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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (49)

Search Parameters:
Keywords = asymptotic perturbation theory

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 774 KiB  
Article
An Infinitely Old Universe with Planck Fields Before and After the Big Bang
by Dragana Pilipović
Universe 2024, 10(10), 400; https://doi.org/10.3390/universe10100400 - 17 Oct 2024
Viewed by 1041
Abstract
The Robertson–Walker minimum length (RWML) theory considers stochastically perturbed spacetime to describe an expanding universe governed by geometry and diffusion. We explore the possibility of static, torsionless universe eras with conserved energy density. We find that the RWML theory provides asymptotically static equations [...] Read more.
The Robertson–Walker minimum length (RWML) theory considers stochastically perturbed spacetime to describe an expanding universe governed by geometry and diffusion. We explore the possibility of static, torsionless universe eras with conserved energy density. We find that the RWML theory provides asymptotically static equations of state under positive curvature both far in the past and far into the future, with a Big Bang singularity in between. Full article
(This article belongs to the Special Issue Probing the Early Universe)
Show Figures

Figure 1

Figure 1
<p>Universe with curvature: equation of state (<b>left</b>) and acceleration (<b>right</b>). In both plots, the <span class="html-italic">x</span>-axis is the ratio of diffusion parameter D to the Hubble parameter H: X <math display="inline"><semantics> <mrow> <mo>≡</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>D</mi> <mi>H</mi> </mfrac> </mstyle> </mrow> </semantics></math>. The equation of state plot shows the ratio of energy density to pressure, both provided by the extended RWML Friedmann equations. For the acceleration plot, <span class="html-italic">y</span>-axis is given by −(1+q) <math display="inline"><semantics> <mrow> <mo>≡</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mover accent="true"> <mi>H</mi> <mo>˙</mo> </mover> <msup> <mi>H</mi> <mn>2</mn> </msup> </mfrac> </mstyle> </mrow> </semantics></math>. The arrows on the plots correspond to the arrow of time.</p>
Full article ">Figure 2
<p>Flat universe equation of state–zoom in view after the Big Bang.</p>
Full article ">Figure 3
<p>Flat universe equation of state–zoom in view before the Big Bang.</p>
Full article ">Figure 4
<p>Universe with curvature equation of state–zoom in view after the Big Bang.</p>
Full article ">Figure 5
<p>Universe with curvature equation of state–zoom in view before the Big Bang.</p>
Full article ">Figure 6
<p>Universe acceleration with positive and zero curvature.</p>
Full article ">Figure A1
<p>Nineteen SNe Ia galaxies with distances calibrated using Cepheids data from Table 1 in Riess et al. (2022) [<a href="#B7-universe-10-00400" class="html-bibr">7</a>,<a href="#B21-universe-10-00400" class="html-bibr">21</a>,<a href="#B22-universe-10-00400" class="html-bibr">22</a>,<a href="#B23-universe-10-00400" class="html-bibr">23</a>,<a href="#B24-universe-10-00400" class="html-bibr">24</a>,<a href="#B25-universe-10-00400" class="html-bibr">25</a>,<a href="#B26-universe-10-00400" class="html-bibr">26</a>] corresponding to redshift data <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mrow> <mi>h</mi> <mi>e</mi> <mi>l</mi> <mi>i</mi> <mi>o</mi> </mrow> </msub> <mo>&lt;</mo> <mn>0.0177</mn> </mrow> </semantics></math>, along with <math display="inline"><semantics> <msub> <mi>z</mi> <mrow> <mi>h</mi> <mi>e</mi> <mi>l</mi> <mi>i</mi> <mi>o</mi> </mrow> </msub> </semantics></math> data from PantheonPlusSHOES Github [<a href="#B27-universe-10-00400" class="html-bibr">27</a>]. The RWML model fit was performed using <math display="inline"><semantics> <msub> <mi>H</mi> <mn>0</mn> </msub> </semantics></math> = 67.4 km s<sup>−1</sup> Mpc<sup>−1</sup>. The <math display="inline"><semantics> <mo>Λ</mo> </semantics></math>CDM model used <math display="inline"><semantics> <mrow> <mi>q</mi> <mo>=</mo> <mo>−</mo> <mn>0.55</mn> </mrow> </semantics></math>.</p>
Full article ">Figure A2
<p>Natural log of luminosity distance fitted with an implementation of the RWML model for a positive curvature in ROOT with user-defined functions using <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mrow> <mi>h</mi> <mi>e</mi> <mi>l</mi> <mi>i</mi> <mi>o</mi> </mrow> </msub> <mo>&lt;</mo> <mn>0.077</mn> </mrow> </semantics></math> for 57 SNe Ia Hubble flow galaxies shown in Table 2 of Riess et al. (2022) [<a href="#B7-universe-10-00400" class="html-bibr">7</a>,<a href="#B21-universe-10-00400" class="html-bibr">21</a>,<a href="#B22-universe-10-00400" class="html-bibr">22</a>,<a href="#B23-universe-10-00400" class="html-bibr">23</a>,<a href="#B24-universe-10-00400" class="html-bibr">24</a>,<a href="#B25-universe-10-00400" class="html-bibr">25</a>,<a href="#B26-universe-10-00400" class="html-bibr">26</a>]. The RWML model fit was performed using <math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>67.4</mn> </mrow> </semantics></math>km s<sup>−1</sup> Mpc<sup>−1</sup>.</p>
Full article ">
37 pages, 4154 KiB  
Article
Stochastic Optimal Control Analysis for HBV Epidemic Model with Vaccination
by Sayed Murad Ali Shah, Yufeng Nie, Anwarud Din and Abdulwasea Alkhazzan
Symmetry 2024, 16(10), 1306; https://doi.org/10.3390/sym16101306 - 3 Oct 2024
Cited by 1 | Viewed by 553
Abstract
In this study, we explore the concept of symmetry as it applies to the dynamics of the Hepatitis B Virus (HBV) epidemic model. By incorporating symmetric principles in the stochastic model, we ensure that the control strategies derived are not only effective but [...] Read more.
In this study, we explore the concept of symmetry as it applies to the dynamics of the Hepatitis B Virus (HBV) epidemic model. By incorporating symmetric principles in the stochastic model, we ensure that the control strategies derived are not only effective but also consistent across varying conditions, and ensure the reliability of our predictions. This paper presents a stochastic optimal control analysis of an HBV epidemic model, incorporating vaccination as a pivotal control measure. We formulate a stochastic model to capture the complex dynamics of HBV transmission and its progression to acute and chronic stages. By leveraging stochastic differential equations, we examine the model’s stationary distribution and asymptotic behavior, elucidating the impact of random perturbations on disease dynamics. Optimal control theory is employed to derive control strategies aimed at minimizing the disease burden and vaccination costs. Through rigorous numerical simulations using the fourth-order Runge–Kutta method, we demonstrate the efficacy of the proposed control measures. Our findings highlight the critical role of vaccination in controlling HBV spread and provide insights into the optimization of vaccination strategies under stochastic conditions. The symmetry within the proposed model equations allows for a balanced approach to analyzing both acute and chronic stages of HBV. Full article
(This article belongs to the Section Mathematics)
Show Figures

