[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

Journals

Article Types

Countries / Regions

Search Results (18)

Search Parameters:
Keywords = Hopf’s invariant

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 823 KiB  
Article
Acoustic Drift: Generating Helicity and Transferring Energy
by Andrey Morgulis
Axioms 2024, 13(11), 767; https://doi.org/10.3390/axioms13110767 - 4 Nov 2024
Viewed by 684
Abstract
This article studies the general properties of the Stokes drift field. This name is commonly used for the correction added to the mean Eulerian velocity for describing the averaged transport of the material particles by the oscillating fluid flows. Stokes drift is widely [...] Read more.
This article studies the general properties of the Stokes drift field. This name is commonly used for the correction added to the mean Eulerian velocity for describing the averaged transport of the material particles by the oscillating fluid flows. Stokes drift is widely known mainly in connection with another feature of oscillating flows known as steady streaming, which has been and remains the focus of a multitude of studies. However, almost nothing is known about Stokes drift in general, e.g., about its energy or helicity (Hopf’s invariant). We address these quantities for acoustic drift driven by simple sound waves with finite discrete Fourier spectra. The results discover that the mean drift energy is partly localized on a certain resonant set, which we have described explicitly. Moreover, the mean drift helicity turns out to be completely localized on the same set. We also present several simple examples to discover the effect of the power spectrum and positioning of the spectral atoms. It is revealed that tuning them can drastically change both resonant and non-resonant energies, zero the helicity, or even increase it unboundedly. Full article
(This article belongs to the Special Issue Fluid Dynamics: Mathematics and Numerical Experiment)
Show Figures

Figure 1

Figure 1
<p>This figure illustrates the drift energy that arises from configuration (<a href="#FD26-axioms-13-00767" class="html-disp-formula">26</a>)–(<a href="#FD29-axioms-13-00767" class="html-disp-formula">29</a>) when the number of atoms tends to <span class="html-italic">∞</span> along a sequence of the doubled primes. The distribution of power spectrum over the atoms is the normal periodic one (right panel) or the uniform one (left panel). The graphs depict the total mean drift energy and its resonant or non-resonant fractions vs. the geometric parameter <math display="inline"><semantics> <mi>β</mi> </semantics></math>. Recall that the drift includes no uniform component for the uniform power spectrum distribution (see Remark 1).</p>
Full article ">Figure 2
<p>This figure displays samples of 20 level sets of the stream function (<a href="#FD59-axioms-13-00767" class="html-disp-formula">59</a>) for the drift field emergent from 22 atoms configuration defined by equalities (<a href="#FD26-axioms-13-00767" class="html-disp-formula">26</a>)–(<a href="#FD29-axioms-13-00767" class="html-disp-formula">29</a>) for equal amplitudes, <math display="inline"><semantics> <msub> <mi>φ</mi> <mi>j</mi> </msub> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>10</mn> </mrow> </semantics></math>. Upper-left, upper-right, lower-left, and lower-right frames display the samples taken for <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mn>1.69</mn> <mo>,</mo> <mn>3.38</mn> <mo>,</mo> <mn>5.08</mn> </mrow> </semantics></math>, correspondingly. So, the sampling interval is approximately eight times smaller then the period. The next semi-period demonstrates the reverse changes.</p>
Full article ">
16 pages, 4720 KiB  
Article
Dynamics of a New Four-Thirds-Degree Sub-Quadratic Lorenz-like System
by Guiyao Ke, Jun Pan, Feiyu Hu and Haijun Wang
Axioms 2024, 13(9), 625; https://doi.org/10.3390/axioms13090625 - 12 Sep 2024
Cited by 1 | Viewed by 578
Abstract
Aiming to explore the subtle connection between the number of nonlinear terms in Lorenz-like systems and hidden attractors, this paper introduces a new simple sub-quadratic four-thirds-degree Lorenz-like system, where x˙=a(yx), [...] Read more.
Aiming to explore the subtle connection between the number of nonlinear terms in Lorenz-like systems and hidden attractors, this paper introduces a new simple sub-quadratic four-thirds-degree Lorenz-like system, where x˙=a(yx), y˙=cxx3z, z˙=bz+x3y, and uncovers the following property of these systems: decreasing the powers of the nonlinear terms in a quadratic Lorenz-like system where x˙=a(yx), y˙=cxxz, z˙=bz+xy, may narrow, or even eliminate the range of the parameter c for hidden attractors, but enlarge it for self-excited attractors. By combining numerical simulation, stability and bifurcation theory, most of the important dynamics of the Lorenz system family are revealed, including self-excited Lorenz-like attractors, Hopf bifurcation and generic pitchfork bifurcation at the origin, singularly degenerate heteroclinic cycles, degenerate pitchfork bifurcation at non-isolated equilibria, invariant algebraic surface, heteroclinic orbits and so on. The obtained results may verify the generalization of the second part of the celebrated Hilbert’s sixteenth problem to some degree, showing that the number and mutual disposition of attractors and repellers may depend on the degree of chaotic multidimensional dynamical systems. Full article
(This article belongs to the Section Mathematical Analysis)
Show Figures

