline"><semantics> <mrow> <mrow> <mo stretchy="false">|</mo> <mi>ϵ</mi> </mrow> <msub> <mi>S</mi> <mi>x</mi> </msub> <mrow> <mo>(</mo> <msup> <mi>ω</mi> <mn>2</mn> </msup> <mo>)</mo> </mrow> <msup> <mrow> <mo stretchy="false">|</mo> </mrow> <mn>2</mn> </msup> </mrow> </semantics></math> vs. <span class="html-italic">q</span>, where <math display="inline"><semantics> <mrow> <mi>ω</mi> <mo>=</mo> <msub> <mi>c</mi> <mi>t</mi> </msub> <mi>q</mi> </mrow> </semantics></math> for the parameters <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>k</mi> <mi>y</mi> </msub> <mo>,</mo> <mi>ϵ</mi> <mo>,</mo> <mi>d</mi> <mo>,</mo> <msub> <mi>ξ</mi> <mi>ρ</mi> </msub> <mo>,</mo> <mi>α</mi> <mo>)</mo> </mrow> </semantics></math> = (1.73405, 0.001, 29,880, 0.280, 0.59915) (black solid line). The red dashed line corresponds to a numerical Lorentzian fit. The optimal fitting parameters are found to be <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msub> <mi>ω</mi> <mn>0</mn> </msub> <mo>,</mo> <mo>Γ</mo> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>4.5490425</mn> <msub> <mi>c</mi> <mi>t</mi> </msub> <mo>,</mo> <mn>6.72071</mn> <mo>·</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> <msubsup> <mi>c</mi> <mrow> <mi>t</mi> </mrow> <mn>2</mn> </msubsup> <mo>)</mo> </mrow> </mrow> </semantics></math>. The blue dotted line corresponds to the result of the perturbation theory where <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msub> <mi>ω</mi> <mn>0</mn> </msub> <mo>,</mo> <mo>Γ</mo> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>4.5490430</mn> <msub> <mi>c</mi> <mi>t</mi> </msub> <mo>,</mo> <mn>6.62285</mn> <mo>·</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> <msubsup> <mi>c</mi> <mrow> <mi>t</mi> </mrow> <mn>2</mn> </msubsup> <mo>)</mo> </mrow> </mrow> </semantics></math> are computed to the leading order.</p>
Full article ">Figure 3
<p>Plot of <math display="inline"><semantics> <mrow> <msup> <mi>s</mi> <mo>±</mo> </msup> <mo>/</mo> <mover accent="true"> <mi>s</mi> <mo>˜</mo> </mover> </mrow> </semantics></math> vs. <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>s</mi> <mo>˜</mo> </mover> <mi>d</mi> </mrow> </semantics></math>. The red (or lower) curve is the graph of <math display="inline"><semantics> <mrow> <msup> <mi>s</mi> <mo>+</mo> </msup> <mo>/</mo> <mover accent="true"> <mi>s</mi> <mo>˜</mo> </mover> </mrow> </semantics></math>, and the black (upper) curve is the graph of <math display="inline"><semantics> <mrow> <msup> <mi>s</mi> <mo>−</mo> </msup> <mo>/</mo> <mover accent="true"> <mi>s</mi> <mo>˜</mo> </mover> </mrow> </semantics></math>.</p>
Full article ">
19 pages, 425 KiB  
Article
Inverse Boundary Conditions Interface Problems for the Heat Equation with Cylindrical Symmetry
by Miglena N. Koleva and Lubin G. Vulkov
Symmetry 2024, 16(8), 1065; https://doi.org/10.3390/sym16081065 - 18 Aug 2024
Cited by 1 | Viewed by 962
Abstract
In this paper, we study inverse interface problems with unknown boundary conditions, using point observations for parabolic equations with cylindrical symmetry. In the one-dimensional, two-layer interface problem, the left interval 0<r<l1, i.e., the zero degeneracy, causes serious [...] Read more.
In this paper, we study inverse interface problems with unknown boundary conditions, using point observations for parabolic equations with cylindrical symmetry. In the one-dimensional, two-layer interface problem, the left interval 0<r<l1, i.e., the zero degeneracy, causes serious solution difficulty. For this, we investigate the well-posedness of the direct (forward) problem. Next, we formulate and solve five inverse boundary condition problems for the interface heat equation with cylindrical symmetry from internal measurements. The finite volume difference method is developed to construct second-order schemes for direct and inverse problems. The correctness of the proposed numerical solution decomposition algorithms for the inverse problems is discussed. Several numerical examples are presented to illustrate the efficiency of the approach. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>Left</b>): Numerical solution <span class="html-italic">u</span> of the direct problem (solid red line) and recovered solution (line with blue circles); (<b>right</b>): error between the numerically recovered solution and numerical solution of the direct problem, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mi>h</mi> </mrow> </semantics></math>, IP solved by (<a href="#FD29-symmetry-16-01065" class="html-disp-formula">29</a>)–(<a href="#FD32-symmetry-16-01065" class="html-disp-formula">32</a>), Example 2.</p>
Full article ">Figure 2
<p>(<b>Left</b>): Exact function <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mi>L</mi> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (solid red line) and recovered function <math display="inline"><semantics> <msubsup> <mi>g</mi> <mi>L</mi> <mi>n</mi> </msubsup> </semantics></math> (line with blue circles); (<b>right</b>): the corresponding error, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mi>h</mi> </mrow> </semantics></math>, IP solved by (<a href="#FD29-symmetry-16-01065" class="html-disp-formula">29</a>)–(<a href="#FD32-symmetry-16-01065" class="html-disp-formula">32</a>), Example 2.</p>
Full article ">Figure 3
<p>(<b>Left</b>): Numerical solution <span class="html-italic">u</span> of the direct problem (solid red line) and recovered solution (line with blue circles); (<b>right</b>): error between the numerically recovered solution and numerical solution of the direct problem, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mi>h</mi> </mrow> </semantics></math>, IP solved by (<a href="#FD37-symmetry-16-01065" class="html-disp-formula">37</a>)–(<a href="#FD42-symmetry-16-01065" class="html-disp-formula">42</a>), Example 2.</p>
Full article ">Figure 4
<p>(<b>Left</b>): Exact function <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (solid red line) and recovered function <math display="inline"><semantics> <msubsup> <mi>g</mi> <mn>0</mn> <mi>n</mi> </msubsup> </semantics></math> (line with blue circles); (<b>right</b>): the corresponding error, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mi>h</mi> </mrow> </semantics></math>, IP solved by (<a href="#FD37-symmetry-16-01065" class="html-disp-formula">37</a>)–(<a href="#FD42-symmetry-16-01065" class="html-disp-formula">42</a>), Example 2.</p>
Full article ">Figure 5
<p>(<b>Left</b>): Exact function <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mi>L</mi> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (solid red line) and recovered function <math display="inline"><semantics> <msubsup> <mi>g</mi> <mi>L</mi> <mi>n</mi> </msubsup> </semantics></math> (line with blue circles); (<b>right</b>): the corresponding error, <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mi>h</mi> </mrow> </semantics></math>, IP solved by (<a href="#FD37-symmetry-16-01065" class="html-disp-formula">37</a>)–(<a href="#FD42-symmetry-16-01065" class="html-disp-formula">42</a>), Example 2.</p>
Full article ">Figure 6
<p>(<b>Left</b>): Exact function <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mi>L</mi> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (solid red line) and recovered function <math display="inline"><semantics> <msubsup> <mi>g</mi> <mi>L</mi> <mi>n</mi> </msubsup> </semantics></math> (line with blue circles); (<b>right</b>): the corresponding error, <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, IP solved by (<a href="#FD37-symmetry-16-01065" class="html-disp-formula">37</a>)–(<a href="#FD42-symmetry-16-01065" class="html-disp-formula">42</a>), Example 3.</p>
Full article ">Figure 7
<p>(<b>Left</b>): Exact function <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (solid red line) and recovered function <math display="inline"><semantics> <msubsup> <mi>g</mi> <mn>0</mn> <mi>n</mi> </msubsup> </semantics></math> (line with blue circles); (<b>right</b>): the corresponding error, <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.003</mn> </mrow> </semantics></math>, IP solved by (<a href="#FD37-symmetry-16-01065" class="html-disp-formula">37</a>)–(<a href="#FD42-symmetry-16-01065" class="html-disp-formula">42</a>), Example 3.</p>
Full article ">Figure 8
<p>(<b>Left</b>): Exact function <math display="inline"><semantics> <mrow> <msub> <mi>g</mi> <mi>L</mi> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (solid red line) and recovered function <math display="inline"><semantics> <msubsup> <mi>g</mi> <mi>L</mi> <mi>n</mi> </msubsup> </semantics></math> (line with blue circles); (<b>right</b>): the corresponding error, <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.003</mn> </mrow> </semantics></math>, IP solved by (<a href="#FD37-symmetry-16-01065" class="html-disp-formula">37</a>)–(<a href="#FD42-symmetry-16-01065" class="html-disp-formula">42</a>), Example 3.</p>
Full article ">
21 pages, 4570 KiB  
Article
A Static Stability Analysis Method for Passively Stabilized Sounding Rockets
by Riccardo Cadamuro, Maria Teresa Cazzola, Nicolò Lontani and Carlo E. D. Riboldi
Aerospace 2024, 11(3), 242; https://doi.org/10.3390/aerospace11030242 - 20 Mar 2024
Viewed by 2334
Abstract
Sounding rockets constitute a class of rocket with a generally simple layout, being composed of a cylindrical center-body, a nosecone, a number of fins placed symmetrically around the longitudinal axis (usually three or four), and possibly a boat-tail. This type of flying craft [...] Read more.
Sounding rockets constitute a class of rocket with a generally simple layout, being composed of a cylindrical center-body, a nosecone, a number of fins placed symmetrically around the longitudinal axis (usually three or four), and possibly a boat-tail. This type of flying craft is typically not actively controlled; instead, a passive stabilization effect is obtained through suitable positioning and sizing of the fins. Therefore, in the context of dynamic performance analysis, the margin of static stability is an index of primary interest. However, the classical approach to static stability analysis, which consists in splitting computations in two decoupled domains, namely, around the pitch and yaw axis, provides a very limited insight to the missile performance for this type of vehicle due to the violation of the classical assumptions of planar symmetry and symmetric flight conditions commonly adopted for winged aircraft. To tackle this issue, this paper introduces a method for analyzing static stability through a novel index, capable of more generally assessing the level of static stability for sounding rockets, exploiting the same information on aerodynamic coefficients typically required for more usual (i.e., decoupled) static stability analyses, and suggests a way to assess the validity and shortcoming of the method in each case at hand. Full article
(This article belongs to the Special Issue Aircraft Modeling, Simulation and Control II)
Show Figures