Figure 1

Figure 1
<p>Flowcharts of models (<a href="#FD1-symmetry-16-01306" class="html-disp-formula">1</a>) and (<a href="#FD2-symmetry-16-01306" class="html-disp-formula">2</a>) showing HBV transmission rate. (<b>a</b>) Model (<a href="#FD1-symmetry-16-01306" class="html-disp-formula">1</a>) flowchart. (<b>b</b>) Model (<a href="#FD2-symmetry-16-01306" class="html-disp-formula">2</a>) flowchart.</p>
Full article ">Figure 2
<p>Theses graphs show the paths of deterministic and stochastic models (<a href="#FD1-symmetry-16-01306" class="html-disp-formula">1</a>) and (<a href="#FD2-symmetry-16-01306" class="html-disp-formula">2</a>) when <math display="inline"><semantics> <msubsup> <mi>R</mi> <mn>0</mn> <mi>E</mi> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mi>R</mi> <mrow> <mn>0</mn> </mrow> <mi>D</mi> </msubsup> </semantics></math> are less than one.</p>
Full article ">Figure 3
<p>Tracking trajectories of susceptible, vaccinated, acute infections, chronic carriers, and recovered individuals for models (<a href="#FD1-symmetry-16-01306" class="html-disp-formula">1</a>) and (<a href="#FD2-symmetry-16-01306" class="html-disp-formula">2</a>).</p>
Full article ">Figure 4
<p>Ergodic stationary distribution of model (<a href="#FD1-symmetry-16-01306" class="html-disp-formula">1</a>).</p>
Full article ">Figure 5
<p>The plot visually depicts the temporal evolution of populations for <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>, utilizing both deterministic and stochastic models.</p>
Full article ">Figure 6
<p>The trajectories of <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">A</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Z</mi> <mo>(</mo> <mi mathvariant="sans-serif">t</mi> <mo>)</mo> </mrow> </semantics></math> projected by the stochastic model and their corresponding deterministic counterpart are depicted.</p>
Full article ">Figure 7
<p>Simulated susceptible, vaccinated, and acutely infected populations for both deterministic and stochastic models.</p>
Full article ">Figure 8
<p>Simulated chronic, hospitalized, and recovered populations for both deterministic and stochastic models.</p>
Full article ">Figure 9
<p>These graphs show an optimal control of deterministic and stochastics systems with and without control.</p>
Full article ">
18 pages, 522 KiB  
Article
Uncovering Hidden Patterns: Approximate Resurgent Resummation from Truncated Series
by Alessio Maiezza and Juan Carlos Vasquez
Mathematics 2024, 12(19), 3087; https://doi.org/10.3390/math12193087 - 2 Oct 2024
Viewed by 475
Abstract
We analyze truncated series generated as divergent formal solutions of non-linear ordinary differential equations. Motivating the study is a specific non-linear, first-order differential equation, which is the basis of the resurgent formulation of renormalized perturbation theory in quantum field theory. We use the [...] Read more.
We analyze truncated series generated as divergent formal solutions of non-linear ordinary differential equations. Motivating the study is a specific non-linear, first-order differential equation, which is the basis of the resurgent formulation of renormalized perturbation theory in quantum field theory. We use the Borel–Padé approximant and classical analysis to determine the analytic structure of the solution using the first few terms of its asymptotic series. Afterward, we build an approximant, consistent with the resurgent properties of the equation. The procedure gives an approximate expression for the Borel–Ecalle resummation of the solution useful for practical applications. Connections with other physical applications are also discussed. Full article
(This article belongs to the Section Mathematical Physics)
Show Figures

Figure 1

Figure 1
<p>(<b>Left</b>): accumulation of poles of the Borel–Padé approximants for the asymptotic series related to (<a href="#FD38-mathematics-12-03087" class="html-disp-formula">38</a>); the inset shows the same but with finer bins, zooming on the dominant peak. (<b>Right</b>): accumulation of the Borel–Padé approximants poles of the derivative-of-log applied to the truncated Borel series from (<a href="#FD39-mathematics-12-03087" class="html-disp-formula">39</a>), as in (<a href="#FD37-mathematics-12-03087" class="html-disp-formula">37</a>); the inset shows the same but with finer bins, zooming on the dominant peak.</p>
Full article ">Figure 2
<p>Padé-based estimate of the root branch point (<math display="inline"><semantics> <mrow> <mo>|</mo> <mi>b</mi> <mo>|</mo> </mrow> </semantics></math> from (<a href="#FD40-mathematics-12-03087" class="html-disp-formula">40</a>)) coming from the Borel transform of the truncated series from (<a href="#FD39-mathematics-12-03087" class="html-disp-formula">39</a>).</p>
Full article ">Figure 3
<p>(<b>Left</b>) panel: goodness of the approximant in (<a href="#FD45-mathematics-12-03087" class="html-disp-formula">45</a>) in terms of the ratios of the predicted and actual coefficients of the Borel transform. (<b>Right</b>) panel: same thing but referring to the approximant in (<a href="#FD48-mathematics-12-03087" class="html-disp-formula">48</a>) with <math display="inline"><semantics> <mrow> <msup> <mi>N</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>8</mn> <mi>N</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Ratios of the predicted and actual coefficients of the Borel transform for a varying <math display="inline"><semantics> <msup> <mi>N</mi> <mo>′</mo> </msup> </semantics></math>. The first ratio is not present now since it is used to evaluate <math display="inline"><semantics> <msup> <mi>N</mi> <mo>′</mo> </msup> </semantics></math>. The predicted coefficients have a better convergence than the one in <a href="#mathematics-12-03087-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>Comparison between the exact solution of (<a href="#FD38-mathematics-12-03087" class="html-disp-formula">38</a>), approximate resurgent resummation, and the Cauchy principal value of <span class="html-italic">P</span> (denoted as <math display="inline"><semantics> <msub> <mi>y</mi> <mn>0</mn> </msub> </semantics></math>) in (<a href="#FD45-mathematics-12-03087" class="html-disp-formula">45</a>).</p>
Full article ">Figure 6
<p>(<b>Left</b>) panel is the equivalent of <a href="#mathematics-12-03087-f005" class="html-fig">Figure 5</a>, but one cannot appreciate the difference between the actual result (black) and the resummation (blue); the (<b>Right</b>) panel is magnified to appreciate this difference.</p>
Full article ">Figure 7
<p>(<b>Left</b>) panel shows the exact result, the approximated resurgent resummation, and <math display="inline"><semantics> <msub> <mi>y</mi> <mn>0</mn> </msub> </semantics></math> (the leading contribution, <math display="inline"><semantics> <msup> <mi>C</mi> <mn>0</mn> </msup> </semantics></math>, in (<a href="#FD31-mathematics-12-03087" class="html-disp-formula">31</a>)) Again, one cannot appreciate the difference between the actual result (black) and the resummation (blue); the (<b>Right</b>) panel is a zoom to see this difference.</p>
Full article ">
21 pages, 21669 KiB  
Article
Dynamics of a Predator–Prey System with Impulsive Stocking Prey and Nonlinear Harvesting Predator at Different Moments
by Zeli Zhou, Jianjun Jiao, Xiangjun Dai and Lin Wu
Mathematics 2024, 12(15), 2369; https://doi.org/10.3390/math12152369 - 30 Jul 2024
Viewed by 751
Abstract
In this article, we study a predator–prey system, which includes impulsive stocking prey and a nonlinear harvesting predator at different moments. Firstly, we derive a sufficient condition of the global asymptotical stability of the predator–extinction periodic solution utilizing the comparison theorem of the [...] Read more.
In this article, we study a predator–prey system, which includes impulsive stocking prey and a nonlinear harvesting predator at different moments. Firstly, we derive a sufficient condition of the global asymptotical stability of the predator–extinction periodic solution utilizing the comparison theorem of the impulsive differential equations and the Floquet theory. Secondly, the condition, which is to maintain the permanence of the system, is derived. Finally, some numerical simulations are displayed to examine our theoretical results and research the effect of several important parameters for the investigated system, which shows that the period of the impulse control and impulsive perturbations of the stocking prey and nonlinear harvesting predator have a significant impact on the behavioral dynamics of the system. The results of this paper give a reliable tactical basis for actual biological resource management. Full article
Show Figures