Figure 1

Figure 1
<p>For <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>4</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>5</mn> <mo>]</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>1</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>1</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>1</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.13</mn> <mo>,</mo> <mn>1.3</mn> <mo>,</mo> <mn>1.6</mn> <mo>)</mo> </mrow> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>7</mn> </mrow> </msup> </mrow> </semantics></math>; (<b>a</b>–<b>c</b>) bifurcation diagrams; (<b>d</b>) Lyapunov exponents versus <span class="html-italic">c</span> of system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>). In contrast to system (<a href="#FD2-axioms-13-00625" class="html-disp-formula">2</a>) [<a href="#B22-axioms-13-00625" class="html-bibr">22</a>] (Figure 3, p. 363), these figures suggest that the solutions for the system in (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) display stable equilibria and period orbits, rather than the self-excited and hidden attractors shown in the system in (<a href="#FD2-axioms-13-00625" class="html-disp-formula">2</a>).</p>
Full article ">Figure 2
<p>For <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>3</mn> <mo>,</mo> <mn>1.5</mn> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>∈</mo> <mo>[</mo> <mn>0.1</mn> <mo>,</mo> <mn>599.1</mn> <mo>]</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1.314</mn> <mo>,</mo> <mn>2.236</mn> <mo>,</mo> <mn>4.669</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>a</b>–<b>c</b>) bifurcation diagrams; (<b>d</b>) Lyapunov exponents versus <span class="html-italic">c</span> of system (<a href="#FD2-axioms-13-00625" class="html-disp-formula">2</a>). The subfigures (<b>a</b>–<b>c</b>) are consistent with the subfigure (<b>d</b>), showing that system (<a href="#FD2-axioms-13-00625" class="html-disp-formula">2</a>) mainly experiences periodic behaviors.</p>
Full article ">Figure 3
<p>For <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>3</mn> <mo>,</mo> <mn>1.5</mn> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>∈</mo> <mo>[</mo> <mn>0.1</mn> <mo>,</mo> <mn>599.1</mn> <mo>]</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1.314</mn> <mo>,</mo> <mn>2.236</mn> <mo>,</mo> <mn>4.669</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>a</b>–<b>c</b>) bifurcation diagrams; (<b>d</b>) Lyapunov exponents versus <span class="html-italic">c</span> of system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>). In contrast with <a href="#axioms-13-00625-f002" class="html-fig">Figure 2</a>, the four subfigures show that system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) mainly behaves in a similar way to self-excited attractors, verifying the introduced property, i.e., a decrease in powers of nonlinear terms of the quadratic Lorenz-like system (<a href="#FD2-axioms-13-00625" class="html-disp-formula">2</a>) may narrow or even eliminate the range of the parameter <span class="html-italic">c</span> for hidden attractors, but enlarge it for self-excited attractors.</p>
Full article ">Figure 4
<p>Phase portraits of system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) for <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>c</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>3</mn> <mo>,</mo> <mn>100</mn> <mo>,</mo> <mn>1.5</mn> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1.314</mn> <mo>,</mo> <mn>2.236</mn> <mo>,</mo> <mn>4.669</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> illustrating the existence of two-scroll self-excited attractor suggested in <a href="#axioms-13-00625-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 4 Cont.
<p>Phase portraits of system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) for <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>c</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>3</mn> <mo>,</mo> <mn>100</mn> <mo>,</mo> <mn>1.5</mn> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1.314</mn> <mo>,</mo> <mn>2.236</mn> <mo>,</mo> <mn>4.669</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> illustrating the existence of two-scroll self-excited attractor suggested in <a href="#axioms-13-00625-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>Poincaré cross-sections of system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) for <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>c</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>3</mn> <mo>,</mo> <mn>100</mn> <mo>,</mo> <mn>1.5</mn> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>2</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1.314</mn> <mo>,</mo> <mn>2.236</mn> <mo>,</mo> <mn>4.669</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> showing the geometrical structure of the Lorenz-like attractor depicted in <a href="#axioms-13-00625-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 6
<p>Phase portraits of system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) for <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>c</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>1</mn> <mo>,</mo> <mn>36</mn> <mo>)</mo> </mrow> </semantics></math>, (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>0.07</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>1</mn> <mo>,</mo> <mn>3</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>1</mn> <mo>,</mo> <mn>3</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>3</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mo>±</mo> <mn>1.3</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>8</mn> </mrow> </msup> <mo>,</mo> <mo>±</mo> <mn>1.3</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>7</mn> </mrow> </msup> <mo>,</mo> <mo>−</mo> <mn>1</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. Both figures imply that collapsing singularly degenerate heteroclinic cycles in system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) create limited cycles rather than strange attractors.</p>
Full article ">Figure 7
<p>Phase portraits of system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) for <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>c</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>4</mn> <mo>,</mo> <mn>1.8182</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math> and (<b>a</b>) <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>1</mn> <mo>,</mo> <mn>3</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>1</mn> <mo>,</mo> <mn>3</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>4</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mo>±</mo> <mn>0.13</mn> <mo>,</mo> <mo>±</mo> <mn>1.3</mn> <mo>,</mo> <mn>1.6</mn> <mo>)</mo> </mrow> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>7</mn> </mrow> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>3</mn> <mo>,</mo> <mn>4</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>3</mn> <mo>,</mo> <mn>4</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>4</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mo>±</mo> <mn>2.4</mn> <mo>,</mo> <mo>±</mo> <mn>2.41</mn> <mo>,</mo> <mn>3.3</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>5</mn> <mo>,</mo> <mn>6</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>5</mn> <mo>,</mo> <mn>6</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>5</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mo>±</mo> <mn>2.39</mn> <mo>,</mo> <mo>±</mo> <mn>2.4</mn> <mo>,</mo> <mn>3.32</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>3</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>3</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>4</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>2.4</mn> <mo>,</mo> <mn>2.41</mn> <mo>,</mo> <mn>3.3</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>5</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>5</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>5</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>2.39</mn> <mo>,</mo> <mn>2.4</mn> <mo>,</mo> <mn>3.32</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>4</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>4</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>4</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mo>−</mo> <mn>2.4</mn> <mo>,</mo> <mo>−</mo> <mn>2.41</mn> <mo>,</mo> <mn>3.3</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mn>6</mn> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mn>6</mn> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>5</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mo>−</mo> <mn>2.39</mn> <mo>,</mo> <mo>−</mo> <mn>2.4</mn> <mo>,</mo> <mn>3.32</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, showing at least five limit cycles for system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>) when Hopf bifurcation occurs at <math display="inline"><semantics> <msub> <mi>E</mi> <mo>±</mo> </msub> <mo>,</mo> </semantics></math> i.e., two around <math display="inline"><semantics> <msub> <mi>E</mi> <mo>+</mo> </msub> </semantics></math>, two around <math display="inline"><semantics> <msub> <mi>E</mi> <mo>−</mo> </msub> </semantics></math> and one around <math display="inline"><semantics> <msub> <mi>E</mi> <mo>±</mo> </msub> </semantics></math>.</p>
Full article ">Figure 8
<p>For <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>c</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>6</mn> <mo>,</mo> <mn>100</mn> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>8</mn> <mo>,</mo> <mn>9</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>1</mn> <mo>,</mo> <mn>3</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>y</mi> <mrow> <mn>0</mn> </mrow> <mrow> <mn>1</mn> <mo>,</mo> <mn>3</mn> </mrow> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mn>0</mn> </mrow> <mn>3</mn> </msubsup> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mo>±</mo> <mn>0.13</mn> <mo>,</mo> <mo>±</mo> <mn>1.3</mn> <mo>,</mo> <mn>1.6</mn> <mo>)</mo> </mrow> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>7</mn> </mrow> </msup> </mrow> </semantics></math>, phase portraits of system (<a href="#FD1-axioms-13-00625" class="html-disp-formula">1</a>), verifying the existence of a pair of heteroclinic orbits to unstable <math display="inline"><semantics> <msub> <mi>E</mi> <mn>0</mn> </msub> </semantics></math> and stable <math display="inline"><semantics> <msub> <mi>E</mi> <mo>±</mo> </msub> </semantics></math> when <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>3</mn> <mi>b</mi> <mo>≥</mo> <mn>4</mn> <mi>a</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">
10 pages, 250 KiB  
Article
(Non-Symmetric) Yetter–Drinfel’d Module Category and Invariant Coinvariant Jacobians
by Zhongwei Wang and Yong Wang
Symmetry 2024, 16(5), 515; https://doi.org/10.3390/sym16050515 - 24 Apr 2024
Viewed by 710
Abstract
In this paper, we generalize the homomorphisms of modules over groups and Lie algebras as being morphisms in the category of (non-symmetric) Yetter–Drinfel’d modules. These module homomorphisms play a key role in the conjecture of Yau. Full article
(This article belongs to the Special Issue Symmetry/Asymmetry Study in Hopf-Type Algebras and Groups)
578 KiB  
Proceeding Paper
A Diseased Three-Species Harvesting Food Web Model with Various Response Functions
by Thangavel Megala, Thangaraj Nandha Gopal, Manickasundaram Siva Pradeep and Arunachalam Yasotha
Biol. Life Sci. Forum 2024, 30(1), 17; https://doi.org/10.3390/IOCAG2023-16876 - 11 Mar 2024
Cited by 2 | Viewed by 736
Abstract
The purpose of this work is to present a three-species harvesting food web model that takes into account the interactions of susceptible prey, infected prey, and predator species. Prey species are assumed to expand logistically in the absence of predator species. The Crowley–Martin [...] Read more.
The purpose of this work is to present a three-species harvesting food web model that takes into account the interactions of susceptible prey, infected prey, and predator species. Prey species are assumed to expand logistically in the absence of predator species. The Crowley–Martin and Beddington–DeAngelis functional responses are used by predators to consume both susceptible and infected prey. Additionally, susceptible prey is consumed by infected prey in the formation of a Holling type II response. Both prey species are considered when prey harvesting is taken into account. Boundedness, positivity, and positive invariance are considered in this study. The investigation covers all the equilibrium points that are biologically feasible. Local stability is evaluated by analyzing the distribution of eigen values, while global stability is evaluated using suitable Lyapunov functions. Also, Hopf bifurcation is analyzed at the harvesting rate H1. At the end, we evaluate the numerical solutions based on our findings. Full article
(This article belongs to the Proceedings of The 2nd International Online Conference on Agriculture)
Show Figures

Figure 1

Figure 1
<p>Flowchart of the model with different functional responses.</p>
Full article ">Figure 2
<p>Dynamical changes in Model (<a href="#FD2-blsf-30-00017" class="html-disp-formula">2</a>) at harvesting rate <math display="inline"><semantics> <msub> <mi mathvariant="script">H</mi> <mn>1</mn> </msub> </semantics></math> = 0.47.</p>
Full article ">
13 pages, 1514 KiB  
Article
On the Bifurcations of a 3D Symmetric Dynamical System
by Dana Constantinescu
Symmetry 2023, 15(4), 923; https://doi.org/10.3390/sym15040923 - 15 Apr 2023
Cited by 2 | Viewed by 1547
Abstract
The paper studies the bifurcations that occur in the T-system, a 3D dynamical system symmetric in respect to the Oz axis. Results concerning some local bifurcations (pitchfork and Hopf bifurcation) are presented and our attention is focused on a special bifurcation, when the [...] Read more.
The paper studies the bifurcations that occur in the T-system, a 3D dynamical system symmetric in respect to the Oz axis. Results concerning some local bifurcations (pitchfork and Hopf bifurcation) are presented and our attention is focused on a special bifurcation, when the system has infinitely many equilibrium points. It is shown that, at the bifurcation limit, the phase space is foliated by infinitely many invariant surfaces, each of them containing two equilibrium points (an attractor and a saddle). For values of the bifurcation parameter close to the bifurcation limit, the study of the system’s dynamics is done according to the singular perturbation theory. The dynamics is characterized by mixed mode oscillations (also called fast-slow oscillations or oscillations-relaxations) and a finite number of equilibrium points. The specific features of the bifurcation are highlighted and explained. The influence of the pitchfork and Hopf bifurcations on the fast-slow dynamics is also pointed out. Full article
(This article belongs to the Special Issue Three-Dimensional Dynamical Systems and Symmetry)
Show Figures