Figure 1

Figure 1
<p>Rocket body reference frame.</p>
Full article ">Figure 2
<p>Two sets of equivalent aerodynamic angles to describe the direction of a perturbation. (<b>a</b>) <math display="inline"><semantics> <mi>α</mi> </semantics></math> and <math display="inline"><semantics> <mi>β</mi> </semantics></math>. (<b>b</b>) <math display="inline"><semantics> <msub> <mi>α</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> </mrow> </msub> </semantics></math> and <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math>.</p>
Full article ">Figure 3
<p>Longitudinal and directional static margins as functions of the aerodynamic roll angle <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math> for a four-finned rocket.</p>
Full article ">Figure 4
<p>Cross-derivatives of the aerodynamic pitching and yawing moment of a four-finned rocket for changing values of the aerodynamic angles.</p>
Full article ">Figure 5
<p>Longitudinal and directional static margins as functions of the aerodynamic roll angle <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math> for a three-finned rocket.</p>
Full article ">Figure 6
<p>Cross-derivatives of the aerodynamic pitching and yawing moment of a three-finned rocket for changing values of the aerodynamic angles.</p>
Full article ">Figure 7
<p>(<b>a</b>) Rotated forces; (<b>b</b>) rotated moments.</p>
Full article ">Figure 8
<p>Description of parameters in <a href="#aerospace-11-00242-t003" class="html-table">Table 3</a>.</p>
Full article ">Figure 9
<p>The considered layouts, visualized within <tt>Tucan OpenVogel</tt>: (<b>a</b>) no dihedral wing, (<b>b</b>) +20° dihedral wing, (<b>c</b>) −20° dihedral wing, and (<b>d</b>) set of three fins.</p>
Full article ">Figure 10
<p>Normal components of the velocity (red), force (black), and moment (blue).</p>
Full article ">Figure 11
<p>Definition of angle <math display="inline"><semantics> <mo>Δ</mo> </semantics></math>.</p>
Full article ">Figure 12
<p>Parameter <span class="html-italic">T</span> as a function of the aerodynamic roll angle.</p>
Full article ">Figure 13
<p>Axial symmetric slender body of revolution.</p>
Full article ">Figure 14
<p>Data at <span class="html-italic">v</span> = 32 m/s and <math display="inline"><semantics> <msub> <mi>α</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> </mrow> </msub> </semantics></math> = 16 deg: (<b>a</b>) force coefficients; (<b>b</b>) moment coefficient; (<b>c</b>) force coefficients rotated; (<b>d</b>) moment coefficients rotated.</p>
Full article ">Figure 15
<p>Comparison between the classical and the compounded static margins for the three-finned missile.</p>
Full article ">Figure 16
<p>Distribution of <span class="html-italic">T</span> index, considering all testing conditions and configurations in <a href="#aerospace-11-00242-t004" class="html-table">Table 4</a> and <a href="#aerospace-11-00242-t005" class="html-table">Table 5</a>.</p>
Full article ">Figure 17
<p>(<b>a</b>) <span class="html-italic">Gemini</span> lift-off from the launch rail; (<b>b</b>) <span class="html-italic">Gemini</span> airborne during the powered ascent phase.</p>
Full article ">Figure 18
<p>Longitudinal (<b>a</b>), directional (<b>b</b>), and compounded (<b>c</b>) static margin behavior as functions of the wind intensity and direction.</p>
Full article ">Figure A1
<p>Comparison between the values of <math display="inline"><semantics> <msub> <mi>C</mi> <mi>Y</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>C</mi> <mi>N</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>C</mi> <msub> <mi>m</mi> <mi mathvariant="bold-italic">CG</mi> </msub> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>C</mi> <msub> <mi>n</mi> <mi mathvariant="bold-italic">CG</mi> </msub> </msub> </semantics></math> and the longitudinal and directional stability margins <math display="inline"><semantics> <mrow> <mi>S</mi> <msub> <mi>M</mi> <mrow> <mi>l</mi> <mi>o</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>S</mi> <msub> <mi>M</mi> <mrow> <mi>d</mi> <mi>i</mi> <mi>r</mi> </mrow> </msub> </mrow> </semantics></math> for the reference three-finned configuration, obtained from <tt>Missile DATCOM 97</tt> and CFD simulations in <tt>ANSYS Fluent 2022 R1</tt>.</p>
Full article ">Figure A2
<p>Comparison between the values of the compound static margin and the <span class="html-italic">T</span> index for the reference three-finned configuration, obtained from <tt>Missile DATCOM 97</tt> and CFD simulations in <tt>ANSYS Fluent 2022 R1</tt>.</p>
Full article ">Figure A3
<p>Fluid domain hexahedral dominant mesh. (<b>a</b>) Polyhedral elements near the body; (<b>b</b>) hexahedral elements sufficiently far from it.</p>
Full article ">
25 pages, 736 KiB  
Article
Linear Stability Analysis of Relativistic Magnetized Jets: Methodology
by Nektarios Vlahakis
Universe 2023, 9(9), 386; https://doi.org/10.3390/universe9090386 - 26 Aug 2023
Cited by 4 | Viewed by 1044
Abstract
The stability of astrophysical jets in the linear regime is investigated by presenting a methodology to find the growth rates of the various instabilities. We perturb a cylindrical axisymmetric steady jet, linearize the relativistic ideal magnetohydrodynamic (MHD) equations, and analyze the evolution of [...] Read more.
The stability of astrophysical jets in the linear regime is investigated by presenting a methodology to find the growth rates of the various instabilities. We perturb a cylindrical axisymmetric steady jet, linearize the relativistic ideal magnetohydrodynamic (MHD) equations, and analyze the evolution of the eigenmodes of the perturbation by deriving the differential equations that need to be integrated, subject to the appropriate boundary conditions, in order to find the dispersion relation. We also apply the WKBJ approximation and, additionally, give analytical solutions in some subcases corresponding to unperturbed jets with constant bulk velocity along the symmetry axis. Full article
(This article belongs to the Section Compact Objects)
Show Figures

Figure 1

Figure 1
<p>The unperturbed state for the example case in <a href="#sec6-universe-09-00386" class="html-sec">Section 6</a>.</p>
Full article ">Figure 2
<p>Left panel: the solutions of the dispersion relation. Right panels: the eigenfunctions for the three modes.</p>
Full article ">Figure 3
<p>Left panel: the numerical solution for the case of <a href="#universe-09-00386-f002" class="html-fig">Figure 2</a>. right panel: Similarly for the nonaxisymmetric mode with <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>The WKBJ solution for the case in <a href="#universe-09-00386-f002" class="html-fig">Figure 2</a>.</p>
Full article ">
19 pages, 5536 KiB  
Article
A Station-Keeping Control Strategy for a Symmetrical Spacecraft Utilizing Hybrid Low-Thrust Propulsion in the Heliocentric Displaced Orbit
by Tengfei Zhang, Rongjun Mu, Yilin Zhou, Zizheng Liao, Zhewei Zhang, Bo Liao and Chuang Yao
Symmetry 2023, 15(8), 1549; https://doi.org/10.3390/sym15081549 - 6 Aug 2023
Viewed by 1173
Abstract
The solar sail spacecraft utilizing a hybrid approach of solar sail and solar electric propulsion in the heliocentric displaced orbit is affected by external disturbances, internal unmodeled dynamics, initial injection errors, and input saturation. To solve the station-keeping control problem under such complex [...] Read more.
The solar sail spacecraft utilizing a hybrid approach of solar sail and solar electric propulsion in the heliocentric displaced orbit is affected by external disturbances, internal unmodeled dynamics, initial injection errors, and input saturation. To solve the station-keeping control problem under such complex conditions, an adaptive control strategy is proposed. First, the dynamical equations of the spacecraft utilizing hybrid low-thrust propulsion in the cylindrical coordinate system are derived. Second, the combined disturbance acceleration introduced by external disturbances and internal unmodeled dynamics is constructed, and a radial basis function neural network estimator is designed to estimate it online in real time. Third, an adaptive high-performance station-keeping controller based on an improved integral sliding surface and multivariate super-twisting sliding mode approaching law is designed. Then, stability analysis is conducted using Lyapunov theory, adaptive laws are designed, and the introduced virtual control accelerations are converted into actual control variables. Finally, simulations are conducted under different simulation conditions based on the disturbance sources. The results show that although the use of hybrid low-thrust propulsion breaks the symmetry of the solar sail in configuration, the proposed control strategy can effectively achieve the station-keeping and disturbance estimation of the spacecraft with only a small amount of propellant consumed and position tracking errors up to decimeters. Full article
(This article belongs to the Special Issue Advances in Mechanics and Control II)
Show Figures

Figure 1

Figure 1
<p>Displaced orbit and orbital coordinate system of the solar sail spacecraft.</p>
Full article ">Figure 2
<p>Block diagram of the station-keeping control system for a solar sail spacecraft.</p>
Full article ">Figure 3
<p>Change curves of improved integral sliding surfaces.</p>
Full article ">Figure 4
<p>Reference orbit and controlled trajectory.</p>
Full article ">Figure 5
<p>Change curves of position errors and angular velocity: (<b>a</b>) change curves of orbital radius error and displaced height error; (<b>b</b>) change curve of angular velocity.</p>
Full article ">Figure 6
<p>Change curves of external disturbances and their estimations.</p>
Full article ">Figure 7
<p>Station-keeping control variables: (<b>a</b>) change curves of attitude angles; (<b>b</b>) change curves of propulsion force components of argon Hall thruster.</p>
Full article ">Figure 8
<p>Mass variation in a spacecraft under different propulsion methods.</p>
Full article ">Figure 9
<p>Change curves of position errors and angular velocity: (<b>a</b>) change curves of orbital radius error and displaced height error; (<b>b</b>) change curve of angular velocity.</p>
Full article ">Figure 10
<p>Change curves of external disturbances and their estimations.</p>
Full article ">Figure 11
<p>Station-keeping control variables: (<b>a</b>) change curves of attitude angles; (<b>b</b>) change curves of propulsion force components of argon Hall thruster.</p>
Full article ">Figure 12
<p>Mass variation in a spacecraft.</p>
Full article ">
11 pages, 310 KiB  
Article
Noether Symmetries and Conservation Laws in Static Cylindrically Symmetric Spacetimes via Rif Tree Approach
by Muhammad Farhan, Suhad Subhi Aiadi, Tahir Hussain and Nabil Mlaiki
Symmetry 2023, 15(1), 184; https://doi.org/10.3390/sym15010184 - 8 Jan 2023
Cited by 2 | Viewed by 1366
Abstract
A new approach is adopted to completely classify the Lagrangian associated with the static cylindrically symmetric spacetime metric via Noether symmetries. The determining equations representing Noether symmetries are analyzed using a Maple algorithm that imposes different conditions on metric coefficients under which static [...] Read more.
A new approach is adopted to completely classify the Lagrangian associated with the static cylindrically symmetric spacetime metric via Noether symmetries. The determining equations representing Noether symmetries are analyzed using a Maple algorithm that imposes different conditions on metric coefficients under which static cylindrically symmetric spacetimes admit Noether symmetries of different dimensions. These conditions are used to solve the determining equations, giving the explicit form of vector fields representing Noether symmetries. The obtained Noether symmetry generators are used in Noether’s theorem to find the expressions for corresponding conservation laws. The singularity of the obtained metrics is discussed by finding their Kretschmann scalar. Full article
(This article belongs to the Special Issue Noether and Space-Time Symmetries in Physics)
Show Figures