Figure 1

Figure 1
<p>Globally asymptotically stable periodic solution <math display="inline"><semantics> <mrow> <mo>(</mo> <mover> <mrow> <msub> <mi>x</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mi>v</mi> <mo>)</mo> </mrow> </mrow> <mo>¯</mo> </mover> <mo>,</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math> of system (<a href="#FD2-mathematics-12-02369" class="html-disp-formula">2</a>) setting parameters as <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>3</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>K</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>5</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.6</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>0.461</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.68</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <mn>0.7</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>5</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>The permanence of the model (<a href="#FD2-mathematics-12-02369" class="html-disp-formula">2</a>) at <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>3</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>K</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>5</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.5</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.35</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>0.65</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <mn>0.51</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Bifurcation pictures of model (<a href="#FD2-mathematics-12-02369" class="html-disp-formula">2</a>) at <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>0.4</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>0.2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>3</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>K</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>10</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.8</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.15</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>0.41</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.15</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.53</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mn>0.01</mn> <mo>≤</mo> <msub> <mi>h</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mo>≤</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Period-doubling cascade results in system from an <span class="html-italic">F</span>-periodic solution to a <math display="inline"><semantics> <mrow> <mn>2</mn> <mi>F</mi> </mrow> </semantics></math>-periodic solution setting final time <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Bifurcation pictures of system (<a href="#FD2-mathematics-12-02369" class="html-disp-formula">2</a>) on parameters <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>K</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>25</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.8</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.15</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>0.41</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <mn>0.45</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.53</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mn>0.01</mn> <mo>≤</mo> <msub> <mi>μ</mi> <mn>0</mn> </msub> <mo>≤</mo> <mn>9</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Period-halving bifurcation reduces system into a <math display="inline"><semantics> <mrow> <mn>2</mn> <mi>F</mi> </mrow> </semantics></math>-periodic solution with <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Chaos of prey–predator system (<a href="#FD2-mathematics-12-02369" class="html-disp-formula">2</a>) with <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Period-halving bifurcation results in system from a <math display="inline"><semantics> <mrow> <mn>2</mn> <mi>F</mi> </mrow> </semantics></math>-periodic solution into an <span class="html-italic">F</span>-periodic solution with final time <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Bifurcation pictures of model (<a href="#FD2-mathematics-12-02369" class="html-disp-formula">2</a>) at <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>0.6</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>0.4</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>6</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>K</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>25</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.4</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.15</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>=</mo> <mn>0.41</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.4</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <mn>0.6</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.2</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.53</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>0.01</mn> <mo>≤</mo> <mi>F</mi> <mo>≤</mo> <mn>27</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Chaos of prey–predator system (<a href="#FD2-mathematics-12-02369" class="html-disp-formula">2</a>) with <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>=</mo> <mn>2800</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>A period-doubling cascade leads system from an <span class="html-italic">F</span>-periodic to a <math display="inline"><semantics> <mrow> <mn>2</mn> <mi>F</mi> </mrow> </semantics></math>-periodic solution.</p>
Full article ">Figure 12
<p>A period-halving cascade leads system from chaos to cycle.</p>
Full article ">Figure 13
<p>An <span class="html-italic">F</span>-periodic solution suddenly alters into chaos.</p>
Full article ">Figure 14
<p>An <span class="html-italic">F</span>-periodic solution coexists with a strange attractor for <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>9.8</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>=</mo> <mn>6860</mn> </mrow> </semantics></math>.</p>
Full article ">
23 pages, 944 KiB  
Article
Thermodynamic Stability Theories of Irreversible Processes and the Fourth Law of Thermodynamics
by Vijay M. Tangde, Anil A. Bhalekar and Bjarne Andresen
Entropy 2024, 26(6), 442; https://doi.org/10.3390/e26060442 - 24 May 2024
Viewed by 1340
Abstract
Three approaches for determining the thermodynamic stability of irreversible processes are described in generalized formulations. The simplest is the Gibbs–Duhem theory, specialized to irreversible trajectories, which uses the concept of virtual displacement in the reverse direction. Its only drawback is that even a [...] Read more.