Figure 1

Figure 1
<p>The pitchfork and Hopf bifurcation curves in the parameters’ plane.</p>
Full article ">Figure 2
<p>Orbits of the fast subsystem (5) restricted to the plane <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <msub> <mi>z</mi> <mn>0</mn> </msub> </mrow> </semantics></math>. (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1.6</mn> <mo>&gt;</mo> <mi>m</mi> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <msub> <mi>z</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1.4</mn> <mo>&lt;</mo> <mi>m</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>(<b>a</b>) The orbit of (0.2, 5, 0.2) in system (2) for <span class="html-italic">n</span> = 0.03 and <span class="html-italic">m</span> = 1.5 (fast-slow oscillations); (<b>b</b>) time series <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mi>z</mi> <mo stretchy="false">(</mo> <mi>t</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> from the orbit of (0.2, 3, 0.1) in system (2) for <span class="html-italic">m</span> = 1.5 and <span class="html-italic">n</span> = 0 (red curve), <span class="html-italic">n</span> = 0.00001 (magenta curve), <span class="html-italic">n</span> = 0.0001 (green curve), <span class="html-italic">n</span> = 0.001 (blue curve) and <span class="html-italic">n</span> = 0.01 (black curve).</p>
Full article ">Figure 4
<p>The synthesis of the dynamical behavior of some points situated in the plane <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mi>m</mi> </mrow> </semantics></math>, for <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> and (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>0.09</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>. The red points are in the basin of attraction of <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mo>+</mo> </msub> </mrow> </semantics></math>, the green points are in the basin of attraction of <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mo>−</mo> </msub> </mrow> </semantics></math> and the orbits of the blue points are ocsillating.</p>
Full article ">Figure 5
<p>The orbit of (−1, 4, −1) in system (2) for <span class="html-italic">n</span> = 0.03 and (<b>a</b>) <span class="html-italic">m</span> = 0.05 (after pitchfork bifurcation); (<b>b</b>) <span class="html-italic">m</span> = 1.06185 (before Hopf bifurcation).</p>
Full article ">
13 pages, 316 KiB  
Article
About the Jacobi Stability of a Generalized Hopf–Langford System through the Kosambi–Cartan–Chern Geometric Theory
by Florian Munteanu, Alexander Grin, Eduard Musafirov, Andrei Pranevich and Cătălin Şterbeţi
Symmetry 2023, 15(3), 598; https://doi.org/10.3390/sym15030598 - 26 Feb 2023
Cited by 6 | Viewed by 1559
Abstract
In this work, we will consider an autonomous three-dimensional quadratic system of first-order ordinary differential equations, with five parameters and with symmetry relative to the z-axis, which generalize the Hopf–Langford system. By reformulating the system as a system of two second-order ordinary [...] Read more.
In this work, we will consider an autonomous three-dimensional quadratic system of first-order ordinary differential equations, with five parameters and with symmetry relative to the z-axis, which generalize the Hopf–Langford system. By reformulating the system as a system of two second-order ordinary differential equations and using the Kosambi–Cartan–Chern (KCC) geometric theory, we will investigate this system from the perspective of Jacobi stability. We will compute the five invariants of KCC theory which determine the own geometrical properties of this system, especially the deviation curvature tensor. Additionally, we will search for necessary and sufficient conditions on the five parameters of the system in order to reach the Jacobi stability around each equilibrium point. Full article
(This article belongs to the Special Issue Geometric Algebra and Its Applications)
45 pages, 425 KiB  
Article
Wick Theorem and Hopf Algebra Structure in Causal Perturbative Quantum Field Theory
by D. R. Grigore
Universe 2023, 9(3), 117; https://doi.org/10.3390/universe9030117 - 24 Feb 2023
Cited by 4 | Viewed by 3197
Abstract
We consider the general framework of perturbative quantum field theory for the pure Yang–Mills model. We give a more precise version of the Wick theorem using Hopf algebra notations for chronological products and not for Feynman graphs. Next, we prove that the Wick [...] Read more.
We consider the general framework of perturbative quantum field theory for the pure Yang–Mills model. We give a more precise version of the Wick theorem using Hopf algebra notations for chronological products and not for Feynman graphs. Next, we prove that the Wick expansion property can be preserved for all cases in order n=2. However, gauge invariance is broken for chronological products of Wick submonomials. Full article
(This article belongs to the Section Field Theory)
17 pages, 328 KiB  
Article
Hopf Quasigroup Galois Extensions and a Morita Equivalence
by Huaiwen Guo and Shuanhong Wang
Mathematics 2023, 11(2), 273; https://doi.org/10.3390/math11020273 - 5 Jan 2023
Cited by 2 | Viewed by 1222
Abstract
For H, a Hopf coquasigroup, and A, a left quasi-H-module algebra, we show that the smash product A#H is linked to the algebra of H invariants AH by a Morita context. We use the Morita setting [...] Read more.
For H, a Hopf coquasigroup, and A, a left quasi-H-module algebra, we show that the smash product A#H is linked to the algebra of H invariants AH by a Morita context. We use the Morita setting to prove that for finite dimensional H, there are equivalent conditions for A/AH to be Galois parallel in the case of H finite dimensional Hopf algebra. Full article
(This article belongs to the Section A: Algebra and Logic)
14 pages, 285 KiB  
Article
The ∗-Ricci Operator on Hopf Real Hypersurfaces in the Complex Quadric
by Rongsheng Ma and Donghe Pei
Mathematics 2023, 11(1), 90; https://doi.org/10.3390/math11010090 - 26 Dec 2022
Viewed by 1335
Abstract
We study the ∗-Ricci operator on Hopf real hypersurfaces in the complex quadric. We prove that for Hopf real hypersurfaces in the complex quadric, the ∗-Ricci tensor is symmetric if and only if the unit normal vector field is singular. In the following, [...] Read more.
We study the ∗-Ricci operator on Hopf real hypersurfaces in the complex quadric. We prove that for Hopf real hypersurfaces in the complex quadric, the ∗-Ricci tensor is symmetric if and only if the unit normal vector field is singular. In the following, we obtain that if the ∗-Ricci tensor of Hopf real hypersurfaces in the complex quadric is symmetric, then the ∗-Ricci operator is both Reeb-flow-invariant and Reeb-parallel. As the correspondence to the semi-symmetric Ricci tensor, we give a classification of real hypersurfaces in the complex quadric with the semi-symmetric ∗-Ricci tensor. Full article
(This article belongs to the Special Issue Submanifolds in Metric Manifolds)
16 pages, 599 KiB  
Article
Stability Analysis of a Stage-Structure Predator–Prey Model with Holling III Functional Response and Cannibalism
by Yufen Wei and Yu Li
Axioms 2022, 11(8), 421; https://doi.org/10.3390/axioms11080421 - 21 Aug 2022
Viewed by 1804
Abstract
This paper considers the time taken for young predators to become adult predators as the delay and constructs a stage-structured predator–prey system with Holling III response and time delay. Using the persistence theory for infinite-dimensional systems and the Hurwitz criterion, the permanent persistence [...] Read more.
This paper considers the time taken for young predators to become adult predators as the delay and constructs a stage-structured predator–prey system with Holling III response and time delay. Using the persistence theory for infinite-dimensional systems and the Hurwitz criterion, the permanent persistence condition of this system and the local stability condition of the system’s coexistence equilibrium are given. Further, it is proven that the system undergoes a Hopf bifurcation at the coexistence equilibrium. By using Lyapunov functions and the LaSalle invariant principle, it is shown that the trivial equilibrium and the coexistence equilibrium are globally asymptotically stable, and sufficient conditions are derived for the global stability of the coexistence equilibrium. Some numerical simulations are carried out to illustrate the main results. Full article
Show Figures