Figure 1

Figure 1
<p>Rif tree.</p>
Full article ">
13 pages, 3429 KiB  
Article
Using Cylindrical and Spherical Symmetries in Numerical Simulations of Quasi-Infinite Mechanical Systems
by Alexander E. Filippov and Valentin L. Popov
Symmetry 2022, 14(8), 1557; https://doi.org/10.3390/sym14081557 - 28 Jul 2022
Cited by 3 | Viewed by 1845
Abstract
The application of cylindrical and spherical symmetries for numerical studies of many-body problems is presented. It is shown that periodic boundary conditions corresponding to formally cylindrical symmetry allow for reducing the problem of a huge number of interacting particles, minimizing the effect of [...] Read more.
The application of cylindrical and spherical symmetries for numerical studies of many-body problems is presented. It is shown that periodic boundary conditions corresponding to formally cylindrical symmetry allow for reducing the problem of a huge number of interacting particles, minimizing the effect of boundary conditions, and obtaining reasonably correct results from a practical point of view. A physically realizable cylindrical configuration is also studied. The advantages and disadvantages of symmetric realizations are discussed. Finally, spherical symmetry, which naturally realizes a three-dimensional system without boundaries on its two-dimensional surface, is studied. As an example, tectonic dynamics are considered, and interesting patterns resembling real ones are found. It is stressed that perturbations of the axis of planet rotation may be responsible for the formation of such patterns. Full article
(This article belongs to the Special Issue Axisymmetry in Mechanical Engineering)
Show Figures