Three approaches for determining the thermodynamic stability of irreversible processes are described in generalized formulations. The simplest is the Gibbs–Duhem theory, specialized to irreversible trajectories, which uses the concept of virtual displacement in the reverse direction. Its only drawback is that even a trajectory leading to an explosion is identified as a thermodynamically stable motion. In the second approach, we use a thermodynamic Lyapunov function and its time rate from the Lyapunov thermodynamic stability theory (LTS, previously known as CTTSIP). In doing so, we demonstrate that the second differential of entropy, a frequently used Lyapunov function, is useful only for investigating the stability of equilibrium states. Nonequilibrium steady states do not qualify. Without using explicit perturbation coordinates, we further identify asymptotic thermodynamic stability and thermodynamic stability under constantly acting disturbances of unperturbed trajectories as well as of nonequilibrium steady states. The third approach is also based on the Lyapunov function from LTS, but here we additionally use the rates of perturbation coordinates, based on the Gibbs relations and without using their explicit expressions, to identify not only asymptotic thermodynamic stability but also thermodynamic stability under constantly acting disturbances. Only those trajectories leading to an infinite rate of entropy production (unstable states) are excluded from this conclusion. Finally, we use these findings to formulate the Fourth Law of thermodynamics based on the thermodynamic stability. It is a comprehensive statement covering all nonequilibrium trajectories, close to as well as far from equilibrium. Unlike previous suggested “fourth laws”, this one meets the same level of generality that is associated with the original zeroth to third laws. The above is illustrated using the Schlögl reaction with its multiple steady states in certain regions of operation. Full article
(This article belongs to the Section Thermodynamics)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of an autonomous system of motion: stable, asymptotically stable, and unstable motions. A perturbation of magnitude <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>(</mo> <msub> <mi>t</mi> <mn>0</mn> </msub> <mo>)</mo> </mrow> </semantics></math> away from the unperturbed trajectory indicated as the origin 0 is effected at time <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <msub> <mi>t</mi> <mn>0</mn> </msub> </mrow> </semantics></math> within a sufficiently small region <math display="inline"><semantics> <mi>δ</mi> </semantics></math>. If the perturbed trajectory remains within the region of magnitude <math display="inline"><semantics> <mi>ε</mi> </semantics></math>, the real trajectory is said to be <span class="html-italic">stable</span>. If the perturbed motion within a short time tends back toward the unperturbed trajectory (the origin), the unperturbed motion is said to be <span class="html-italic">asymptotically stable</span>. And if the perturbed trajectory diverges from the region determined by <math display="inline"><semantics> <mi>ε</mi> </semantics></math>, the motion is said to be <span class="html-italic">unstable</span>.</p>
Full article ">Figure 2
<p>A schematic representation of the Schlögl reaction in a continuously stirred tank reactor (CSTR). The input of <math display="inline"><semantics> <mi mathvariant="normal">A</mi> </semantics></math> and output of <math display="inline"><semantics> <mi mathvariant="normal">B</mi> </semantics></math> are adjusted so that within the CSTR, they maintain constant concentrations <math display="inline"><semantics> <msup> <mi mathvariant="normal">A</mi> <mn>0</mn> </msup> </semantics></math> and <math display="inline"><semantics> <msup> <mi mathvariant="normal">B</mi> <mn>0</mn> </msup> </semantics></math>, respectively. The intermediate component <span class="html-italic">X</span> is neither added nor withdrawn during the process. It adjusts itself as part of the reaction scheme.</p>
Full article ">Figure 3
<p>Steady states of the intermediate X of the Schlögl reaction are shown as the driving force (pump parameter) is increased. The solid parts of the curve are stable steady states, the dashed part is unstable.</p>
Full article ">Figure 4
<p>Evolution of the rate of excess entropy production <math display="inline"><semantics> <msub> <mi mathvariant="script">L</mi> <mi>S</mi> </msub> </semantics></math> (green), its time derivative <math display="inline"><semantics> <mover accent="true"> <msub> <mi mathvariant="script">L</mi> <mi>S</mi> </msub> <mo>˙</mo> </mover> </semantics></math> (red), and its gradient <math display="inline"><semantics> <mfrac> <mrow> <mo>∂</mo> <msub> <mi mathvariant="script">L</mi> <mi>S</mi> </msub> </mrow> <mrow> <mo>∂</mo> <mo>(</mo> <mi>δ</mi> <mi>X</mi> <mo>)</mo> </mrow> </mfrac> </semantics></math> (black) using rate constants <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>k</mi> <mn>2</mn> </msub> <mo>=</mo> <msub> <mi>k</mi> <mn>4</mn> </msub> <mo>=</mo> <mn>1</mn> <mo>,</mo> <msub> <mi>k</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, and fixed concentrations <math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">A</mi> <mn>0</mn> </msup> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">B</mi> <mn>0</mn> </msup> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> in the Schlögl reaction. The left graph is calculated around the first steady state point, a stable one. Here, the rate of excess entropy production <math display="inline"><semantics> <msub> <mi mathvariant="script">L</mi> <mi>S</mi> </msub> </semantics></math> is positive and its time rate is of the opposite sign throughout, both converging to zero. Thus, it is a case of <span class="html-italic">asymptotic thermodynamic stability</span>. Further, the gradient of <math display="inline"><semantics> <msub> <mi mathvariant="script">L</mi> <mi>S</mi> </msub> </semantics></math> remains finite throughout, making it also <span class="html-italic">thermodynamically stable under constantly acting disturbance</span>. The right graph is calculated around the second steady-state point, displaying divergent behavior and hence instability.</p>
Full article ">
17 pages, 779 KiB  
Article
Effects of Small Random Perturbations in the Extended Glass–Kauffman Model of Gene Regulatory Networks
by Arcady Ponosov, Irina Shlykova and Ramazan I. Kadiev
Mathematics 2024, 12(8), 1223; https://doi.org/10.3390/math12081223 - 18 Apr 2024
Viewed by 741
Abstract
A mathematical justification of some basic structural properties of stochastically perturbed gene regulatory networks, including those with autoregulation and delay, is offered in this paper. By using the theory of stochastic differential equations, it is, in particular, shown how to control the asymptotic [...] Read more.
A mathematical justification of some basic structural properties of stochastically perturbed gene regulatory networks, including those with autoregulation and delay, is offered in this paper. By using the theory of stochastic differential equations, it is, in particular, shown how to control the asymptotic behavior of the diffusion terms in order to not destroy certain qualitative features of the networks, for instance, their sliding modes. The results also confirm that the level of randomness is gradually reduced if the gene activation times become much smaller than the time of interaction of genes. Finally, the suggested analysis explains why the deterministic numerical schemes based on replacing smooth, steep response functions by the simpler yet discontinuous Heaviside function, the well-known simplification algorithm, are robust with respect to uncertainties in data. The main technical difficulties of the analysis are handled by applying the uniform version of the stochastic Tikhonov theorem in singular perturbation analysis suggested by Yu. Kabanov and S. Pergamentshchikov. Full article
(This article belongs to the Special Issue Nonlinear Stochastic Dynamics and Control and Its Applications)
Show Figures