Figure 1

Figure 1
<p>The positive equilibrium <math display="inline"><semantics> <msup> <mi>E</mi> <mo>∗</mo> </msup> </semantics></math> is stable at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>The positive equilibrium <math display="inline"><semantics> <msup> <mi>E</mi> <mo>∗</mo> </msup> </semantics></math> is stable at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>9</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>The positive equilibrium <math display="inline"><semantics> <msup> <mi>E</mi> <mo>∗</mo> </msup> </semantics></math> loses its stability and a Hopf bifurcation occurs at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>12.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>The temporal solution found by the numerical integration of Model (2) with <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msub> <mi>ϕ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>ϕ</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>ϕ</mi> <mn>3</mn> </msub> <mo>,</mo> <msub> <mi>ϕ</mi> <mn>4</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. (<b>a</b>) The non-negative equilibrium <math display="inline"><semantics> <msub> <mi>E</mi> <mn>1</mn> </msub> </semantics></math> is globally asymptotically stable. (<b>b</b>) The positive equilibrium <math display="inline"><semantics> <msup> <mi>E</mi> <mo>∗</mo> </msup> </semantics></math> is globally asymptotically stable.</p>
Full article ">
17 pages, 13358 KiB  
Article
Symmetry and Asymmetry in the Fluid Mechanical Sewing Machine
by Neil M. Ribe, Pierre-Thomas Brun and Basile Audoly
Symmetry 2022, 14(4), 772; https://doi.org/10.3390/sym14040772 - 8 Apr 2022
Cited by 5 | Viewed by 4296
Abstract
The ‘fluid mechanical sewing machine’ is a device in which a thin thread of viscous fluid falls onto a horizontal belt moving in its own plane, creating a rich variety of ‘stitch’ patterns depending on the fall height and the belt speed. This [...] Read more.
The ‘fluid mechanical sewing machine’ is a device in which a thin thread of viscous fluid falls onto a horizontal belt moving in its own plane, creating a rich variety of ‘stitch’ patterns depending on the fall height and the belt speed. This review article surveys the complex phenomenology of the patterns, their symmetries, and the mathematical models that have been used to understand them. The various patterns obey different symmetries that include (slightly imperfect) fore–aft symmetry relative to the direction of belt motion and invariance under reflection across a vertical plane containing the velocity vector of the belt, followed by a shift of one-half the wavelength. As the belt speed decreases, the first (Hopf) bifurcation is to a ‘meandering’ state whose frequency is equal to the frequency Ωc of steady coiling on a motionless surface. More complex patterns can be studied using direct numerical simulation via a novel ‘discrete viscous threads’ algorithm that yields the Fourier spectra of the longitudinal and transverse components of the motion of the contact point of the thread with the belt. The most intriguing case is the ‘alternating loops’ pattern, the spectra of which are dominated by the first five multiples of Ωc/3. A reduced (three-degrees-of-freedom) model succeeds in predicting the sequence of patterns observed as the belt speed decreases for relatively low fall heights for which inertia in the thread is negligible. Patterns that appear at greater fall heights seem to owe their existence to weakly nonlinear interaction between different ‘distributed pendulum’ modes of the quasi-vertical ‘tail’ of the thread. Full article
(This article belongs to the Special Issue Symmetry and Symmetry-Breaking in Fluid Dynamics)
Show Figures