Figure 1

Figure 1
<p>Conceptual picture of a system with periodic boundary conditions in one-dimensional and repulsing boundaries in the orthogonal one. The origin of the particular system and meaning of the colors in different subplots are described in the main text.</p>
Full article ">Figure 2
<p>Localized excitation moving along rigid stripe inserted between two soft layers. It is seen directly that velocity of the rigid stripe inserted between two soft ones (green color) is almost constant and much smaller than velocities of the outer layers moving in opposite directions (blue and red colors, respectively).</p>
Full article ">Figure 3
<p>Square root <math display="inline"><semantics> <mrow> <mi>h</mi> <msup> <mrow> <mo stretchy="false">(</mo> <mo>|</mo> <msub> <mi>p</mi> <mi>i</mi> </msub> <mo>|</mo> <mo stretchy="false">)</mo> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math> of the histogram <math display="inline"><semantics> <mrow> <mi>h</mi> <mo stretchy="false">(</mo> <mo>|</mo> <msub> <mi>p</mi> <mi>i</mi> </msub> <mo>|</mo> <mo stretchy="false">)</mo> </mrow> </semantics></math> (<b>a</b>) and the time dependence of the fraction of the points for which the inequality <math display="inline"><semantics> <mrow> <mo>|</mo> <msub> <mi>p</mi> <mi>i</mi> </msub> <mo>|</mo> <mo>&gt;</mo> <mi>p</mi> <mo>*</mo> </mrow> </semantics></math> holds (<b>b</b>). Note the bursts in the magnitude of this fraction corresponding to the peaks in the time dependence of the total force <math display="inline"><semantics> <mrow> <msub> <mi>F</mi> <mrow> <mi>e</mi> <mi>x</mi> <mi>t</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <mi>P</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>/</mo> <mi>V</mi> </mrow> </semantics></math>, which is presented in (<b>c</b>).</p>
Full article ">Figure 4
<p>Time–space map of the horizontal velocity of selected vertical cross-section <math display="inline"><semantics> <mrow> <msub> <mi>v</mi> <mi>x</mi> </msub> <mo stretchy="false">(</mo> <mi>t</mi> <mo>,</mo> <mi>y</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> recording static information about history of the events (wave propagation) in the system.</p>
Full article ">Figure 5
<p>Time-dependent number of rigid phase domains, histogram of area occupied by the domains of the given size, shown accumulated during long-duration run, as well size of the maximal and averaged domain.</p>
Full article ">Figure 6
<p>Instant configuration of the cylindrical quasi-infinite system presented in the Movie_6.avi [<a href="#B15-symmetry-14-01557" class="html-bibr">15</a>].</p>
Full article ">Figure 7
<p>Sequence of the instant configurations of self-organized spherical quasi-infinite system. Early (starting from the randomly distributed particles on the sphere), two intermediate, and close to the stationary moving configurations are shown in the consequent subplots (<b>a</b>–<b>d</b>), respectively. Parallel belts of the vortices and anti-vortices in the “equatorial” region corresponding to the positive and negative rotor projection perpendicular to the spherical surface (shown by red and blue colors, respectively) are seen directly.</p>
Full article ">Figure 8
<p>Typical intermediate configuration of the two-component spherical system reproduced in Movie_09. Spatial distribution of the rigid and soft (brown and blue, respectively) components is reproduced in subplot (<b>a</b>). Corresponding to the present configuration, distributions of the local velocities and forces are shown in subplots (<b>b</b>,<b>c</b>), respectively. A tendency of the concentration of the “continents” to “equator” at fixed axis of the rotation is seen directly.</p>
Full article ">
26 pages, 7467 KiB  
Article
Hydrodynamic Impacts of Short Laser Pulses on Plasmas
by Gaetano Fiore, Monica De Angelis, Renato Fedele, Gabriele Guerriero and Dušan Jovanović
Mathematics 2022, 10(15), 2622; https://doi.org/10.3390/math10152622 - 27 Jul 2022
Cited by 3 | Viewed by 1485
Abstract
We determine conditions allowing for simplification of the description of the impact of a short and arbitrarily intense laser pulse onto a cold plasma at rest. If both the initial plasma density and pulse profile have plane symmetry, then suitable matched upper bounds [...] Read more.
We determine conditions allowing for simplification of the description of the impact of a short and arbitrarily intense laser pulse onto a cold plasma at rest. If both the initial plasma density and pulse profile have plane symmetry, then suitable matched upper bounds on the maximum and the relative variations of the initial density, as well as on the intensity and duration of the pulse, ensure a strictly hydrodynamic evolution of the electron fluid without wave-breaking or vacuum-heating during its whole interaction with the pulse, while ions can be regarded as immobile. We use a recently developed fully relativistic plane model whereby the system of the Lorentz–Maxwell and continuity PDEs is reduced into a family of highly nonlinear but decoupled systems of non-autonomous Hamilton equations with one degree of freedom, the light-like coordinate ξ=ctz instead of time t as an independent variable, and new a priori estimates (eased by use of a Liapunov function) of the solutions in terms of the input data (i.e., the initial density and pulse profile). If the laser spot radius R is finite and is not too small, the same conclusions hold for the part of the plasma close to the axis z of cylindrical symmetry. These results may help in drastically simplifying the study of extreme acceleration mechanisms of electrons. Full article
(This article belongs to the Topic Fluid Mechanics)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Normalized amplitude of a linearly polarized (i.e., set <math display="inline"><semantics> <mrow> <mi>ψ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> in (<a href="#FD58-mathematics-10-02622" class="html-disp-formula">58</a>)) monochromatic laser pulse slowly modulated by a Gaussian with <span class="html-italic">full width at half maximum</span><math display="inline"><semantics> <msup> <mi>l</mi> <mo>′</mo> </msup> </semantics></math> and peak amplitude <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mspace width="-0.166667em"/> <mo>≡</mo> <mspace width="-0.166667em"/> <mi>λ</mi> <mi>e</mi> <msubsup> <mi>E</mi> <mstyle scriptlevel="2" displaystyle="false"> <mi>M</mi> </mstyle> <mstyle scriptlevel="2" displaystyle="false"> <mo>⊥</mo> </mstyle> </msubsup> <mo>/</mo> <mi>m</mi> <msup> <mi>c</mi> <mn>2</mn> </msup> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <mn>1.3</mn> </mrow> </semantics></math>; this yields a moderately relativistic electron dynamics, and <math display="inline"><semantics> <mrow> <msub> <mo>Δ</mo> <mi>u</mi> </msub> <mo>≡</mo> <mo>Δ</mo> <msup> <mspace width="-0.166667em"/> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mstyle> </msup> <mrow> <mo>(</mo> <mi>l</mi> <mo>)</mo> </mrow> <mo>≃</mo> <mn>0.45</mn> <msup> <mi>l</mi> <mo>′</mo> </msup> </mrow> </semantics></math>. If <math display="inline"><semantics> <mrow> <msup> <mi>l</mi> <mo>′</mo> </msup> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <mn>7.5</mn> <mspace width="3.33333pt"/> <mi mathvariant="sans-serif">μ</mi> </mrow> </semantics></math>m (corresponding to a pulse duration of <math display="inline"><semantics> <mrow> <msup> <mi>τ</mi> <mo>′</mo> </msup> <mo>=</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mi>c</mi> <mo>≃</mo> <mn>2.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>14</mn> </mrow> </msup> </mrow> </semantics></math> s), and the wavelength is <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>0.8</mn> <mspace width="3.33333pt"/> <mi mathvariant="sans-serif">μ</mi> </mrow> </semantics></math>m, then the corresponding peak intensity must be <math display="inline"><semantics> <mrow> <mi>I</mi> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <mn>7.25</mn> <mspace width="-0.166667em"/> <mo>×</mo> <mspace width="-0.166667em"/> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mn>18</mn> </msup> </mrow> </semantics></math> W/cm<math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math>; these are typical values obtainable by Ti:Sapphire lasers in LWFA experiments. (<b>b</b>) The corresponding forcing term <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>(</mo> <mi>ξ</mi> <mo>)</mo> </mrow> </semantics></math> and average-over-cycle (<a href="#FD60-mathematics-10-02622" class="html-disp-formula">60</a>) <math display="inline"><semantics> <mrow> <msub> <mi>v</mi> <mi>a</mi> </msub> <mrow> <mo>(</mo> <mi>ξ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> of the latter. (<b>c</b>) Corresponding solution of (<a href="#FD8-mathematics-10-02622" class="html-disp-formula">8</a>) and (<a href="#FD9-mathematics-10-02622" class="html-disp-formula">9</a>), or equivalently of (<a href="#FD14-mathematics-10-02622" class="html-disp-formula">14</a>) if <math display="inline"><semantics> <mrow> <mi>K</mi> <msub> <mi>n</mi> <mn>0</mn> </msub> <msup> <mi>l</mi> <mo>′</mo> </msup> <msup> <mrow/> <mn>2</mn> </msup> <mo>≃</mo> <mn>4</mn> </mrow> </semantics></math>; this value is obtained if <math display="inline"><semantics> <mrow> <msup> <mi>l</mi> <mo>′</mo> </msup> <mspace width="-0.166667em"/> <mo>=</mo> <mspace width="-0.166667em"/> <mn>7.5</mn> <mspace width="3.33333pt"/> <mi mathvariant="sans-serif">μ</mi> </mrow> </semantics></math>m and <math display="inline"><semantics> <mrow> <mover accent="true"> <msub> <mi>n</mi> <mn>0</mn> </msub> <mo>˜</mo> </mover> <mrow> <mo>(</mo> <mi>Z</mi> <mo>)</mo> </mrow> <mspace width="-0.166667em"/> <mo>≡</mo> <mspace width="-0.166667em"/> <msub> <mi>n</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>2</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>18</mn> </msup> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math> (a typical value of the electron density used in LWFA experiments). As expected, <span class="html-italic">s</span> is insensitive to the rapid oscillations of <math display="inline"><semantics> <msup> <mrow> <mi mathvariant="bold">ϵ</mi> </mrow> <mstyle scriptlevel="2" displaystyle="false"> <mo>⊥</mo> </mstyle> </msup> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>ξ</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mi>l</mi> <mo>]</mo> </mrow> </semantics></math>, while for <math display="inline"><semantics> <mrow> <mi>ξ</mi> <mo>≥</mo> <mi>l</mi> </mrow> </semantics></math> the energy <span class="html-italic">H</span> is conserved and the solution is periodic. The length <span class="html-italic">l</span> is determined on physical grounds; if, e.g., the plasma is created locally by the impact of the pulse itself on a gas (e.g., hydrogen or helium), then <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mi>l</mi> <mo>]</mo> </mrow> </semantics></math> has to contain all points <math display="inline"><semantics> <mi>ξ</mi> </semantics></math> where the pulse intensity is sufficient to ionize the gas. Here, for simplicity and following convention, we have fixed it to be <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>4</mn> <msup> <mi>l</mi> <mo>′</mo> </msup> </mrow> </semantics></math>; the possible inaccuracy of such a cut is very small, because <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mrow> <mo>(</mo> <msup> <mn>0</mn> <mo>+</mo> </msup> <mo>)</mo> </mrow> <mo>=</mo> <mi>ϵ</mi> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>−</mo> </msup> <mo>)</mo> </mrow> </mrow> </semantics></math> is <math display="inline"><semantics> <msup> <mn>2</mn> <mrow> <mo>−</mo> <mn>16</mn> </mrow> </msup> </semantics></math> times the maximum <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>(</mo> <mi>l</mi> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> </semantics></math> of the modulation, i.e., practically zero, which makes <math display="inline"><semantics> <mrow> <msub> <mi>G</mi> <mi>b</mi> </msub> <mo>≡</mo> <msqrt> <mrow> <mi>K</mi> <msub> <mi>n</mi> <mn>0</mn> </msub> </mrow> </msqrt> <mspace width="0.166667em"/> <mi>l</mi> <mo>≃</mo> <mn>8</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Plots of the ratios <math display="inline"><semantics> <mrow> <mover accent="true"> <msub> <mi>n</mi> <mn>0</mn> </msub> <mo>˜</mo> </mover> <mo>/</mo> <msub> <mi>n</mi> <mi>b</mi> </msub> </mrow> </semantics></math> for the following initial densities: (0) <math display="inline"><semantics> <mrow> <mover accent="true"> <msub> <mi>n</mi> <mn>0</mn> </msub> <mo>˜</mo> </mover> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>n</mi> <mi>b</mi> </msub> <mspace width="0.166667em"/> <mi>θ</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>. (1) <math display="inline"><semantics> <mrow> <mover accent="true"> <msub> <mi>n</mi> <mn>0</mn> </msub> <mo>˜</mo> </mover> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>=</mo> <mfrac> <mn>1</mn> <mn>2</mn> </mfrac> <msub> <mi>n</mi> <mi>b</mi> </msub> <mspace width="0.166667em"/> <mi>θ</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mfenced separators="" open="[" close="]"> <mn>1</mn> <mspace width="-0.166667em"/> <mo>+</mo> <mspace width="-0.166667em"/> <mi>θ</mi> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <mi>z</mi> <mo>)</mo> </mrow> <mspace width="0.166667em"/> <mi>z</mi> <mo>/</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mspace width="-0.166667em"/> <mo>+</mo> <mi>θ</mi> <mrow> <mo>(</mo> <mi>z</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> </mfenced> </mrow> </semantics></math>. (2) <math display="inline"><semantics> <mrow> <mover accent="true"> <msub> <mi>n</mi> <mn>0</mn> </msub> <mo>˜</mo> </mover> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>n</mi> <mi>b</mi> </msub> <mspace width="0.166667em"/> <mfenced separators="" open="[" close="]"> <mi>θ</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mspace width="0.166667em"/> <mi>θ</mi> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <mi>z</mi> <mo>)</mo> </mrow> <mspace width="0.166667em"/> <mi>z</mi> <mo>/</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mspace