Figure 1

Figure 1
<p>The Hill function <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mi>i</mi> </msub> <mo>=</mo> <mo>Σ</mo> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mi>i</mi> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mi>i</mi> </msub> <mo>,</mo> <msub> <mi>q</mi> <mi>i</mi> </msub> <mo>)</mo> </mrow> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mi>i</mi> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>) becomes the shifted Heaviside function <math display="inline"><semantics> <mrow> <mo>Σ</mo> <mo>(</mo> <msub> <mi>x</mi> <mi>i</mi> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mi>i</mi> </msub> <mo>,</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math> as <math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mi>i</mi> </msub> <mo>→</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Some phase trajectories of System (<a href="#FD4-mathematics-12-01223" class="html-disp-formula">4</a>) with <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.7</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.6</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>q</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. The trajectories seem to ‘stop’ at the threshold line <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>&gt;</mo> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> <mo>.</mo> </mrow> </semantics></math></p>
Full article ">Figure 3
<p>Some phase trajectories of System (<a href="#FD4-mathematics-12-01223" class="html-disp-formula">4</a>) with <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.7</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.6</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>q</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math>. The ‘true’ trajectories do not stop, but slide along the threshold line <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>&gt;</mo> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (Pplane8 [<a href="#B21-mathematics-12-01223" class="html-bibr">21</a>], MATLAB R2018b).</p>
Full article ">
16 pages, 622 KiB  
Article
Casimir Energy in (2 + 1)-Dimensional Field Theories
by Manuel Asorey, Claudio Iuliano and Fernando Ezquerro
Physics 2024, 6(2), 613-628; https://doi.org/10.3390/physics6020040 - 17 Apr 2024
Cited by 2 | Viewed by 874
Abstract
We explore the dependence of vacuum energy on the boundary conditions for massive scalar fields in (2 + 1)-dimensional spacetimes. We consider the simplest geometrical setup given by a two-dimensional space bounded by two homogeneous parallel wires in order to compare it with [...] Read more.
We explore the dependence of vacuum energy on the boundary conditions for massive scalar fields in (2 + 1)-dimensional spacetimes. We consider the simplest geometrical setup given by a two-dimensional space bounded by two homogeneous parallel wires in order to compare it with the non-perturbative behaviour of the Casimir energy for non-Abelian gauge theories in (2 + 1) dimensions. Our results show the existence of two types of boundary conditions which give rise to two different asymptotic exponential decay regimes of the Casimir energy at large distances. The two families are distinguished by the feature that the boundary conditions involve or not interrelations between the behaviour of the fields at the two boundaries. Non-perturbative numerical simulations and analytical arguments show such an exponential decay for Dirichlet boundary conditions of SU(2) gauge theories. The verification that this behaviour is modified for other types of boundary conditions requires further numerical work. Subdominant corrections in the low-temperature regime are very relevant for numerical simulations, and they are also analysed in this paper. Full article
(This article belongs to the Special Issue 75 Years of the Casimir Effect: Advances and Prospects)
Show Figures

Figure 1

Figure 1
<p>Dependence of the Casimir energy in logarithmic scale as a function of the effective distance <math display="inline"><semantics> <mrow> <mi>m</mi> <mi>L</mi> </mrow> </semantics></math> between the two boundary wires for different boundary conditions.</p>
Full article ">Figure 2
<p>Free energy behaviour of the temperature-dependent part in logarithmic scale as a function of the effective distance <math display="inline"><semantics> <mrow> <mi>m</mi> <mi>L</mi> </mrow> </semantics></math> between the two boundary wires for different boundary conditions, with <math display="inline"><semantics> <mrow> <mi>m</mi> <mi>β</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">
14 pages, 285 KiB  
Article
A Class of Fractional Viscoelastic Kirchhoff Equations Involving Two Nonlinear Source Terms of Different Signs
by Li Zhang and Yang Liu
Axioms 2024, 13(3), 169; https://doi.org/10.3390/axioms13030169 - 5 Mar 2024
Viewed by 1008
Abstract
A class of fractional viscoelastic Kirchhoff equations involving two nonlinear source terms of different signs are studied. Under suitable assumptions on the exponents of nonlinear source terms and the memory kernel, the existence of global solutions in an appropriate functional space is established [...] Read more.
A class of fractional viscoelastic Kirchhoff equations involving two nonlinear source terms of different signs are studied. Under suitable assumptions on the exponents of nonlinear source terms and the memory kernel, the existence of global solutions in an appropriate functional space is established by a combination of the theory of potential wells and the Galerkin approximations. Furthermore, the asymptotic behavior of global solutions is obtained by a combination of the theory of potential wells and the perturbed energy method. Full article
(This article belongs to the Section Mathematical Analysis)
9 pages, 239 KiB  
Article
Quantum Scalar Fields Interacting with Quantum Black Hole Asymptotic Regions
by Rodolfo Gambini and Jorge Pullin
Universe 2024, 10(2), 77; https://doi.org/10.3390/universe10020077 - 6 Feb 2024
Viewed by 1298
Abstract
We continue our work on the study of spherically symmetric loop quantum gravity coupled to two spherically symmetric scalar fields, with one that acts as a clock. As a consequence of the presence of the latter, we can define a true Hamiltonian for [...] Read more.
We continue our work on the study of spherically symmetric loop quantum gravity coupled to two spherically symmetric scalar fields, with one that acts as a clock. As a consequence of the presence of the latter, we can define a true Hamiltonian for the theory. In previous papers, we studied the theory for large values of the radial coordinate, i.e., far away from any black hole or star that may be present. This makes the calculations considerably more tractable. We have shown that in the asymptotic region, the theory admits a large family of quantum vacua for quantum matter fields coupled to quantum gravity, as is expected from the well-known results of quantum field theory on classical curved space-time. Here, we study perturbative corrections involving terms that we neglected in our previous work. Using the time-dependent perturbation theory, we show that the states that represent different possible vacua are essentially stable. This ensures that one recovers from a totally quantized gravitational theory coupled to matter the standard behavior of a matter quantum field theory plus low probability transitions due to gravity between particles that differ at most by a small amount of energy. Full article
(This article belongs to the Section Field Theory)
31 pages, 434 KiB  
Article
Optimized Self-Similar Borel Summation
by Simon Gluzman and Vyacheslav I. Yukalov
Axioms 2023, 12(11), 1060; https://doi.org/10.3390/axioms12111060 - 20 Nov 2023
Cited by 2 | Viewed by 1804
Abstract
The method of Fractional Borel Summation is suggested in conjunction with self-similar factor approximants. The method used for extrapolating asymptotic expansions at small variables to large variables, including the variables tending to infinity, is described. The method is based on the combination of [...] Read more.
The method of Fractional Borel Summation is suggested in conjunction with self-similar factor approximants. The method used for extrapolating asymptotic expansions at small variables to large variables, including the variables tending to infinity, is described. The method is based on the combination of optimized perturbation theory, self-similar approximation theory, and Borel-type transformations. General Borel Fractional transformation of the original series is employed. The transformed series is resummed in order to adhere to the asymptotic power laws. The starting point is the formulation of dynamics in the approximations space by employing the notion of self-similarity. The flow in the approximation space is controlled, and “deep” control is incorporated into the definitions of the self-similar approximants. The class of self-similar approximations, satisfying, by design, the power law behavior, such as the use of self-similar factor approximants, is chosen for the reasons of transparency, explicitness, and convenience. A detailed comparison of different methods is performed on a rather large set of examples, employing self-similar factor approximants, self-similar iterated root approximants, as well as the approximation technique of self-similarly modified Padé–Borel approximations. Full article
(This article belongs to the Special Issue Mathematical Analysis and Applications IV)
41 pages, 2342 KiB  
Article
To the Theory of Decaying Turbulence
by Alexander Migdal
Fractal Fract. 2023, 7(10), 754; https://doi.org/10.3390/fractalfract7100754 - 12 Oct 2023
Cited by 1 | Viewed by 1765
Abstract
We have found an infinite dimensional manifold of exact solutions of the Navier-Stokes loop equation for the Wilson loop in decaying Turbulence in arbitrary dimension d>2. This solution family is equivalent to a fractal curve in complex space Cd [...] Read more.
We have found an infinite dimensional manifold of exact solutions of the Navier-Stokes loop equation for the Wilson loop in decaying Turbulence in arbitrary dimension d>2. This solution family is equivalent to a fractal curve in complex space Cd with random steps parametrized by N Ising variables σi=±1, in addition to a rational number pq and an integer winding number r, related by σi=qr. This equivalence provides a dual theory describing a strong turbulent phase of the Navier-Stokes flow in Rd space as a random geometry in a different space, like ADS/CFT correspondence in gauge theory. From a mathematical point of view, this theory implements a stochastic solution of the unforced Navier-Stokes equations. For a theoretical physicist, this is a quantum statistical system with integer-valued parameters, satisfying some number theory constraints. Its long-range interaction leads to critical phenomena when its size N or its chemical potential μ0. The system with fixed N has different asymptotics at odd and even N, but the limit μ0 is well defined. The energy dissipation rate is analytically calculated as a function of μ using methods of number theory. It grows as ν/μ2 in the continuum limit μ0, leading to anomalous dissipation at μν0. The same method is used to compute all the local vorticity distribution, which has no continuum limit but is renormalizable in the sense that infinities can be absorbed into the redefinition of the parameters. The small perturbation of the fixed manifold satisfies the linear equation we solved in a general form. This perturbation decays as tλ, with a continuous spectrum of indexes λ in the local limit μ0. The spectrum is determined by a resolvent, which is represented as an infinite product of 33 matrices depending of the element of the Euler ensemble. Full article
(This article belongs to the Special Issue Feature Papers for the 'General Mathematics, Analysis' Section)
Show Figures