Figure 1

Figure 1
<p>Behaviors of a viscous thread falling onto a motionless surface. (<b>a</b>) Steady coiling for <math display="inline"><semantics> <mrow> <mi>ν</mi> <mo>=</mo> <mn>6150</mn> </mrow> </semantics></math> cS, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>0.032</mn> </mrow> </semantics></math> mL·s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>H</mi> <mo>=</mo> <mn>7</mn> </mrow> </semantics></math> cm. The definitions of the coil radius <span class="html-italic">R</span> and the radius <math display="inline"><semantics> <msub> <mi>a</mi> <mn>1</mn> </msub> </semantics></math> of the thread at the contact point are shown. The structure comprises a long and nearly vertical tail, the coil, and a pile of fluid previously laid down by the coiling. (<b>b</b>) Axisymmetric stagnation flow for <math display="inline"><semantics> <mrow> <mi>ν</mi> <mo>=</mo> <mn>3450</mn> </mrow> </semantics></math> cS, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>3.55</mn> </mrow> </semantics></math> mL·s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>H</mi> <mo>=</mo> <mn>23.5</mn> </mrow> </semantics></math> cm. (<b>c</b>) Coiling with an unstable pile for <math display="inline"><semantics> <mrow> <mi>ν</mi> <mo>=</mo> <mn>3450</mn> </mrow> </semantics></math> cS, <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>0.68</mn> </mrow> </semantics></math> mL·s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>H</mi> <mo>=</mo> <mn>25</mn> </mrow> </semantics></math> cm.</p>
Full article ">Figure 2
<p>A laboratory apparatus for studying the FMSM. The units on the ruler are centimeters (<b>left</b>) and inches (<b>right</b>). Photograph courtesy of J. R. Lister.</p>
Full article ">Figure 3
<p>Stitch patterns observed in the FMSM. Photographs courtesy of J.R. Lister.</p>
Full article ">Figure 4
<p>Stitch patterns of the FMSM as a function of fall height <span class="html-italic">H</span> and belt speed <span class="html-italic">V</span>, for silicone oil with <math display="inline"><semantics> <mrow> <mi>ν</mi> <mo>=</mo> <mn>277</mn> </mrow> </semantics></math> S and <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>0.027</mn> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math>·s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. Upright and inverted triangles indicate observations made while increasing and decreasing the belt speed, respectively. Grey shaded regions indicate ranges of <span class="html-italic">H</span> for which the frequency of liquid thread coiling, calculated for the parameters of the experiment using the method of Ribe [<a href="#B18-symmetry-14-00772" class="html-bibr">18</a>], is multivalued. Figure adapted from Figure 3 of [<a href="#B20-symmetry-14-00772" class="html-bibr">20</a>].</p>
Full article ">Figure 5
<p>Coiling frequency <math display="inline"><semantics> <mo>Ω</mo> </semantics></math> as a function of the fall height <span class="html-italic">H</span> for <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>=</mo> <mn>0.97</mn> </mrow> </semantics></math> g·cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>ν</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>5</mn> </msup> </mrow> </semantics></math> cS, <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>21.5</mn> </mrow> </semantics></math> dyne·cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>0.068</mn> </mrow> </semantics></math> cm and <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>0.00215</mn> </mrow> </semantics></math> mL·s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. The red curve was calculated numerically using the method of Ribe [<a href="#B18-symmetry-14-00772" class="html-bibr">18</a>]. Portions of the curve corresponding to the different steady coiling regimes are indicated: V (viscous), G (gravitational), IG (inertiogravitational), and I (inertial). Dotted portions of the curve represent coiling that is unstable to small perturbations. Dashed lines with slope <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> are the first three pendulum frequencies of the tail of the thread, with the order of each mode indicated by the number to the left.</p>
Full article ">Figure 6
<p>Onset of meandering for golden syrup (<math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>=</mo> <mn>1.438</mn> </mrow> </semantics></math> g·cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>ν</mi> <mo>=</mo> <mn>350</mn> </mrow> </semantics></math> S), <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>=</mo> <mn>0.044</mn> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math> s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics></math> cm. (<b>a</b>) Critical belt speed <math display="inline"><semantics> <msub> <mi>V</mi> <mi>c</mi> </msub> </semantics></math> as a function of fall height. Circles: experimental measurements. Solid line: prediction of the linear stability analysis described in the text. Dotted line (nearly indistinguishable from the solid line): axial velocity <math display="inline"><semantics> <mrow> <msub> <mi>U</mi> <mn>1</mn> </msub> <mo>=</mo> <mi>Q</mi> <mo>/</mo> <mi>π</mi> <msubsup> <mi>a</mi> <mn>1</mn> <mn>2</mn> </msubsup> </mrow> </semantics></math> at the bottom of a thread coiling on a motionless surface for the same experimental parameters. (<b>b</b>) Angular frequency of oscillation <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>i</mi> </msub> </semantics></math> as a function of fall height. Circles: experimental measurements. Solid line: prediction of the linear stability analysis. Dotted line: angular frequency <math display="inline"><semantics> <mo>Ω</mo> </semantics></math> of steady liquid thread coiling on a motionless surface for the same experimental parameters. Figure adapted from Figures 5 and 7 of [<a href="#B19-symmetry-14-00772" class="html-bibr">19</a>].</p>
Full article ">Figure 7
<p>‘Gravitational heel’ shape of a dragged viscous thread at the onset of meandering, for the parameters of the experiment of <a href="#symmetry-14-00772-f006" class="html-fig">Figure 6</a>. The shape was calculated numerically using the method of Ribe et al. [<a href="#B19-symmetry-14-00772" class="html-bibr">19</a>]. The horizontal scale is exaggerated by a factor of 15 relative to the vertical scale. Figure adapted from Figure 6 of [<a href="#B19-symmetry-14-00772" class="html-bibr">19</a>].</p>
Full article ">Figure 8
<p>Phase diagram for the patterns at low fall heights <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo>≤</mo> <mn>0.8</mn> </mrow> </semantics></math>. Numerically predicted and experimentally observed patterns as functions of <math display="inline"><semantics> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> </semantics></math> and <math display="inline"><semantics> <mover accent="true"> <mi>V</mi> <mo stretchy="false">^</mo> </mover> </semantics></math> are indicated by vertical lines and circles, respectively. The patterns corresponding to each color are shown as insets. White spaces between vertical bars of different colors indicate ranges of belt speeds for which the automatic pattern recognition gave ambiguous results. Figure adapted from Figure 6 of [<a href="#B22-symmetry-14-00772" class="html-bibr">22</a>].</p>
Full article ">Figure 9
<p>Main portion: phase diagram of the FMSM constructed using DVT for <math display="inline"><semantics> <mrow> <msub> <mo>Π</mo> <mn>1</mn> </msub> <mo>=</mo> <mn>670</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mo>Π</mo> <mn>2</mn> </msub> <mo>=</mo> <mn>0.37</mn> </mrow> </semantics></math>, and no surface tension (<math display="inline"><semantics> <mrow> <msub> <mo>Π</mo> <mn>3</mn> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>). The patterns shown as insets include translated coiling (A, red), meanders (B, blue), alternating loops (C, green), disordered patterns (D, grey), double coiling (E, pink), double meanders (F, purple), and stretched coiling (H, yellow). The dimensionless frequency <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mo>Ω</mo> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> of liquid thread coiling is shown beneath the phase diagram, with multivalued intervals highlighted. Inset: phase diagram determined experimentally [<a href="#B20-symmetry-14-00772" class="html-bibr">20</a>] with <math display="inline"><semantics> <mrow> <msub> <mo>Π</mo> <mn>1</mn> </msub> <mo>=</mo> <mn>670</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mo>Π</mo> <mn>2</mn> </msub> <mo>=</mo> <mn>0.37</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mo>Π</mo> <mn>3</mn> </msub> <mo>=</mo> <mn>1.84</mn> </mrow> </semantics></math> (<b>top</b>). Additional patterns include the W pattern (yellow circled in black) and slanted loops (blue circled in black). The corresponding coiling frequency <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mo>Ω</mo> <mo stretchy="false">^</mo> </mover> <mi>c</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mover accent="true"> <mi>H</mi> <mo stretchy="false">^</mo> </mover> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> is also shown (<b>bottom</b>). Figure adapted from Figure 7 of [<a href="#B22-symmetry-14-00772" class="html-bibr">22</a>].</p>
Full article ">Figure 10
<p>Spectral characteristics of three FMSM patterns: translated coiling (<b>A</b>), alternating loops (<b>B</b>), and double coiling (<b>C</b>). See the main text for a detailed discussion. Figure adapted from Figure 10 of [<a href="#B22-symmetry-14-00772" class="html-bibr">22</a>].</p>
Full article ">Figure 11
<p>Geometry of a falling viscous thread in the vicinity of its contact point with a moving belt. Figure adapted from Figure 3 of [<a href="#B23-symmetry-14-00772" class="html-bibr">23</a>].</p>
Full article ">Figure 12
<p>Comparison of the predictions of the GM (green) and full DVT simulations (brown). See the main text for a detailed discussion. Figure adapted from Figure 5 of [<a href="#B23-symmetry-14-00772" class="html-bibr">23</a>].</p>
Full article ">
13 pages, 332 KiB  
Article
Gravitoelectromagnetic Knot Fields
by Adina Crişan, Cresus Godinho and Ion Vancea
Universe 2021, 7(3), 46; https://doi.org/10.3390/universe7030046 - 24 Feb 2021
Cited by 2 | Viewed by 1994
Abstract
We construct a class of knot solutions of the time-dependent gravitoelectromagnetic (GEM) equations in vacuum in the linearized gravity approximation by analogy with the Rañada–Hopf fields. For these solutions, the dual metric tensors of the bi-metric geometry of the gravitational vacuum with knot [...] Read more.
We construct a class of knot solutions of the time-dependent gravitoelectromagnetic (GEM) equations in vacuum in the linearized gravity approximation by analogy with the Rañada–Hopf fields. For these solutions, the dual metric tensors of the bi-metric geometry of the gravitational vacuum with knot perturbations are given and the geodesic equation as a function of two complex parameters of the time-dependent GEM knots are calculated. By taking stationary potentials, which formally amount to particularizing to time-independent GEM equations, we obtain a set of stationary fields subjected to constraints from the time-dependent GEM knots. Finally, the Landau–Lifshitz pseudo-tensor and a scalar invariant of the static fields are computed. Full article
(This article belongs to the Special Issue Frame-Dragging and Gravitomagnetism)
29 pages, 1572 KiB  
Article
Perturbative-Iterative Computation of Inertial Manifolds of Systems of Delay-Differential Equations with Small Delays
by Marc R. Roussel
Algorithms 2020, 13(9), 209; https://doi.org/10.3390/a13090209 - 27 Aug 2020
Viewed by 3629
Abstract
Delay-differential equations belong to the class of infinite-dimensional dynamical systems. However, it is often observed that the solutions are rapidly attracted to smooth manifolds embedded in the finite-dimensional state space, called inertial manifolds. The computation of an inertial manifold yields an ordinary differential [...] Read more.
Delay-differential equations belong to the class of infinite-dimensional dynamical systems. However, it is often observed that the solutions are rapidly attracted to smooth manifolds embedded in the finite-dimensional state space, called inertial manifolds. The computation of an inertial manifold yields an ordinary differential equation (ODE) model representing the long-term dynamics of the system. Note in particular that any attractors must be embedded in the inertial manifold when one exists, therefore reducing the study of these attractors to the ODE context, for which methods of analysis are well developed. This contribution presents a study of a previously developed method for constructing inertial manifolds based on an expansion of the delayed term in small powers of the delay, and subsequent solution of the invariance equation by the Fraser functional iteration method. The combined perturbative-iterative method is applied to several variations of a model for the expression of an inducible enzyme, where the delay represents the time required to transcribe messenger RNA and to translate that RNA into the protein. It is shown that inertial manifolds of different dimensions can be computed. Qualitatively correct inertial manifolds are obtained. Among other things, the dynamics confined to computed inertial manifolds display Andronov–Hopf bifurcations at similar parameter values as the original DDE model. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Trajectories of the ordinary differential equation (ODE) system (11) for <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, along with the singular limit of the slow manifold <math display="inline"><semantics> <mrow> <msub> <mi>s</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mi>e</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (Equation (<a href="#FD14b-algorithms-13-00209" class="html-disp-formula">14b</a>)). (<b>a</b>) Several trajectories along with <math display="inline"><semantics> <mrow> <msub> <mi>s</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mi>e</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> in an <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>×</mo> <mi>e</mi> <mo>×</mo> <mi>s</mi> </mrow> </semantics></math> projection. (<b>b</b>) A trajectory (oscillating) shown with <math display="inline"><semantics> <mrow> <msub> <mi>s</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mi>e</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> in an <math display="inline"><semantics> <mrow> <mi>e</mi> <mo>×</mo> <mi>s</mi> </mrow> </semantics></math> projection.</p>
Full article ">Figure 2
<p>Computed two-dimensional slow manifold of the ODE model (11) for the same parameters as <a href="#algorithms-13-00209-f001" class="html-fig">Figure 1</a>. A trajectory started from a corner of the computed surface is shown to demonstrate the approximate invariance of the surface. The manifold was computed by taking four iterates of Equation (17) starting from the initial function (14). (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mi>e</mi> <mo>)</mo> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mi>e</mi> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Comparison of the trajectories computed with the exact and reduced ODE models, systems (11) and (18) respectively, both started from an initial condition on the computed slow manifold.</p>
Full article ">Figure 4
<p>(<b>a</b>) Approximate inertial manifolds for the three-variable model obtained at the same level of iteration (<math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>) for different orders of expansion of the delayed term. Parameters: <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>1</mn> </msub> <mo>=</mo> <mi>β</mi> <mo>=</mo> <mi>α</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>. On the scale of this figure, regardless of how we rotate the graph, the three surfaces appear very similar to the eye. (<b>b</b>) Difference between <math display="inline"><semantics> <msubsup> <mi mathvariant="script">C</mi> <mn>3</mn> <mrow> <mo>(</mo> <mn>3</mn> <mo>)</mo> </mrow> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mi mathvariant="script">C</mi> <mn>3</mn> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> </semantics></math>.</p>
Full article ">Figure 5
<p>Trajectories of (<b>a</b>) the three-variable model and of (<b>b</b>) the <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> reduced models for the same parameters as <a href="#algorithms-13-00209-f004" class="html-fig">Figure 4</a>. All trajectories were started from initial conditions near the origin, <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. For the three-variable model, <math display="inline"><semantics> <mrow> <mi>e</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> and the initial function was <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>(</mo> <mi>θ</mi> <mo>)</mo> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>&lt;</mo> <mn>0</mn> </mrow> </semantics></math> (i.e. there is a small discontinuity in the initial function at <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 6
<p>Trajectories of (<b>a</b>) the three-variable DDE model and of (<b>b</b>) the <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> reduced ODE models. The <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> model has an Andronov–Hopf bifurcation and can generate limit-cycle oscillations, like the three-variable model, whereas the <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> model never undergoes an Andronov–Hopf bifurcation. All parameters are as in <a href="#algorithms-13-00209-f004" class="html-fig">Figure 4</a>, except <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>0.7</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Scaled iteration differences calculated using Equation (<a href="#FD35-algorithms-13-00209" class="html-disp-formula">35</a>) for <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> for the parameters <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>0.9</mn> </mrow> </semantics></math>. Trajectories used in these calculations were started from <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>e</mi> <mo>,</mo> <mi>s</mi> <mo>,</mo> <mi>c</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>0</mn> <mo>,</mo> <mn>0</mn> <mo>,</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Trajectories of (<b>a</b>) the DDE inducible enzyme model, (<b>b</b>) the reduced model for <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, (<b>c</b>) the reduced model for <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, and (<b>d</b>) the reduced model for <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. For the reduced models, the <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> iterate was used. The parameters are the same as in <a href="#algorithms-13-00209-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 9
<p>Computed inertial manifolds for <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. (<b>a</b>) <math display="inline"><semantics> <msubsup> <mi mathvariant="script">C</mi> <mn>4</mn> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> </semantics></math>, a representative approximation of the inertial manifold <math display="inline"><semantics> <mrow> <mi mathvariant="script">C</mi> <mo>(</mo> <mi>e</mi> <mo>,</mo> <mi>s</mi> <mo>)</mo> </mrow> </semantics></math>. (<b>b</b>) <math display="inline"><semantics> <msubsup> <mi mathvariant="script">R</mi> <mn>4</mn> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> </semantics></math>. (<b>c</b>) <math display="inline"><semantics> <msubsup> <mi mathvariant="script">R</mi> <mn>4</mn> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> </semantics></math>. (<b>d</b>) <math display="inline"><semantics> <msubsup> <mi mathvariant="script">R</mi> <mn>4</mn> <mrow> <mo>(</mo> <mn>3</mn> <mo>)</mo> </mrow> </msubsup> </semantics></math>.</p>
Full article ">Figure 10
<p>Trajectories of the DDE model (21) and of the system reduced to the inertial manifold for <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>. The parameters for the first two curves are those of <a href="#algorithms-13-00209-f009" class="html-fig">Figure 9</a>. In particular, <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. For the third curve, the DDE model was integrated with <math display="inline"><semantics> <mi>δ</mi> </semantics></math> increased to 7.</p>
Full article ">
23 pages, 7663 KiB  
Article
A Haptic Model of Entanglement, Gauge Symmetries and Minimal Interaction Based on Knot Theory
by Stefan Heusler and Malte Ubben
Symmetry 2019, 11(11), 1399; https://doi.org/10.3390/sym11111399 - 12 Nov 2019
Cited by 4 | Viewed by 4059
Abstract
The Heegaard splitting of S U ( 2 ) is a particularly useful representation for quantum phases of spin j-representation arising in the mapping S 1 S 3, which can be related to ( 2 j , 2 ) torus [...] Read more.
The Heegaard splitting of S U ( 2 ) is a particularly useful representation for quantum phases of spin j-representation arising in the mapping S 1 S 3, which can be related to ( 2 j , 2 ) torus knots in Hilbert space. We show that transitions to homotopically equivalent knots can be associated with gauge invariance, and that the same mechanism is at the heart of quantum entanglement. In other words, (minimal) interaction causes entanglement. Particle creation is related to cuts in the knot structure. We show that inner twists can be associated with operations with the quaternions ( I , J , K ), which are crucial to understand the Hopf mapping S 3 S 2. We discuss the relationship between observables on the Bloch sphere S 2, and knots with inner twists in Hilbert space. As applications, we discuss selection rules in atomic physics, and the status of virtual particles arising in Feynman diagrams. Using a simple paper strip model revealing the knot structure of quantum phases in Hilbert space including inner twists, a h a p t i c model of entanglement and gauge symmetries is proposed, which may also be valid for physics education. Full article
Show Figures