width="-0.166667em"/> <mo>+</mo> <mi>θ</mi> <mrow> <mo>(</mo> <mi>z</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> </mfenced> </mrow> </semantics></math>. (3) <math display="inline"><semantics> <mrow> <mover accent="true"> <msub> <mi>n</mi> <mn>0</mn> </msub> <mo>˜</mo> </mover> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>n</mi> <mi>b</mi> </msub> <mspace width="0.166667em"/> <mfenced separators="" open="{" close="}"> <mfrac> <mi>z</mi> <mover accent="true"> <mi>z</mi> <mo>¯</mo> </mover> </mfrac> <mspace width="0.166667em"/> <mfrac> <mrow> <mi>f</mi> <mo>(</mo> <mover accent="true"> <mi>z</mi> <mo>¯</mo> </mover> <mo>)</mo> </mrow> <mrow> <mn>1</mn> <mspace width="-0.166667em"/> <mo>+</mo> <mspace width="-0.166667em"/> <mi>f</mi> <mo>(</mo> <mover accent="true"> <mi>z</mi> <mo>¯</mo> </mover> <mo>)</mo> </mrow> </mfrac> <mi>θ</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mi>θ</mi> <mrow> <mo>(</mo> <mover accent="true"> <mi>z</mi> <mo>¯</mo> </mover> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <mi>z</mi> <mo>)</mo> </mrow> <mo>+</mo> <mi>θ</mi> <mrow> <mo>(</mo> <mi>z</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <mover accent="true"> <mi>z</mi> <mo>¯</mo> </mover> <mo>)</mo> </mrow> <mspace width="0.166667em"/> <mfrac> <mrow> <mi>f</mi> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mrow> <mn>1</mn> <mspace width="-0.166667em"/> <mo>+</mo> <mspace width="-0.166667em"/> <mi>f</mi> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> </mfrac> </mfenced> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>f</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>:</mo> <mo>=</mo> <msup> <mrow> <mo>(</mo> <mn>0.1</mn> <mspace width="0.166667em"/> <mi>z</mi> <mo>/</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> <mn>2</mn> </msup> <mo>+</mo> <msup> <mrow> <mo>(</mo> <mn>0.2</mn> <mspace width="0.166667em"/> <mi>z</mi> <mo>/</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> <mn>4</mn> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>z</mi> <mo>¯</mo> </mover> <mo>=</mo> <mn>6.5</mn> <msup> <mi>l</mi> <mo>′</mo> </msup> </mrow> </semantics></math>; this grows as <span class="html-italic">z</span> for <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>≤</mo> <mover accent="true"> <mi>z</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> and coincides with the next one for <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>&gt;</mo> <mover accent="true"> <mi>z</mi> <mo>¯</mo> </mover> </mrow> </semantics></math>. (4) <math display="inline"><semantics> <mrow> <mover accent="true"> <msub> <mi>n</mi> <mn>0</mn> </msub> <mo>˜</mo> </mover> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>n</mi> <mi>b</mi> </msub> <mspace width="0.166667em"/> <mi>θ</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mfrac> <mrow> <mi>f</mi> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mrow> <mn>1</mn> <mspace width="-0.166667em"/> <mo>+</mo> <mspace width="-0.166667em"/> <mi>f</mi> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> </mfrac> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>The initial electron densities (1), (2) of <a href="#mathematics-10-02622-f002" class="html-fig">Figure 2</a> (first line; left and right, respectively). Below, assuming <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>b</mi> </msub> <mo>=</mo> <mn>2</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>18</mn> </msup> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>, we plot the corresponding <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>,</mo> <mo>Δ</mo> </mrow> </semantics></math>, their upper and lower bounds <math display="inline"><semantics> <mrow> <msup> <mi>s</mi> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </mstyle> </msup> <mo>,</mo> <msup> <mi>s</mi> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </mstyle> </msup> <mo>,</mo> <mo>Δ</mo> <msup> <mspace width="-0.166667em"/> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </mstyle> </msup> <mo>,</mo> <mo>Δ</mo> <msup> <mspace width="-0.166667em"/> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </mstyle> </msup> </mrow> </semantics></math>, and the function <span class="html-italic">Q</span>, vs. <math display="inline"><semantics> <mi>ξ</mi> </semantics></math> during interaction with the pulse in <a href="#mathematics-10-02622-f001" class="html-fig">Figure 1</a> for the same sample values, <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>20</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> of <span class="html-italic">Z</span>. The values <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mi>Z</mi> <mo>)</mo> </mrow> <mspace width="-0.166667em"/> <mo>:</mo> <mo>=</mo> <mspace width="-0.166667em"/> <mi>Q</mi> <mrow> <mo>(</mo> <mi>l</mi> <mo>,</mo> <mspace width="-0.166667em"/> <mi>Z</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> can be read off the plots. As can be seen, the bounds are much better for density (1); the values <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mi>Z</mi> <mo>)</mo> </mrow> <mo>≲</mo> <mn>1</mn> </mrow> </semantics></math> are consistent with all worldlines intersecting rather far from the laser-plasma interaction spacetime region. On the other hand, the large value of <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mi>Z</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for density (2) is an indication that worldlines intersect within or not far from the laser–plasma interaction spacetime region. Our computations lead to <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>20</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>10.64</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>=</mo> </mrow> </semantics></math> with density (1) and <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>20</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>88.53</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>32.35</mn> </mrow> </semantics></math> with density (2).</p>
Full article ">Figure 4
<p>The initial electron densities (3), (4) of <a href="#mathematics-10-02622-f002" class="html-fig">Figure 2</a> (left and right, respectively) with <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>b</mi> </msub> <mo>=</mo> <mn>2</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>18</mn> </msup> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>, corresponding plots of <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>,</mo> <mo>Δ</mo> </mrow> </semantics></math>, their upper and lower bounds <math display="inline"><semantics> <mrow> <msup> <mi>s</mi> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </mstyle> </msup> <mo>,</mo> <msup> <mi>s</mi> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </mstyle> </msup> <mo>,</mo> <mo>Δ</mo> <msup> <mspace width="-0.166667em"/> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </mstyle> </msup> <mo>,</mo> <mo>Δ</mo> <msup> <mspace width="-0.166667em"/> <mstyle scriptlevel="2" displaystyle="false"> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </mstyle> </msup> </mrow> </semantics></math>, and the function <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>(</mo> <mi>ξ</mi> <mo>,</mo> <mi>Z</mi> <mo>)</mo> </mrow> </semantics></math>, vs. <math display="inline"><semantics> <mi>ξ</mi> </semantics></math> for the same sample values <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>20</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> of <span class="html-italic">Z</span>. The values <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mi>Z</mi> <mo>)</mo> </mrow> <mspace width="-0.166667em"/> <mo>:</mo> <mo>=</mo> <mspace width="-0.166667em"/> <mi>Q</mi> <mrow> <mo>(</mo> <mi>l</mi> <mo>,</mo> <mspace width="-0.166667em"/> <mi>Z</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> can be read off the plots. As can be seen, the bounds are much better for density (4); the values <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mi>Z</mi> <mo>)</mo> </mrow> <mo>≤</mo> <mn>1</mn> </mrow> </semantics></math> are consistent with all worldlines intersecting rather far from the laser-plasma interaction spacetime region. On the other hand, the large value of <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mi>Z</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for density (3) is an indication that worldlines intersect not far from the laser–plasma interaction spacetime region. Our computations lead to <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>20</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>4.62</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>3.08</mn> </mrow> </semantics></math> with density (3) and <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>20</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>0.49</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>Q</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>/</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>0.85</mn> </mrow> </semantics></math> with density (4).</p>
Full article ">Figure 5
<p>The initial electron densities (3), (4) of <a href="#mathematics-10-02622-f002" class="html-fig">Figure 2</a> (left and right, respectively) with <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>b</mi> </msub> <mo>=</mo> <mn>2</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>18</mn> </msup> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>, and below, the corresponding plots of <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>,</mo> <mi>σ</mi> <mo>,</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>∂</mo> <msup> <mi>u</mi> <mi>z</mi> </msup> <mo>/</mo> <mo>∂</mo> <mi>z</mi> </mrow> </semantics></math> during interaction with the pulse in <a href="#mathematics-10-02622-f001" class="html-fig">Figure 1</a> for a few sample values of <span class="html-italic">Z</span>. As can be seen, <span class="html-italic">J</span> remains positive at least for all <math display="inline"><semantics> <mrow> <mi>ξ</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>l</mi> <mo>]</mo> </mrow> </semantics></math> if the density is of type (4) (which grows as <math display="inline"><semantics> <msup> <mi>Z</mi> <mn>2</mn> </msup> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>∼</mo> <mn>0</mn> </mrow> </semantics></math>), whereas it becomes negative for <math display="inline"><semantics> <mrow> <mi>ξ</mi> <mo>∼</mo> <mn>6.5</mn> <msup> <mi>l</mi> <mo>′</mo> </msup> </mrow> </semantics></math> and small <span class="html-italic">Z</span> if the density is of type (3) (which grows as <span class="html-italic">Z</span> for <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>∼</mo> <mn>0</mn> </mrow> </semantics></math>). Correspondingly, the right-hand worldlines do not intersect, while the left-hand ones do (see the down <math display="inline"><semantics> <msub> <mi>z</mi> <mi>e</mi> </msub> </semantics></math>-graphs).</p>
Full article ">Figure 6
<p>The initial electron densities (1), (2) of <a href="#mathematics-10-02622-f002" class="html-fig">Figure 2</a> (first line; respectively left, right), and below, the corresponding plots of <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>,</mo> <mi>σ</mi> <mo>,</mo> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>∂</mo> <msup> <mi>u</mi> <mi>z</mi> </msup> <mo>/</mo> <mo>∂</mo> <mi>z</mi> </mrow> </semantics></math> vs. <math display="inline"><semantics> <mi>ξ</mi> </semantics></math> during interaction with the pulse in <a href="#mathematics-10-02622-f001" class="html-fig">Figure 1</a> for a few sample values of <span class="html-italic">Z</span>. As can be seen, the right <span class="html-italic">J</span> remains positive for <math display="inline"><semantics> <mrow> <mi>ξ</mi> <mo>&lt;</mo> <mi>l</mi> </mrow> </semantics></math> and all <span class="html-italic">Z</span>, while the left <span class="html-italic">J</span> becomes negative for very small <span class="html-italic">Z</span> and <math display="inline"><semantics> <mrow> <mi>ξ</mi> <mo>≲</mo> <mi>l</mi> </mrow> </semantics></math>; correspondingly, the right worldlines do not intersect, while the right ones do (see the down <math display="inline"><semantics> <msub> <mi>z</mi> <mi>e</mi> </msub> </semantics></math>-graphs).</p>
Full article ">Figure 7
<p><b>Down</b>: the initial electron density (3) of <a href="#mathematics-10-02622-f002" class="html-fig">Figure 2</a>. <b>Up</b>: The worldlines of <span class="html-italic">Z</span>-electrons interacting with the pulse in <a href="#mathematics-10-02622-f001" class="html-fig">Figure 1</a> for 200 equidistant values of <span class="html-italic">Z</span>; the support <math display="inline"><semantics> <mrow> <mn>0</mn> <mo>≤</mo> <mi>c</mi> <mi>t</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <mi>z</mi> <mo>≤</mo> <mi>l</mi> </mrow> </semantics></math> and the ’effective support’ <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mi>l</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> <mo>/</mo> <mn>2</mn> <mo>≤</mo> <mi>c</mi> <mi>t</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <mi>z</mi> <mo>≤</mo> <mrow> <mo>(</mo> <mi>l</mi> <mspace width="-0.166667em"/> <mo>+</mo> <mspace width="-0.166667em"/> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> of the pulse are pink and red;, respectively, while the spacetime region of the pure-ion layer is yellow. Horizontal arrows pinpoint where particular subsets of worldlines first intersect.</p>
Full article ">Figure 8