Figure 1

Figure 1
<p>Log-log plots of four ratios <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>N</mi> </msub> <mrow> <mo>(</mo> <mi>q</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for even/odd <math display="inline"><semantics> <mrow> <mn>2</mn> <mo>&lt;</mo> <mi>q</mi> <mo>&lt;</mo> <mi>N</mi> </mrow> </semantics></math>, even/odd <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>1000</mn> <mo>,</mo> <mn>1001</mn> </mrow> </semantics></math>. The larger <span class="html-italic">N</span> led to astronomically small ratios, so we did not use them.</p>
Full article ">Figure 2
<p>Partition function <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>(</mo> <mi>N</mi> <mo>,</mo> <mi>N</mi> <mo>)</mo> </mrow> </semantics></math> for odd <span class="html-italic">N</span>, fitted as <math display="inline"><semantics> <mrow> <mi>a</mi> <msqrt> <mi>N</mi> </msqrt> <mo form="prefix">log</mo> <msup> <mrow> <mo>(</mo> <mi>N</mi> <mo>)</mo> </mrow> <mi>b</mi> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Backtracking wires corresponding to vorticity correlation function in (<a href="#FD110-fractalfract-07-00754" class="html-disp-formula">110</a>). With these backtracking wires, the correlation function reduces to the closed loop functional, which is represented by our solution with fractal momentum loop <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>P</mi> <mo stretchy="false">→</mo> </mover> <mrow> <mo>(</mo> <mi>θ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>The relative variance computed in the text in the statistical limit.</p>
Full article ">Figure 5
<p>S(q) for <math display="inline"><semantics> <mrow> <mn>10</mn> <mo>&lt;</mo> <mi>q</mi> <mo>&lt;</mo> </mrow> </semantics></math> 100,000. It does not reach any smooth limit; several bands persist up to infinity, similar to the Euler totient <math display="inline"><semantics> <mrow> <msub> <mi>φ</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mi>q</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>The direct computation of the odd <span class="html-italic">N</span> contribution <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>E</mi> </mrow> </semantics></math> to the enstrophy with 20 digits fitted as <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>E</mi> <mo>≈</mo> <mo>−</mo> <msup> <mi>e</mi> <mi>a</mi> </msup> <msup> <mi>N</mi> <mi>b</mi> </msup> <msup> <mo form="prefix">log</mo> <mi>c</mi> </msup> <mi>N</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>The scaling function <math display="inline"><semantics> <mrow> <mi>F</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> <mo>,</mo> <mi>z</mi> <mo>=</mo> <mfrac> <msup> <mover accent="true"> <mi>J</mi> <mo stretchy="false">→</mo> </mover> <mn>2</mn> </msup> <mrow> <msup> <mrow> <mo>(</mo> <mi>t</mi> <mo>+</mo> <msub> <mi>t</mi> <mn>0</mn> </msub> <mo>)</mo> </mrow> <mn>2</mn> </msup> <msup> <mi>μ</mi> <mn>3</mn> </msup> </mrow> </mfrac> </mrow> </semantics></math>, truncated at <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>200</mn> </mrow> </semantics></math>.</p>
Full article ">
31 pages, 660 KiB  
Article
Machine Learning-Based Model Predictive Control of Two-Time-Scale Systems
by Aisha Alnajdi, Fahim Abdullah, Atharva Suryavanshi and Panagiotis D. Christofides
Mathematics 2023, 11(18), 3827; https://doi.org/10.3390/math11183827 - 6 Sep 2023
Viewed by 1484
Abstract
In this study, we present a general form of nonlinear two-time-scale systems, where singular perturbation analysis is used to separate the dynamics of the slow and fast subsystems. Machine learning techniques are utilized to approximate the dynamics of both subsystems. Specifically, a recurrent [...] Read more.
In this study, we present a general form of nonlinear two-time-scale systems, where singular perturbation analysis is used to separate the dynamics of the slow and fast subsystems. Machine learning techniques are utilized to approximate the dynamics of both subsystems. Specifically, a recurrent neural network (RNN) and a feedforward neural network (FNN) are used to predict the slow and fast state vectors, respectively. Moreover, we investigate the generalization error bounds for these machine learning models approximating the dynamics of two-time-scale systems. Next, under the assumption that the fast states are asymptotically stable, our focus shifts toward designing a Lyapunov-based model predictive control (LMPC) scheme that exclusively employs the RNN to predict the dynamics of the slow states. Additionally, we derive sufficient conditions to guarantee the closed-loop stability of the system under the sample-and-hold implementation of the controller. A nonlinear chemical process example is used to demonstrate the theory. In particular, two RNN models are constructed: one to model the full two-time-scale system and the other to predict solely the slow state vector. Both models are integrated within the LMPC scheme, and we compare their closed-loop performance while assessing the computational time required to execute the LMPC optimization problem. Full article
(This article belongs to the Topic AI and Data-Driven Advancements in Industry 4.0)
Show Figures