Figure 1

Figure 1
<p>Quantum tomography: For an ensemble of identical qubits, successive measurements lead to a random pattern with probabilities for ‘black’ (<math display="inline"><semantics> <mrow> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> eigenvalues) or ‘white’ (<math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> eigenvalues) in the <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>,</mo> <mi>y</mi> <mo>,</mo> <mi>z</mi> </mrow> </semantics></math> direction, respectively. With <math display="inline"><semantics> <msubsup> <mi>p</mi> <mi>k</mi> <mo>+</mo> </msubsup> </semantics></math> as probability for ‘black’, and <math display="inline"><semantics> <msubsup> <mi>p</mi> <mi>k</mi> <mo>−</mo> </msubsup> </semantics></math> as probability for ‘white’, the relation <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>k</mi> </msub> <mo>=</mo> <mi>tr</mi> <mrow> <mo stretchy="false">(</mo> <msub> <mi>σ</mi> <mi>k</mi> </msub> <mi>ρ</mi> <mo stretchy="false">)</mo> </mrow> <mo>=</mo> <msubsup> <mi>p</mi> <mi>k</mi> <mo>+</mo> </msubsup> <mo>−</mo> <msubsup> <mi>p</mi> <mi>k</mi> <mo>−</mo> </msubsup> </mrow> </semantics></math> holds. Here, <math display="inline"><semantics> <mrow> <mrow> <mi>ρ</mi> <mo>=</mo> <mo> </mo> <mo stretchy="false">|</mo> </mrow> <msub> <mn>0</mn> <mi>n</mi> </msub> <mrow> <mo>〉</mo> <mo>〈</mo> </mrow> <msub> <mn>0</mn> <mi>n</mi> </msub> <mrow> <mo stretchy="false">|</mo> </mrow> </mrow> </semantics></math> is the <math display="inline"><semantics> <mrow> <mn>2</mn> <mo>×</mo> <mn>2</mn> </mrow> </semantics></math> density matrix of the single qubit.</p>
Full article ">Figure 2
<p>The operation of the quaternions <math display="inline"><semantics> <mrow> <mi>I</mi> <mo>,</mo> <mi>J</mi> <mo>,</mo> <mi>K</mi> </mrow> </semantics></math> on the Dirac belt in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm. Note that these operations lead to inner twists of the Dirac belt. In particular, for operation <math display="inline"><semantics> <mrow> <mi>I</mi> <mi>J</mi> <mi>K</mi> <mo>=</mo> <msup> <mi>e</mi> <mrow> <mn>2</mn> <mi>π</mi> <mo>/</mo> <mn>2</mn> </mrow> </msup> <mo>=</mo> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> leads to two inner twists.</p>
Full article ">Figure 3
<p>Left: Heegaard splitting of <math display="inline"><semantics> <msub> <mi>S</mi> <mn>3</mn> </msub> </semantics></math>, Right: Hopf mapping to the Bloch sphere. In <math display="inline"><semantics> <msub> <mi>S</mi> <mn>3</mn> </msub> </semantics></math>, a homotopic loop <math display="inline"><semantics> <msub> <mi>S</mi> <mn>1</mn> </msub> </semantics></math> is considered ranging from <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>0</mn> <mo>,</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>, which is mapped to a great circle <math display="inline"><semantics> <msub> <mi>S</mi> <mn>1</mn> </msub> </semantics></math> traversed <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>w</mi> <mi>i</mi> <mi>c</mi> <mi>e</mi> </mrow> </semantics></math> on the Bloch sphere <math display="inline"><semantics> <msub> <mi>S</mi> <mn>2</mn> </msub> </semantics></math>. The Dirac belt describing the quantum phase on the homotopic loop in <math display="inline"><semantics> <msub> <mi>S</mi> <mn>3</mn> </msub> </semantics></math> is equivalent to a Möbius strip when the parts <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo>,</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> separated in <math display="inline"><semantics> <msub> <mi>S</mi> <mn>3</mn> </msub> </semantics></math> are ‘glued together’, [see also <a href="#symmetry-11-01399-f004" class="html-fig">Figure 4</a>]. Superposition of right- and left-twisted Möbius strips leads to a node on the Bloch sphere. The antipode of this node is called ‘direction of the spin’, describing the direction of maximal amplitude.</p>
Full article ">Figure 4
<p>The node on the Bloch sphere [in this case, the node at <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>,</mo> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> arises due to superposition of the quantum phases <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <msqrt> <mn>2</mn> </msqrt> <mrow> <mo stretchy="false">(</mo> <msup> <mi>e</mi> <mrow> <mi>i</mi> <mi>ϕ</mi> <mo>/</mo> <mn>2</mn> </mrow> </msup> <mo>−</mo> <msup> <mi>e</mi> <mrow> <mo>−</mo> <mi>i</mi> <mi>ϕ</mi> <mo>/</mo> <mn>2</mn> </mrow> </msup> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math>]. In the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm, an infinite number of homotopically equivalent quantum phases in <math display="inline"><semantics> <mrow> <msub> <mi>S</mi> <mn>1</mn> </msub> <mo stretchy="false">→</mo> <msub> <mi>S</mi> <mn>3</mn> </msub> </mrow> </semantics></math> arise which all map to the great circle <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mi>θ</mi> <mo>=</mo> <mi>π</mi> <mo>/</mo> <mn>2</mn> <mo>,</mo> <mi>ϕ</mi> <mo>∈</mo> <mo>{</mo> <mn>0</mn> <mo>,</mo> <mn>4</mn> <mi>π</mi> <mo>}</mo> <mo stretchy="false">)</mo> </mrow> </semantics></math>. The superposition on the Bloch sphere translates to a superposition of Dirac belts.</p>
Full article ">Figure 5
<p>Representation of the superposition state <math display="inline"><semantics> <mrow> <mrow> <mo stretchy="false">|</mo> <mo>−</mo> <mo>〉</mo> </mrow> <mo>=</mo> <mfrac> <mn>1</mn> <msqrt> <mn>2</mn> </msqrt> </mfrac> <mrow> <mo stretchy="false">(</mo> <mo stretchy="false">|</mo> <mn>0</mn> <mo>〉</mo> </mrow> <mo>−</mo> <mrow> <mo stretchy="false">|</mo> <mn>1</mn> <mo>〉</mo> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> on the Bloch sphere (<math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm). Positions × of nodes are antipodes to the direction of <math display="inline"><semantics> <mover accent="true"> <mi>n</mi> <mo stretchy="false">→</mo> </mover> </semantics></math>. The superposition of the qubits <math display="inline"><semantics> <mrow> <mo stretchy="false">|</mo> <mn>0</mn> <mo>〉</mo> <mo>,</mo> <mo stretchy="false">|</mo> <mn>1</mn> <mo>〉</mo> </mrow> </semantics></math> changes the position of the direction of the spin, which in turn changes the position × of the node. The relation between nodes on the Bloch sphere and knots in Hilbert space are shown in <a href="#symmetry-11-01399-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 6
<p>For spin <span class="html-italic">j</span>-states, <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>j</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> nodes arise on the Bloch sphere which may be described as a complex function <math display="inline"><semantics> <mrow> <msubsup> <mo>∏</mo> <mrow> <mi>k</mi> <mo>=</mo> <mn>1</mn> </mrow> <mrow> <mn>2</mn> <mi>j</mi> </mrow> </msubsup> <mrow> <mo stretchy="false">(</mo> <mi>z</mi> <mo>−</mo> <msub> <mi>z</mi> <mi>k</mi> </msub> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> in the stellar representation. The corresponding knot structure in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm is described by the Jones polynomial <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mi>j</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>t</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math>. Additionally, the quantum phase has <math display="inline"><semantics> <mrow> <mn>4</mn> <mi>j</mi> </mrow> </semantics></math> inner twists (bosons), or <math display="inline"><semantics> <mrow> <mn>4</mn> <mi>j</mi> <mo>+</mo> <mn>2</mn> </mrow> </semantics></math> inner twists (fermions), respectively, denoted by (+).</p>
Full article ">Figure 7
<p>Left: Double copy of <math display="inline"><semantics> <mi>ν</mi> </semantics></math> inner twists, describing a boson with spin <math display="inline"><semantics> <mrow> <mn>2</mn> <mi>ν</mi> </mrow> </semantics></math>. Right: By joining the two pieces to a fermionic knot, two additional inner twists arise, leading to a knot with <math display="inline"><semantics> <mrow> <mn>4</mn> <mi mathvariant="bold">j</mi> <mo>±</mo> <mn>2</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Due to the interaction <math display="inline"><semantics> <mrow> <mi mathvariant="bold">U</mi> <mo stretchy="false">(</mo> <mi>t</mi> <mo stretchy="false">)</mo> <mo>=</mo> <mo form="prefix">exp</mo> <mrow> <mo>−</mo> <mo stretchy="false">(</mo> <mi>i</mi> <mi mathvariant="bold">H</mi> <mi>t</mi> <mo>/</mo> <mi>ℏ</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi mathvariant="bold">H</mi> <mo>=</mo> <mi>ℏ</mi> <mi>ω</mi> <mo stretchy="false">(</mo> <msub> <mi>σ</mi> <mi>z</mi> </msub> <mo>×</mo> <msub> <mi>σ</mi> <mi>z</mi> </msub> <mo stretchy="false">)</mo> </mrow> </semantics></math>, the initial state <math display="inline"><semantics> <mrow> <mo stretchy="false">|</mo> <mo>+</mo> <mo>〉</mo> <mo stretchy="false">|</mo> <mo>+</mo> <mo>〉</mo> </mrow> </semantics></math> becomes the entangled Bell state <math display="inline"><semantics> <mrow> <mfrac> <mn>1</mn> <msqrt> <mn>2</mn> </msqrt> </mfrac> <mrow> <mo stretchy="false">(</mo> <mo stretchy="false">|</mo> <msub> <mn>0</mn> <mi>a</mi> </msub> <msub> <mn>1</mn> <mi>a</mi> </msub> <mo>〉</mo> </mrow> <mo>+</mo> <mrow> <mo stretchy="false">|</mo> <msub> <mn>1</mn> <mi>a</mi> </msub> <msub> <mn>0</mn> <mi>a</mi> </msub> <mo>〉</mo> </mrow> </mrow> </semantics></math>. We consider the homotopic loop perpendicular to the <math display="inline"><semantics> <mrow> <mrow> <mo stretchy="false">|</mo> </mrow> <msub> <mn>0</mn> <mi>a</mi> </msub> <mrow> <mo>〉</mo> <mo>↔</mo> <mo stretchy="false">|</mo> </mrow> <msub> <mn>1</mn> <mi>a</mi> </msub> <mrow> <mo>〉</mo> </mrow> </mrow> </semantics></math>-direction.</p>
Full article ">Figure 9