<p><b>Down</b>: The initial electron density (4) of <a href="#mathematics-10-02622-f002" class="html-fig">Figure 2</a>. <b>Up</b>: The worldlines of <span class="html-italic">Z</span>-electrons interacting with the pulse in <a href="#mathematics-10-02622-f001" class="html-fig">Figure 1</a> for 200 equidistant values of <span class="html-italic">Z</span>; the support <math display="inline"><semantics> <mrow> <mn>0</mn> <mo>≤</mo> <mi>c</mi> <mi>t</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <mi>z</mi> <mo>≤</mo> <mi>l</mi> </mrow> </semantics></math> and the ’effective support’ <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mi>l</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> <mo>/</mo> <mn>2</mn> <mo>≤</mo> <mi>c</mi> <mi>t</mi> <mspace width="-0.166667em"/> <mo>−</mo> <mspace width="-0.166667em"/> <mi>z</mi> <mo>≤</mo> <mrow> <mo>(</mo> <mi>l</mi> <mspace width="-0.166667em"/> <mo>+</mo> <mspace width="-0.166667em"/> <msup> <mi>l</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> of the pulse are pink and red, respectively, while the spacetime region of the pure-ion layer is yellow. Horizontal arrows pinpoint where particular subsets of worldlines first intersect; as can be seen, small <span class="html-italic">Z</span> worldlines first intersect quite farther from the laser–plasma interaction spacetime region (shown in pink) than in the linear homogenous case (3).</p>
Full article ">
21 pages, 80963 KiB  
Article
Tailoring of Inverse Energy Flow Profiles with Vector Lissajous Beams
by Svetlana N. Khonina, Alexey P. Porfirev, Andrey V. Ustinov, Mikhail S. Kirilenko and Nikolay L. Kazanskiy
Photonics 2022, 9(2), 121; https://doi.org/10.3390/photonics9020121 - 20 Feb 2022
Cited by 6 | Viewed by 2367
Abstract
In recent years, structured laser beams for shaping inverse energy flow regions: regions with a direction of energy flow opposite to the propagation direction of a laser beam, have been actively studied. Unfortunately, many structured laser beams generate inverse energy flow regions with [...] Read more.
In recent years, structured laser beams for shaping inverse energy flow regions: regions with a direction of energy flow opposite to the propagation direction of a laser beam, have been actively studied. Unfortunately, many structured laser beams generate inverse energy flow regions with dimensions of the order of the wavelength. Moreover, there are significant limitations to the location of these regions. Here, we investigate the possibility of controlling inverse energy flow distributions by using the generalization of well-known cylindrical vector beams with special polarization symmetry—vector Lissajous beams (VLBs)—defined by two polarization orders (p, q). We derive the conditions for the indices (p, q) in order, not only to shape separate isolated regions with a reverse energy flow, but also regions that are infinitely extended along a certain direction in the focal plane. In addition, we show that the maximum intensity curves of the studied VLBs are useful for predicting the properties of focused beams. Full article
(This article belongs to the Special Issue Polarized Light and Optical Systems)
19 pages, 7280 KiB  
Article
Thermophysical Characterization of Paraffin Wax Based on Mass-Accommodation Methods Applied to a Cylindrical Thermal Energy-Storage Unit
by Valter Silva-Nava, Ernesto M. Hernández-Cooper, Jesús Enrique Chong-Quero and José A. Otero
Molecules 2022, 27(4), 1189; https://doi.org/10.3390/molecules27041189 - 10 Feb 2022
Viewed by 2023
Abstract
Two mass-accommodation methods are proposed to describe the melting of paraffin wax used as a phase-change material in a centrally heated annular region. The two methods are presented as models where volume changes produced during the phase transition are incorporated through total mass [...] Read more.
Two mass-accommodation methods are proposed to describe the melting of paraffin wax used as a phase-change material in a centrally heated annular region. The two methods are presented as models where volume changes produced during the phase transition are incorporated through total mass conservation. The mass of the phase-change material is imposed as a constant, which brings an additional equation of motion. Volume changes in a cylindrical unit are pictured in two different ways. On the one hand, volume changes in the radial direction are proposed through an equation of motion where the outer radius of the cylindrical unit is promoted as a dynamical variable of motion. On the other hand, volume changes along the axial symmetry axis of the cylindrical unit are proposed through an equation of motion, where the excess volume of liquid constitutes the dynamical variable. The energy–mass balance at the liquid–solid interface is obtained according to each method of conceiving volume changes. The resulting energy–mass balance at the interface constitutes an equation of motion for the radius of the region delimited by the liquid–solid interface. Subtle differences are found between the equations of motion for the interface. The differences are consistent with mass conservation and local mass balance at the interface. Stationary states for volume changes and the radius of the region delimited by the liquid–solid interface are obtained for each mass-accommodation method. We show that the relationship between these steady states is proportional to the relationship between liquid and solid densities when the system is close to the high melting regime. Experimental tests are performed in a vertical annular region occupied by a paraffin wax. The boundary conditions used in the experimental tests produce a thin liquid layer during a melting process. The experimental results are used to characterize the phase-change material through the proposed models in this work. Finally, the thermodynamic properties of the paraffin wax are estimated by minimizing the quadratic error between the temperature readings within the phase-change material and the temperature field predicted by the proposed model. Full article
(This article belongs to the Special Issue Phase Change Materials 2.0)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of liquid volume growth in the axial direction. The height of the column that represents the excess liquid at any time <span class="html-italic">t</span> is <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>z</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>. The volume of this liquid at some time <span class="html-italic">t</span> is <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>V</mi> <mo>ℓ</mo> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>=</mo> <mfenced separators="" open="(" close=")"> <mover> <mi>r</mi> <mo>¯</mo> </mover> <msup> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mn>2</mn> </msup> <mo>−</mo> <msubsup> <mi>r</mi> <mn>0</mn> <mn>2</mn> </msubsup> </mfenced> <mrow> <mo>(</mo> <mi>L</mi> <mo>+</mo> <mo>Δ</mo> <mi>z</mi> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> and represents the liquid that will scatter throughout the top surface of the cylinder or the volume of liquid that must be removed from the cylindrical unit.</p>
Full article ">Figure 2
<p>Logarithmic relation between the two mass-accommodation methods previously discussed <math display="inline"><semantics> <mrow> <mo form="prefix">ln</mo> <mfenced separators="" open="(" close=")"> <msup> <mover> <mi>r</mi> <mo>¯</mo> </mover> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msup> <mo>/</mo> <msup> <mover> <mi>r</mi> <mo>¯</mo> </mover> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msup> </mfenced> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo form="prefix">ln</mo> <mfenced separators="" open="(" close=")"> <msup> <mi>R</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msup> <mo>/</mo> <msup> <mi>R</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msup> </mfenced> </mrow> </semantics></math>, obtained from the exact steady-state solutions given through Equations (<a href="#FD21-molecules-27-01189" class="html-disp-formula">21</a>) and (<a href="#FD22-molecules-27-01189" class="html-disp-formula">22</a>), and the asymptotic time values estimated through the FDM. Asterisk and cross symbols are used to represent the relation obtained through the exact steady-state values. Empty circles and squares represent the relation between the asymptotic time limits according to the numerical solutions for each mass-accommodation method. The dashed line corresponds to the predicted relation for high melting fractions according to Equation (<a href="#FD28-molecules-27-01189" class="html-disp-formula">28</a>).</p>
Full article ">Figure 3
<p>(<b>a</b>) Experimental setup with the cylindrical unit and thermocouple array. (<b>b</b>) Schematic representation of the experimental setup with the following components: 1. Lauda Thermostatic Bath, 2. Thermal energy-storage unit, 3. SCXI-1000 National Instruments module for thermocouple signal conditioning, 4. Laptop for data processing, A. 1 K-type thermocouple for heat bath temperature sensing, B. 1 K-type thermocouple for HTF inlet temperature sensing, C. 22 K-type thermocouple array for temperature sensing at the copper–PCM interface, aluminium–PCM interface and PCM temperature field estimation and D. 1 K-type thermocouple for HTF outlet temperature sensing. E. 1 K-type thermocouple for ambient temperature sensing.</p>
Full article ">Figure 4
<p>Cylindrical unit with thermocouple array for temperature sensing.</p>
Full article ">Figure 5
<p>Data acquisition and processing system. (<b>a</b>) Data acquisition system. (<b>b</b>) Laptop for temperature data processing.</p>
Full article ">Figure 6
<p>Nonhomogeneous isothermal boundary conditions. (<b>a</b>) Symbols represent time-dependent temperature values registered by the thermocouple located at the copper–PCM interface and the solid line corresponds to the best fit with the highest correlation. (<b>b</b>) Time-dependent temperature readings at the aluminium–PCM interface. Symbols represent the temperature values registered by the thermocouples located at <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0.108</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> and the solid line is the best fit with the highest correlation.</p>
Full article ">Figure 7
<p>Time evolution of the temperature at each thermocouple radial position according to the the experimental and numerical results. Experimental values of the average temperature at each radial coordinate (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>4.1</mn> <mspace width="0.166667em"/> <mi>cm</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>5.8</mn> <mspace width="0.166667em"/> <mi>cm</mi> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>7.5</mn> <mspace width="0.166667em"/> <mi>cm</mi> </mrow> </semantics></math> and (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mn>4</mn> </msub> <mo>=</mo> <mn>9.2</mn> <mspace width="0.166667em"/> <mi>cm</mi> </mrow> </semantics></math>, respectively are shown in red circles and the result obtained through the FDM is shown in solid lines. The numerical result was obtained through the solution of the model described by Equations (<a href="#FD9-molecules-27-01189" class="html-disp-formula">9</a>), (<a href="#FD14-molecules-27-01189" class="html-disp-formula">14</a>) and (<a href="#FD15-molecules-27-01189" class="html-disp-formula">15</a>) and through the set of thermodynamic variables <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mo>ℓ</mo> </msub> <mo>,</mo> <mspace width="0.166667em"/> <msub> <mi>k</mi> <mi>s</mi> </msub> <mo>,</mo> <mspace width="0.166667em"/> <msub> <mi>C</mi> <mo>ℓ</mo> </msub> <mspace width="0.166667em"/> <mspace width="0.166667em"/> <mi>and</mi> <mspace width="0.166667em"/> <mspace width="0.166667em"/> <msub> <mi>C</mi> <mi>s</mi> </msub> </mrow> </semantics></math> with the lowest quadratic error shown by Equation (<a href="#FD40-molecules-27-01189" class="html-disp-formula">40</a>).</p>
Full article ">Figure 8
<p>(<b>a</b>) Numerical solution to the liquid–solid interface motion. The results illustrate the relation between the liquid’s thickness and the outer radius of the thermal unit, according to the numerical solution of the proposed model with the set of thermodynamic variables shown in <a href="#molecules-27-01189-t002" class="html-table">Table 2</a>. (<b>b</b>) Time evolution of <math display="inline"><semantics> <msub> <mi>f</mi> <mi>s</mi> </msub> </semantics></math> obtained from the numerical solutions to the proposed model and with the set of parameters shown in <a href="#molecules-27-01189-t002" class="html-table">Table 2</a>.</p>
Full article ">
16 pages, 4150 KiB  
Article
Analysis of the Scattering from a Two Stacked Thin Resistive Disks Resonator by Means of the Helmholtz–Galerkin Regularizing Technique
by Mario Lucido
Appl. Sci. 2021, 11(17), 8173; https://doi.org/10.3390/app11178173 - 3 Sep 2021
Cited by 5 | Viewed by 1344
Abstract
In this paper, the scattering of a plane wave from a lossy Fabry–Perót resonator, realized with two equiaxial thin resistive disks with the same radius, is analyzed by means of the generalization of the Helmholtz–Galerkin regularizing technique recently developed by the author. The [...] Read more.
In this paper, the scattering of a plane wave from a lossy Fabry–Perót resonator, realized with two equiaxial thin resistive disks with the same radius, is analyzed by means of the generalization of the Helmholtz–Galerkin regularizing technique recently developed by the author. The disks are modelled as 2-D planar surfaces described in terms of generalized boundary conditions. Taking advantage of the revolution symmetry, the problem is equivalently formulated as a set of independent systems of 1-D equations in the vector Hankel transform domain for the cylindrical harmonics of the effective surface current densities. The Helmholtz decomposition of the unknowns, combined with a suitable choice of the expansion functions in a Galerkin scheme, lead to a fast-converging Fredholm second-kind matrix operator equation. Moreover, an analytical technique specifically devised to efficiently evaluate the integrals of the coefficient matrix is adopted. As shown in the numerical results section, near-field and far-field parameters are accurately and efficiently reconstructed even at the resonance frequencies of the natural modes, which are searched for the peaks of the total scattering cross-section and the absorption cross-section. Moreover, the proposed method drastically outperforms the general-purpose commercial software CST Microwave Studio in terms of both CPU time and memory occupation. Full article
Show Figures