Figure 1

Figure 1
<p>Recurrent neural network structure.</p>
Full article ">Figure 2
<p>Feedforward neural network structure.</p>
Full article ">Figure 3
<p>The continuous-stirred tank reactor with jacket.</p>
Full article ">Figure 4
<p>States and input trajectories of the CSTR under the Lyapunov-based MPC scheme using the first-principles model of the full process (<math display="inline"><semantics> <mrow> <mi>F</mi> <msub> <mi>P</mi> <mi>F</mi> </msub> </mrow> </semantics></math>-based LMPC, blue line) and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>F</mi> </msub> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>F</mi> </msub> </mrow> </semantics></math>-based LMPC, red dashed line).</p>
Full article ">Figure 5
<p>States and input trajectories of the CSTR under the Lyapunov-based MPC scheme using the first-principles model of the slow-subsystem (<math display="inline"><semantics> <mrow> <mi>F</mi> <msub> <mi>P</mi> <mi>S</mi> </msub> </mrow> </semantics></math>-based LMPC, blue line) and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>S</mi> </msub> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>S</mi> </msub> </mrow> </semantics></math>-based LMPC, red dashed line).</p>
Full article ">Figure 6
<p>Considering the initial condition <math display="inline"><semantics> <mrow> <mi>I</mi> <msub> <mi>C</mi> <mn>3</mn> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mo>−</mo> <mn>3</mn> <mo>,</mo> <mn>30</mn> <mo>,</mo> <mn>5</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, (<b>a</b>) illustrates the time-varying profiles of the states and the input under the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>F</mi> </msub> </mrow> </semantics></math>-based LMPC (solid line), whereas (<b>b</b>) shows the time-varying profiles of the states and the input under the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>S</mi> </msub> </mrow> </semantics></math>-based LMPC (dashed line).</p>
Full article ">Figure 7
<p>Considering the initial condition <math display="inline"><semantics> <mrow> <mi>I</mi> <msub> <mi>C</mi> <mn>4</mn> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mo>−</mo> <mn>2</mn> <mo>,</mo> <mo>−</mo> <mn>10</mn> <mo>,</mo> <mn>100</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, (<b>a</b>) illustrates the time-varying profiles of the states and the input under the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>F</mi> </msub> </mrow> </semantics></math>-based LMPC (solid line), whereas (<b>b</b>) shows the time-varying profiles of the states and the input under the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>S</mi> </msub> </mrow> </semantics></math>-based LMPC (dashed line).</p>
Full article ">Figure 8
<p>Considering the initial condition <math display="inline"><semantics> <mrow> <mi>I</mi> <msub> <mi>C</mi> <mn>5</mn> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mn>1</mn> <mo>,</mo> <mn>90</mn> <mo>,</mo> <mn>10</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, (<b>a</b>) illustrates the time-varying profiles of the states and the input under the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>F</mi> </msub> </mrow> </semantics></math>-based LMPC (solid line), whereas (<b>b</b>) shows the time-varying profiles of the states and the input under the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>S</mi> </msub> </mrow> </semantics></math>-based LMPC (dashed line).</p>
Full article ">Figure 9
<p>Considering the initial condition <math display="inline"><semantics> <mrow> <mi>I</mi> <msub> <mi>C</mi> <mn>10</mn> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mo>−</mo> <mn>1</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>80</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, (<b>a</b>) illustrates the time-varying profiles of the states and the input under the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>F</mi> </msub> </mrow> </semantics></math>-based LMPC (solid line), whereas (<b>b</b>) shows the time-varying profiles of the states and the input under the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>N</mi> <msub> <mi>N</mi> <mi>S</mi> </msub> </mrow> </semantics></math>-based LMPC (dashed line).</p>
Full article ">
6 pages, 1563 KiB  
Proceeding Paper
Investigation of the Influence of Stratospheric Shear on Baroclinic Instability
by Christos Gkoulekas and Nikolaos A. Bakas
Environ. Sci. Proc. 2023, 26(1), 73; https://doi.org/10.3390/environsciproc2023026073 - 25 Aug 2023
Viewed by 832
Abstract
Baroclinic instability is one of the main mechanisms for the formation of synoptic scale systems. Previous studies examined the exponential growth of small perturbations for a stably stratified troposphere in the case of a constant meridional temperature gradient ignoring the stratosphere (Eady’s model). [...] Read more.
Baroclinic instability is one of the main mechanisms for the formation of synoptic scale systems. Previous studies examined the exponential growth of small perturbations for a stably stratified troposphere in the case of a constant meridional temperature gradient ignoring the stratosphere (Eady’s model). However, since stratospheric flow also affects to some extent the motions in the troposphere, in this work we investigate the effect of stratospheric wind shear on baroclinic instability using the tools of Generalized Stability Theory (GST). GST is a linear stability theory that addresses both the exponential growth of perturbations that is pertinent in the large time asymptotic limit and the transient growth of perturbations at finite time. The optimal initial perturbations leading to the largest growth over a specified time interval are calculated for three main cases of stratospheric shear: positive, zero and negative shear over the stratosphere. It is found that the inclusion of stratospheric shear in all three cases decreases perturbation growth and influences the scale of the structures that will dominate the flow. For optimizing times of the order of a week, the development of systems with larger spatial scale compared to the prediction of the Eady model is expected, while for optimizing times of the order of a day, smaller scale systems are found to develop. Full article
Show Figures