<p>Using transitions to homotopically equivalent knots similar to Type-2 Reidemeister moves in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm, the constant phase can also be viewed as a combination of two qubits <math display="inline"><semantics> <mrow> <mo stretchy="false">|</mo> <mn>0</mn> <mo>〉</mo> <mo stretchy="false">|</mo> <mn>1</mn> <mo>〉</mo> </mrow> </semantics></math> or <math display="inline"><semantics> <mrow> <mo stretchy="false">|</mo> <mn>1</mn> <mo>〉</mo> <mo stretchy="false">|</mo> <mn>0</mn> <mo>〉</mo> </mrow> </semantics></math>. Since both possibilities are indistinguishable, the superposition <math display="inline"><semantics> <mrow> <mo stretchy="false">|</mo> <mn>0</mn> <mo>〉</mo> <mo stretchy="false">|</mo> <mn>1</mn> <mo>〉</mo> <mo>+</mo> <mo stretchy="false">|</mo> <mn>1</mn> <mo>〉</mo> <mo stretchy="false">|</mo> <mn>0</mn> <mo>〉</mo> </mrow> </semantics></math> emerges, which is an entangled state. After taking the particle trace, a mixed state arises. The latter result is basis-independent.</p>
Full article ">Figure 10
<p>Paper strip model of the interaction of two qubits: The entangled state is described by one common state with one quantum phase. By rotating the phase once and cutting into two pieces, two separate particles with left- and right twists arise. Here, we use the model of the paper strip near the boundary, where the phases <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo>,</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> are glued together. The general homotopy in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm is shown in <a href="#sec6dot3-symmetry-11-01399" class="html-sec">Section 6.3</a>.</p>
Full article ">Figure 11
<p>In the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm, the constant phase in <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> can superpose constructively (symmetric wave function) <math display="inline"><semantics> <mrow> <mrow> <mo stretchy="false">|</mo> </mrow> <msup> <mi mathvariant="sans-serif">Ψ</mi> <mo>+</mo> </msup> <mrow> <mo>〉</mo> </mrow> </mrow> </semantics></math> or destructively (anti symmetric wave function) <math display="inline"><semantics> <mrow> <mrow> <mo stretchy="false">|</mo> </mrow> <msup> <mi mathvariant="sans-serif">Ψ</mi> <mo>−</mo> </msup> <mrow> <mo>〉</mo> </mrow> </mrow> </semantics></math> with the constant phase in <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo>,</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>. The situation on the Bloch sphere in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm is shown in <a href="#symmetry-11-01399-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 12
<p>Paper strip model of minimal interaction: The insertion of additional twists is compensated by a gauge field. Only after a torus splitting, the gauge field with <math display="inline"><semantics> <mrow> <mo>−</mo> <mi>T</mi> </mrow> </semantics></math> twists is separated from the original particle, which then has <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>+</mo> <mi>T</mi> </mrow> </semantics></math> twists due to the gauge interaction.</p>
Full article ">Figure 13
<p>The quantum phase of the state <math display="inline"><semantics> <mrow> <mo stretchy="false">|</mo> <mn>3</mn> <mi>s</mi> <mo>〉</mo> </mrow> </semantics></math> is homotopically equivalent to <math display="inline"><semantics> <mrow> <mfrac> <mn>1</mn> <msqrt> <mn>2</mn> </msqrt> </mfrac> <mrow> <mrow> <mo stretchy="false">(</mo> <mo stretchy="false">|</mo> <mn>2</mn> <mi>p</mi> <mo>,</mo> <mo>+</mo> <mn>1</mn> <mo>〉</mo> </mrow> <mo stretchy="false">|</mo> <mi>L</mi> <mo>〉</mo> <mo>+</mo> <mo stretchy="false">|</mo> <mn>2</mn> <mi>p</mi> <mo>,</mo> <mo>−</mo> <mn>1</mn> <mo>〉</mo> <mo stretchy="false">|</mo> <mi>R</mi> <mo>〉</mo> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math>, see also <a href="#symmetry-11-01399-f009" class="html-fig">Figure 9</a>. The entangled state decays into a mixed state due to interaction of the photon with the environment.</p>
Full article ">Figure 14
<p>Paper strip model of the quantum phase of the decay <math display="inline"><semantics> <mrow> <mo stretchy="false">|</mo> <mn>3</mn> <mi>s</mi> <mo>〉</mo> <mo stretchy="false">→</mo> <mo stretchy="false">|</mo> <mn>2</mn> <mi>p</mi> <mo>〉</mo> </mrow> </semantics></math> in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm. If <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>R</mi> </mrow> </semantics></math> is associated with the right-circular polarized photon, the <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>L</mi> </mrow> </semantics></math> described the quantum phase of the electron in the state <math display="inline"><semantics> <mrow> <mo stretchy="false">|</mo> <mn>2</mn> <mi>p</mi> <mo>,</mo> <mo>−</mo> <mn>1</mn> <mo>〉</mo> </mrow> </semantics></math>. With <math display="inline"><semantics> <mrow> <mn>50</mn> <mo>%</mo> </mrow> </semantics></math> probability, the roles of the pieces <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>R</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>L</mi> </mrow> </semantics></math> are interchanged, see <a href="#symmetry-11-01399-f013" class="html-fig">Figure 13</a>.</p>
Full article ">Figure 15
<p>Leading-order Feynman diagrams for electron-electron and electron-positron interaction. In view of entanglement, a direct interpretation of Feynman diagrams in space time is impossible. The quantum phase of the entangled state can be modeled in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm as shown in <a href="#symmetry-11-01399-f010" class="html-fig">Figure 10</a>, and in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm as shown in <a href="#symmetry-11-01399-f016" class="html-fig">Figure 16</a>.</p>
Full article ">Figure 16
<p>Haptic model of the quantum phase of an entangled pair of qubits in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm. Here, we consider any homotopy in the bulk of <math display="inline"><semantics> <msub> <mi>B</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>B</mi> <mn>2</mn> </msub> </semantics></math> in the Heegaard splitting, see <a href="#symmetry-11-01399-f003" class="html-fig">Figure 3</a>. After Hopf mapping, the pieces <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo>,</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> from <math display="inline"><semantics> <msub> <mi>B</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>B</mi> <mn>2</mn> </msub> </semantics></math> are ‘glued together’, leading to the description in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm as shown in <a href="#symmetry-11-01399-f010" class="html-fig">Figure 10</a>. As discussed in the text, we associate with the two qubits either a pair of electrons, or an electron-positron pair. (<b>A</b>) The constant phase, see also <a href="#symmetry-11-01399-f011" class="html-fig">Figure 11</a>. (<b>B</b>) One rotation leads to a homotopically equivalent configuration, which can be seen as a pair of entangled spin <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> particles, see also <a href="#symmetry-11-01399-f010" class="html-fig">Figure 10</a>A. Due to the homotopic equivalences shown in <a href="#symmetry-11-01399-f009" class="html-fig">Figure 9</a>, we may view this quantum state also a combination of two spin <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> particles with a (virtual) gauge particle. (<b>C</b>) Encounter of the quantum phases, see also <a href="#symmetry-11-01399-f010" class="html-fig">Figure 10</a>B. (<b>D</b>) First splitting of the quantum phase in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm, see also <a href="#symmetry-11-01399-f0A1" class="html-fig">Figure A1</a> for the corresponding Jones polynomials. This configuration cannot be mapped to the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm. (<b>E</b>) Second splitting of the quantum phase in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm. The situation in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm, where both splitting are combined to one torus splitting is shown in <a href="#symmetry-11-01399-f010" class="html-fig">Figure 10</a>C,D. (<b>f</b>) Quantum phase of two distinguishable spin <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> particles in the <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>4</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>-realm. The inner twists in the Dirac belts have opposite sign, i.e., <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mo>+</mo> <mo>+</mo> <mo>+</mo> <mo>+</mo> <mo stretchy="false">)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mo>−</mo> <mo>−</mo> <mo>−</mo> <mo>−</mo> <mo stretchy="false">)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure A1
<p>Calculation of Jones polynomials for the torus splitting described in the main text (<a href="#symmetry-11-01399-f016" class="html-fig">Figure 16</a>).</p>
Full article ">
18 pages, 714 KiB  
Article
A Delayed Epidemic Model for Propagation of Malicious Codes in Wireless Sensor Network
by Zizhen Zhang, Soumen Kundu and Ruibin Wei
Mathematics 2019, 7(5), 396; https://doi.org/10.3390/math7050396 - 1 May 2019
Cited by 30 | Viewed by 3028
Abstract
In this paper, we investigate a delayed SEIQRS-V epidemic model for propagation of malicious codes in a wireless sensor network. The communication radius and distributed density of nodes is considered in the proposed model. With this model, first we find a feasible region [...] Read more.
In this paper, we investigate a delayed SEIQRS-V epidemic model for propagation of malicious codes in a wireless sensor network. The communication radius and distributed density of nodes is considered in the proposed model. With this model, first we find a feasible region which is invariant and where the solutions of our model are positive. To show that the system is locally asymptotically stable, a Lyapunov function is constructed. After that, sufficient conditions for local stability and existence of Hopf bifurcation are derived by analyzing the distribution of the roots of the corresponding characteristic equation. Finally, numerical simulations are presented to verify the obtained theoretical results and to analyze the effects of some parameters on the dynamical behavior of the proposed model in the paper. Full article
Show Figures