Figure 1

Figure 1
<p>Geometry of the problem.</p>
Full article ">Figure 2
<p>Relative computation error for a two stacked thin resistive disks resonator with different values of the radius of the disks (<math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mrow> <mi>λ</mi> <mo>/</mo> <mn>2</mn> </mrow> <mo>,</mo> <mi>λ</mi> <mo>,</mo> <mn>2</mn> <mi>λ</mi> </mrow> </semantics></math>) and of the distance between the disks (<math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mo>∞</mo> </mrow> </semantics></math>). <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>100</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>200</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <mo> </mo> <msub> <mrow> <munder accentunder="true"> <mi>E</mi> <mo stretchy="true">_</mo> </munder> </mrow> <mn>0</mn> </msub> </mrow> <mo>|</mo> </mrow> <mo>=</mo> <mn>1</mn> <mo> </mo> <mrow> <mi mathvariant="normal">V</mi> <mo>/</mo> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>30</mn> <mo>°</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ϕ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>°</mo> </mrow> </semantics></math>, and TE incidence: (<b>a</b>) <math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mo>∞</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Components of the effective electric current densities on a two stacked thin resistive disks resonator for different values of the radius of the disks (<math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mrow> <mi>λ</mi> <mo>/</mo> <mn>2</mn> </mrow> <mo>,</mo> <mi>λ</mi> <mo>,</mo> <mn>2</mn> <mi>λ</mi> </mrow> </semantics></math>). <math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>100</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>200</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <mo> </mo> <msub> <mrow> <munder accentunder="true"> <mi>E</mi> <mo stretchy="true">_</mo> </munder> </mrow> <mn>0</mn> </msub> </mrow> <mo>|</mo> </mrow> <mo>=</mo> <mn>1</mn> <mo> </mo> <mrow> <mi mathvariant="normal">V</mi> <mo>/</mo> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ϕ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>°</mo> </mrow> </semantics></math>, and TE incidence: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mrow> <mi>λ</mi> <mo>/</mo> <mn>2</mn> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <msub> <mi>J</mi> <mrow> <mi>i</mi> <mo>,</mo> <mi>ρ</mi> </mrow> </msub> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mrow> <mi>λ</mi> <mo>/</mo> <mn>2</mn> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <msub> <mi>J</mi> <mrow> <mi>i</mi> <mo>,</mo> <mi>ϕ</mi> </mrow> </msub> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mi>λ</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <msub> <mi>J</mi> <mrow> <mi>i</mi> <mo>,</mo> <mi>ρ</mi> </mrow> </msub> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mi>λ</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <msub> <mi>J</mi> <mrow> <mi>i</mi> <mo>,</mo> <mi>ϕ</mi> </mrow> </msub> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>; (<b>e</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>2</mn> <mi>λ</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <msub> <mi>J</mi> <mrow> <mi>i</mi> <mo>,</mo> <mi>ρ</mi> </mrow> </msub> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>; (<b>f</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>2</mn> <mi>λ</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <msub> <mi>J</mi> <mrow> <mi>i</mi> <mo>,</mo> <mi>ϕ</mi> </mrow> </msub> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Components of the effective electric current densities on the upper disk (disk 1) of a two stacked thin resistive disks resonator for different values of the distance between the disks (<math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mo>∞</mo> </mrow> </semantics></math>). <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mi>λ</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>100</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>200</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <mo> </mo> <msub> <mrow> <munder accentunder="true"> <mi>E</mi> <mo stretchy="true">_</mo> </munder> </mrow> <mn>0</mn> </msub> </mrow> <mo>|</mo> </mrow> <mo>=</mo> <mn>1</mn> <mo> </mo> <mrow> <mi mathvariant="normal">V</mi> <mo>/</mo> <mi mathvariant="normal">m</mi> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>30</mn> <mo>°</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ϕ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>°</mo> </mrow> </semantics></math>, and TE incidence: (<b>a</b>) <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <msub> <mi>J</mi> <mrow> <mn>1</mn> <mo>,</mo> <mi>ρ</mi> </mrow> </msub> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mrow> <msub> <mi>J</mi> <mrow> <mn>1</mn> <mo>,</mo> <mi>ϕ</mi> </mrow> </msub> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>BRCS of a two stacked thin resistive disks resonator for different values of the radius of the disks (<math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mrow> <mi>λ</mi> <mo>/</mo> <mn>2</mn> </mrow> <mo>,</mo> <mi>λ</mi> <mo>,</mo> <mn>2</mn> <mi>λ</mi> </mrow> </semantics></math>). <math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>100</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>200</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>30</mn> <mo>°</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ϕ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>°</mo> </mrow> </semantics></math>, and TE incidence: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mrow> <mi>λ</mi> <mo>/</mo> <mn>2</mn> </mrow> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mi>λ</mi> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>2</mn> <mi>λ</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Normalized TSCS and ACS of a two stacked thin resistive disks resonator for varying values of the normalized frequency (<math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> </mrow> </semantics></math>) when a plane wave orthogonally impinges onto the structure. <math display="inline"><semantics> <mrow> <mrow> <mi>d</mi> <mo>/</mo> <mi>a</mi> </mrow> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>R</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> <mo> </mo> <mi mathvariant="sans-serif">Ω</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Near E-field behavior in the Cartesian coordinate planes <span class="html-italic">xz</span> and <span class="html-italic">xy</span> of a two stacked thin resistive disks resonator at three resonance frequencies (<math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>9.662491</mn> <mo>,</mo> <mn>10.643395</mn> <mo>,</mo> <mn>11.998105</mn> </mrow> </semantics></math>), when a plane wave orthogonally impinges onto the structure with <math display="inline"><semantics> <mrow> <msub> <mrow> <munder accentunder="true"> <mi>E</mi> <mo stretchy="true">_</mo> </munder> </mrow> <mn>0</mn> </msub> <mo>=</mo> <msub> <mi>E</mi> <mn>0</mn> </msub> <mover accent="true"> <mi>y</mi> <mo>^</mo> </mover> </mrow> </semantics></math>: (<b>a</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>9.662491</mn> </mrow> </semantics></math>; (<b>b</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>9.662491</mn> </mrow> </semantics></math>; (<b>c</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>10.643395</mn> </mrow> </semantics></math>; (<b>d</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>10.643395</mn> </mrow> </semantics></math>; (<b>e</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>11.998105</mn> </mrow> </semantics></math>; (<b>f</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>11.998105</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7 Cont.
<p>Near E-field behavior in the Cartesian coordinate planes <span class="html-italic">xz</span> and <span class="html-italic">xy</span> of a two stacked thin resistive disks resonator at three resonance frequencies (<math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>9.662491</mn> <mo>,</mo> <mn>10.643395</mn> <mo>,</mo> <mn>11.998105</mn> </mrow> </semantics></math>), when a plane wave orthogonally impinges onto the structure with <math display="inline"><semantics> <mrow> <msub> <mrow> <munder accentunder="true"> <mi>E</mi> <mo stretchy="true">_</mo> </munder> </mrow> <mn>0</mn> </msub> <mo>=</mo> <msub> <mi>E</mi> <mn>0</mn> </msub> <mover accent="true"> <mi>y</mi> <mo>^</mo> </mover> </mrow> </semantics></math>: (<b>a</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>9.662491</mn> </mrow> </semantics></math>; (<b>b</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>9.662491</mn> </mrow> </semantics></math>; (<b>c</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>10.643395</mn> </mrow> </semantics></math>; (<b>d</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>10.643395</mn> </mrow> </semantics></math>; (<b>e</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>11.998105</mn> </mrow> </semantics></math>; (<b>f</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>11.998105</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Near E-field behavior in the Cartesian coordinate planes <span class="html-italic">xz</span> and <span class="html-italic">xy</span> of a two stacked thin resistive disks resonator at three resonance frequencies (<math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>15.862148</mn> <mo>,</mo> <mn>16.503645</mn> <mo>,</mo> <mn>17.532887</mn> </mrow> </semantics></math>), when a plane wave orthogonally impinges onto the structure with <math display="inline"><semantics> <mrow> <msub> <mrow> <munder accentunder="true"> <mi>E</mi> <mo stretchy="true">_</mo> </munder> </mrow> <mn>0</mn> </msub> <mo>=</mo> <msub> <mi>E</mi> <mn>0</mn> </msub> <mover accent="true"> <mi>y</mi> <mo>^</mo> </mover> </mrow> </semantics></math>: (<b>a</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>15.862148</mn> </mrow> </semantics></math>; (<b>b</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>15.862148</mn> </mrow> </semantics></math>; (<b>c</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>16.503645</mn> </mrow> </semantics></math>; (<b>d</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>16.503645</mn> </mrow> </semantics></math>; (<b>e</b>) Near E-field in the <span class="html-italic">xz</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>17.532887</mn> </mrow> </semantics></math>; (<b>f</b>) Near E-field in the <span class="html-italic">xy</span> plane, <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>0</mn> </msub> <mi>a</mi> <mo>=</mo> <mn>17.532887</mn> </mrow> </semantics></math>.</p>
Full article ">
15 pages, 308 KiB  
Communication
Generalized Rest Mass and Dirac’s Monopole in 5D Theory and Cosmology
by Boris G. Aliyev
Universe 2021, 7(8), 295; https://doi.org/10.3390/universe7080295 - 11 Aug 2021
Cited by 1 | Viewed by 1546
Abstract
It is shown that the 5D geodetic equations and 5D Ricci identities give us a way to create a new viewpoint on some problems of modern physics, astrophysics, and cosmology. Specifically, the application of the 5D geodetic equations in (4+1) and (3+1+1) splintered [...] Read more.
It is shown that the 5D geodetic equations and 5D Ricci identities give us a way to create a new viewpoint on some problems of modern physics, astrophysics, and cosmology. Specifically, the application of the 5D geodetic equations in (4+1) and (3+1+1) splintered forms obtained with the help of the monad and dyad methods made it possible to introduce a new, effective generalized concept of the rest mass of the elementary particle. The latter leads one to novel connections between the general relativity and quantum field theories, and all that, including the (4+1) splitting of the 5D Ricci identities, brings about a better understanding of the magnetic monopole problem and the vital difference in the origins of the Maxwell equations and gives rise to surprising connections between them. The obtained results also provide new insight into the mechanism of the 4D universe’s expansion and its following acceleration. Full article
20 pages, 10568 KiB  
Article
A Novel Single-Loop Mechanism and the Associated Cylindrical Deployable Mechanisms
by Long Huang, Bei Liu, Lairong Yin and Jinhang Wang
Symmetry 2021, 13(7), 1255; https://doi.org/10.3390/sym13071255 - 13 Jul 2021
Cited by 5 | Viewed by 2474
Abstract
This paper presents a new type of 2-DOF single-loop mechanism inspired by the Sarrus mechanism, and it utilizes this mechanism to construct 2-DOF cylindrical deployable mechanisms. First, the motion pattern of the single-loop mechanism is analyzed utilizing screw theory. According to the structural [...] Read more.
This paper presents a new type of 2-DOF single-loop mechanism inspired by the Sarrus mechanism, and it utilizes this mechanism to construct 2-DOF cylindrical deployable mechanisms. First, the motion pattern of the single-loop mechanism is analyzed utilizing screw theory. According to the structural symmetries, the cylindrical deployable mechanisms are constructed through the linear pattern combination and circular pattern combination of the single-loop mechanisms. After the geometrical analysis and interference condition analysis, the axial, circumferential and area magnification ratios are defined and, furthermore, applied to the parameter optimization of the deployable mechanisms, forming an example surface. Finally, a simplified 1-DOF single-loop mechanism is derived from the proposed 2-DOF mechanism, which is used to construct 1-DOF cylindrical deployable mechanisms with singular free workspaces. Full article
Show Figures

Figure 1

Figure 1
<p>The schematic diagram of the Sarrus mechanism.</p>
Full article ">Figure 2
<p>The schematic diagram of the 8R mechanism.</p>
Full article ">Figure 3
<p>Structural design of the 8R mechanism.</p>
Full article ">Figure 4
<p>Coordinate frames of the 8R mechanism and the equivalent mechanism.</p>
Full article ">Figure 5
<p>Demonstration of variables.</p>
Full article ">Figure 6
<p>Diagram of the triangles.</p>
Full article ">Figure 7
<p>Singular configurations: (<b>a</b>) singular case I; (<b>b</b>) singular case II; and (<b>c</b>) the fully deployed configuration.</p>
Full article ">Figure 8
<p>The cylindrical DM composed of four DUs.</p>
Full article ">Figure 9
<p>The topological constraint graph of the cylindrical DM.</p>
Full article ">Figure 10
<p>The cylindrical circumscribed rectangle.</p>
Full article ">Figure 11
<p>Interference case I.</p>
Full article ">Figure 12
<p>Interference case II.</p>
Full article ">Figure 13
<p>The relationship of magnification ratios.</p>
Full article ">Figure 14
<p>Influences of m and n on the area magnification ratio.</p>
Full article ">Figure 15
<p>Simulation of the deployment. (<b>a</b>) Fully folded configuration. (<b>b</b>) Partially deployed configuration. (<b>c</b>) Fully deployed configuration.</p>
Full article ">Figure 16
<p>The deployment of the prototype. (<b>a</b>) Fully folded configuration. (<b>b</b>) Partially deployed configuration. (<b>c</b>) Fully deployed configuration.</p>
Full article ">Figure 17
<p>The 6R DU.</p>
Full article ">Figure 18
<p>The 1-DOF DM composed of four 6R DUs.</p>
Full article ">
31 pages, 7432 KiB  
Article
Efficient Computation of Heat Distribution of Processed Materials under Laser Irradiation
by Jianxin Zhu and Wencheng Lin
Mathematics 2021, 9(12), 1368; https://doi.org/10.3390/math9121368 - 12 Jun 2021
Cited by 4 | Viewed by 1709
Abstract
In this paper, a solution is provided to solve the heat conduction equation in the three-dimensional cylinder region, where the laser intensity of the material irradiation surface is expressed as a Gaussian distribution. Based on the symmetry of heat distribution, firstly, the form [...] Read more.
In this paper, a solution is provided to solve the heat conduction equation in the three-dimensional cylinder region, where the laser intensity of the material irradiation surface is expressed as a Gaussian distribution. Based on the symmetry of heat distribution, firstly, the form of the heat equation in the common rectangular coordinate system is changed to another form in the two-dimensional cylindrical coordinate system. Secondly, the ADI with the backward Euler method and with Crank–Nicolson method are established to discretize the model in the cylindrical coordinate system, after which the simulation results are obtained, where the first kind of boundary value condition is used to verify the accuracy of these two algorithms. Then, the above two methods are used to solve the model with the third kind of boundary value condition. Finally, the comparison is performed with the results obtained by the MATLAB’s PDETOOL, which shows that the solution is more feasible and efficient. Full article
Show Figures

Figure 1

Figure 1
<p>Sketch of the domain of heat equation.</p>
Full article ">Figure 2
<p>Numerical solutions of difference equations <math display="inline"><semantics> <mrow> <mstyle mathvariant="bold"> <mfenced> <mrow> <mi mathvariant="bold-italic">h</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>64</mn> <mo>,</mo> <mi mathvariant="bold-italic">τ</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>2048</mn> </mrow> </mfenced> </mstyle> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Numerical solutions obtained by PDETOOL of MATLAB.</p>
Full article ">Figure 4
<p>Longitudinal section contrast.</p>
Full article ">Figure 5
<p>Cross-sectional contrast.</p>
Full article ">Figure 6
<p>Numerical solutions of difference equations <math display="inline"><semantics> <mrow> <mstyle mathvariant="bold"> <mfenced> <mrow> <mi mathvariant="bold-italic">h</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>64</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">τ</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>128</mn> </mrow> </mfenced> </mstyle> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Numerical solutions obtained by PDETOOL of MATLAB.</p>
Full article ">Figure 8
<p>Longitudinal section contrast.</p>
Full article ">Figure 9
<p>Cross-sectional contrast.</p>
Full article ">Figure 10
<p>Numerical solutions of difference equations <math display="inline"><semantics> <mrow> <mstyle mathvariant="bold"> <mfenced> <mrow> <mi mathvariant="bold-italic">h</mi> <mo>=</mo> <mn>0.001</mn> <mo>/</mo> <mn>100</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">τ</mi> <mo>=</mo> <mn>0.01</mn> <mo>/</mo> <mn>1000</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">H</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </mfenced> </mstyle> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Numerical solutions of difference equations <math display="inline"><semantics> <mrow> <mstyle mathvariant="bold"> <mfenced> <mrow> <mi mathvariant="bold-italic">h</mi> <mo>=</mo> <mn>0.001</mn> <mo>/</mo> <mn>100</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">τ</mi> <mo>=</mo> <mn>0.01</mn> <mo>/</mo> <mn>500</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">H</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </mfenced> </mstyle> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>Numerical solutions obtained by PDETOOL of MATLAB.</p>
Full article ">Figure 13
<p>Longitudinal sectional contrast.</p>
Full article ">Figure 14
<p>Cross section contrast.</p>
Full article ">Figure 15
<p>Numerical solutions of difference equations <math display="inline"><semantics> <mrow> <mstyle mathvariant="bold"> <mfenced> <mrow> <mi mathvariant="bold-italic">h</mi> <mo>=</mo> <mn>0.001</mn> <mo>/</mo> <mn>100</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">τ</mi> <mo>=</mo> <mn>0.01</mn> <mo>/</mo> <mn>500</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">H</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </mfenced> </mstyle> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>Numerical solutions obtained by PDETOOL of MATLAB.</p>
Full article ">Figure 17
<p>Numerical solutions of difference equations <math display="inline"><semantics> <mrow> <mstyle mathvariant="bold"> <mfenced> <mrow> <mi mathvariant="bold-italic">h</mi> <mo>=</mo> <mn>0.001</mn> <mo>/</mo> <mn>100</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">τ</mi> <mo>=</mo> <mn>0.01</mn> <mo>/</mo> <mn>500</mn> <mo>,</mo> <mo> </mo> <mi mathvariant="bold-italic">H</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </mfenced> </mstyle> </mrow> </semantics></math>.</p>
Full article ">
11 pages, 2162 KiB  
Article
The Influence of the Implant Macrogeometry on Insertion Torque, Removal Torque, and Periotest Implant Primary Stability: A Mechanical Simulation on High-Density Artificial Bone
by Margherita Tumedei, Morena Petrini, Davide Pietropaoli, Alessandro Cipollina, Castrenze La Torre, Maria Stella Di Carmine, Adriano Piattelli and Giovanna Iezzi
Symmetry 2021, 13(5), 776; https://doi.org/10.3390/sym13050776 - 30 Apr 2021
Cited by 6 | Viewed by 3508
Abstract
Background: The primary stability is a determinant clinical condition for the success of different dental implants macro-design in different bone density using a validated and repeatable in vitro technique employing solid rigid polyurethane blocks. Materials and Methods: Five implants 3.8 × 13 mm [...] Read more.
Background: The primary stability is a determinant clinical condition for the success of different dental implants macro-design in different bone density using a validated and repeatable in vitro technique employing solid rigid polyurethane blocks. Materials and Methods: Five implants 3.8 × 13 mm2 for each macro-design (i.e., IK—tapered; IC—cylindric; and IA—active blade shape) were positioned into 20- and 30- pounds per cubic foot (PCF) polyurethane blocks. Bucco-lingual (BL) and mesial-distal (MD) implant stability quotient score (ISQ) was assessed by resonance frequency analysis while, insertion/removal torques were evaluated by dynamometric ratchet. Results: IC implants shown better primary stability in terms of ISQ compared to IA and IK in lower density block (20 PCF), while IK was superior to IA in higher density (30 PCF). IC shown higher removal torque in 30-PCF compared to IA and IC. Conclusions: The study effectiveness on polyurethane artificial bone with isotropic symmetry structure showed that the implants macro-design might represent a key factor on primary stability, in particular on low-density alveolar bone. Clinicians should consider patients features and implant geometry during low-density jaws rehabilitation. Further investigations are needed to generalize these findings. Full article
(This article belongs to the Special Issue Symmetry in Dentistry: From the Clinic to the Lab)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of the design of the study. PCF pounds per cubic foot block; IT insertion torque; RT removal torque; ISQ implant stability quotient score.</p>
Full article ">Figure 2
<p>Graphical representation of the IC, IK, and IA implants.</p>
Full article ">Figure 3
<p>Details of the insertion and removal torque measured by the dedicated dynamometric ratchet.</p>
Full article ">Figure 4
<p>Charts representing the BL and MD ISQ, calculated for the study groups (Kruskal–Wallis test, <span class="html-italic">p</span> &lt; 0.01). (<b>A</b>) Buccal-Lingual ISQ score. (<b>B</b>) Mesial-Distal ISQ Score.</p>
Full article ">Figure 5
<p>Charts representing the IT and RT, calculated for the study groups (Kruskal–Wallis test, <span class="html-italic">p</span> &lt; 0.01). (<b>A</b>) Insertion Torque of IA, IC, and IK implants. (<b>B</b>) Removal Torque of IA, IC, and IK implants.</p>
Full article ">
Back to TopTop