Figure 1

Figure 1
<p>Time plots of <span class="html-italic">S</span>, <span class="html-italic">E</span>, <span class="html-italic">I</span>, <span class="html-italic">Q</span>, <span class="html-italic">R</span> and <span class="html-italic">V</span> with <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>12.85</mn> <mo>&lt;</mo> <msub> <mi>τ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>13.1047</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Time plots of <span class="html-italic">S</span>, <span class="html-italic">E</span>, <span class="html-italic">I</span>, <span class="html-italic">Q</span>, <span class="html-italic">R</span> and <span class="html-italic">V</span> with <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>13.75</mn> <mo>&gt;</mo> <msub> <mi>τ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>13.1047</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Bifurcation diagram with respect to time delay of system (<a href="#FD22-mathematics-07-00396" class="html-disp-formula">22</a>): (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>−</mo> <mi>τ</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>−</mo> <mi>τ</mi> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>I</mi> <mo>−</mo> <mi>τ</mi> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>−</mo> <mi>τ</mi> </mrow> </semantics></math>, (<b>e</b>) <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>−</mo> <mi>τ</mi> </mrow> </semantics></math>, (<b>f</b>) <math display="inline"><semantics> <mrow> <mi>V</mi> <mo>−</mo> <mi>τ</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Time plots of <span class="html-italic">I</span> for different <math display="inline"><semantics> <msub> <mi>η</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>η</mi> <mn>2</mn> </msub> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>12.85</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Dynamic behavior of system (<a href="#FD22-mathematics-07-00396" class="html-disp-formula">22</a>): projection on I-Q-R for different <math display="inline"><semantics> <msub> <mi>η</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>η</mi> <mn>2</mn> </msub> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>13.75</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Time plots of <span class="html-italic">I</span> for different <math display="inline"><semantics> <mi>φ</mi> </semantics></math> and <math display="inline"><semantics> <mi>ε</mi> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>12.85</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Dynamic behavior of system (<a href="#FD22-mathematics-07-00396" class="html-disp-formula">22</a>): projection on I-Q-R for different <math display="inline"><semantics> <mi>φ</mi> </semantics></math> and <math display="inline"><semantics> <mi>ε</mi> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>12.85</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Time plots of <span class="html-italic">I</span> for different <math display="inline"><semantics> <mi>φ</mi> </semantics></math> and <math display="inline"><semantics> <mi>ε</mi> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>8.85</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Dynamic behavior of system (<a href="#FD22-mathematics-07-00396" class="html-disp-formula">22</a>): projection on I-Q-R for different <math display="inline"><semantics> <mi>φ</mi> </semantics></math> and <math display="inline"><semantics> <mi>ε</mi> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>9.25</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Other parameters are as in the text.</p>
Full article ">
Back to TopTop