Figure 1

Figure 1
<p>Vertical profile of the zonal wind speed <math display="inline"><semantics> <mrow> <mi>U</mi> <mrow> <mo>(</mo> <mi>z</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for the cases of (<b>a</b>) positive, (<b>b</b>) zero and (<b>c</b>) negative stratospheric shear. (<b>d</b>) The vertical profile of the Brunt–Väisälä frequency.</p>
Full article ">Figure 2
<p>Maximum exponential growth rate of perturbations as a function of wavenumber <math display="inline"><semantics> <mi>K</mi> </semantics></math> for perturbations with <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> in the three cases of stratospheric flow and in the case of the Eady model.</p>
Full article ">Figure 3
<p>Optimal energy growth <math display="inline"><semantics> <mi>G</mi> </semantics></math> as a function of the wavenumber <math display="inline"><semantics> <mi>K</mi> </semantics></math> and the orientation <math display="inline"><semantics> <mi>θ</mi> </semantics></math> of the perturbations for <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>o</mi> <mi>p</mi> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> day. Shown is the growth for (<b>a</b>) the Eady model and for (<b>b</b>) positive, (<b>c</b>) zero and (<b>d</b>) negative stratospheric shear.</p>
Full article ">Figure 4
<p>Optimal energy growth <math display="inline"><semantics> <mi>G</mi> </semantics></math> as a function of the wavenumber <math display="inline"><semantics> <mi>K</mi> </semantics></math> and the orientation <math display="inline"><semantics> <mi>θ</mi> </semantics></math> of the perturbations for <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>o</mi> <mi>p</mi> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> days. Shown is the growth for (<b>a</b>) the Eady model and for (<b>b</b>) positive, (<b>c</b>) zero and (<b>d</b>) negative stratospheric shear.</p>
Full article ">
15 pages, 717 KiB  
Article
Quasinormal Modes of a Charged Black Hole with Scalar Hair
by Wen-Di Guo and Qin Tan
Universe 2023, 9(7), 320; https://doi.org/10.3390/universe9070320 - 3 Jul 2023
Cited by 2 | Viewed by 1090
Abstract
Based on the five-dimensional Einstein–Maxwell theory, Bah et al. constructed a singularity-free topology star/black hole [Phys. Rev. Lett. 126, 151101 (2021)]. After performing the Kaluza–Klein reduction, i.e., integrating the extra space dimension, it can obtain an effective four-dimensional spherically static charged black hole [...] Read more.
Based on the five-dimensional Einstein–Maxwell theory, Bah et al. constructed a singularity-free topology star/black hole [Phys. Rev. Lett. 126, 151101 (2021)]. After performing the Kaluza–Klein reduction, i.e., integrating the extra space dimension, it can obtain an effective four-dimensional spherically static charged black hole with scalar hair. In this paper, we study the quasinormal modes (QNMs) of the scalar, electromagnetic, and gravitational fields in the background of this effective four-dimensional charged black hole. The radial parts of the perturbed fields all satisfy a Schrödinger-like equation. Using the asymptotic iteration method, we obtain the QNM frequencies semianalytically. For low-overtone QNMs, the results obtained using both the asymptotic iteration method and the Wentzel–Kramers–Brillouin approximation method agree well. In the null coordinates, the evolution of a Gaussian package is also studied. The QNM frequencies obtained by fitting the evolution data also agree well with the results obtained using the asymptotic iteration method. Full article
(This article belongs to the Section Gravitation)
Show Figures

Figure 1

Figure 1
<p>The effective potentials in the tortoise coordinate <math display="inline"><semantics><msub><mi>r</mi><mo>*</mo></msub></semantics></math>. The parameter <math display="inline"><semantics><msub><mi>r</mi><mi>B</mi></msub></semantics></math> is set to <math display="inline"><semantics><mrow><msub><mi>r</mi><mi>B</mi></msub><mo>=</mo><mn>0.1</mn><msub><mi>r</mi><mi>S</mi></msub></mrow></semantics></math> (black solid lines), <math display="inline"><semantics><mrow><msub><mi>r</mi><mi>B</mi></msub><mo>=</mo><mn>0.5</mn><msub><mi>r</mi><mi>S</mi></msub></mrow></semantics></math> (blue dashed lines), and <math display="inline"><semantics><mrow><msub><mi>r</mi><mi>B</mi></msub><mo>=</mo><mn>0.9</mn><msub><mi>r</mi><mi>S</mi></msub></mrow></semantics></math> (red dashed lines). (<b>a</b>) The scalar field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>0</mn></mrow></semantics></math>. (<b>b</b>) The electromagnetic field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>1</mn></mrow></semantics></math>. (<b>c</b>) The gravitational field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math>.</p>
Full article ">Figure 2
<p>The first twenty QNMs for the charged black hole with scalar hair. The parameter <math display="inline"><semantics><msub><mi>r</mi><mi>B</mi></msub></semantics></math> is set to <math display="inline"><semantics><mrow><msub><mi>r</mi><mi>B</mi></msub><mo>=</mo><mn>0.5</mn><msub><mi>r</mi><mi>S</mi></msub></mrow></semantics></math>. (<b>a</b>) QNMs for the scalar field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>0</mn></mrow></semantics></math> (black dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>1</mn></mrow></semantics></math> (blue dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math> (red dots). (<b>b</b>) QNMs for the electromagnetic field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>1</mn></mrow></semantics></math> (black dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math> (blue dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>3</mn></mrow></semantics></math> (the red dots). (<b>c</b>) QNMs for the gravitational field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math> (black dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>3</mn></mrow></semantics></math> (blue dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>4</mn></mrow></semantics></math> (red dots).</p>
Full article ">Figure 2 Cont.
<p>The first twenty QNMs for the charged black hole with scalar hair. The parameter <math display="inline"><semantics><msub><mi>r</mi><mi>B</mi></msub></semantics></math> is set to <math display="inline"><semantics><mrow><msub><mi>r</mi><mi>B</mi></msub><mo>=</mo><mn>0.5</mn><msub><mi>r</mi><mi>S</mi></msub></mrow></semantics></math>. (<b>a</b>) QNMs for the scalar field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>0</mn></mrow></semantics></math> (black dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>1</mn></mrow></semantics></math> (blue dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math> (red dots). (<b>b</b>) QNMs for the electromagnetic field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>1</mn></mrow></semantics></math> (black dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math> (blue dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>3</mn></mrow></semantics></math> (the red dots). (<b>c</b>) QNMs for the gravitational field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math> (black dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>3</mn></mrow></semantics></math> (blue dots), <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>4</mn></mrow></semantics></math> (red dots).</p>
Full article ">Figure 3
<p>The effect of the parameter <math display="inline"><semantics><msub><mi>r</mi><mi>B</mi></msub></semantics></math> on the frequencies of fundamental QNMs. (<b>a</b>) Real parts of frequencies for the scalar field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>0</mn></mrow></semantics></math>. (<b>b</b>) Imaginary parts of frequencies for the scalar field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>0</mn></mrow></semantics></math>. (<b>c</b>) Real parts of frequencies for the electromagnetic field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>1</mn></mrow></semantics></math>. (<b>d</b>) Imaginary parts of frequencies for the electromagnetic field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>1</mn></mrow></semantics></math>. (<b>e</b>) Real parts of frequencies for the gravitational field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math>. (<b>f</b>) Imaginary parts of frequencies for the gravitational field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math>.</p>
Full article ">Figure 4
<p>Time evolution of the Gauss package extracted at <math display="inline"><semantics><mrow><msub><mi>r</mi><mo>*</mo></msub><mo>=</mo><mn>20</mn></mrow></semantics></math>. The parameter <math display="inline"><semantics><msub><mi>r</mi><mi>B</mi></msub></semantics></math> is set to <math display="inline"><semantics><mrow><msub><mi>r</mi><mi>B</mi></msub><mo>=</mo><mn>0.2</mn><msub><mi>r</mi><mi>S</mi></msub></mrow></semantics></math> (black lines), <math display="inline"><semantics><mrow><msub><mi>r</mi><mi>B</mi></msub><mo>=</mo><mn>0.5</mn><msub><mi>r</mi><mi>S</mi></msub></mrow></semantics></math> (blue lines), and <math display="inline"><semantics><mrow><msub><mi>r</mi><mi>B</mi></msub><mo>=</mo><mn>0.8</mn><msub><mi>r</mi><mi>S</mi></msub></mrow></semantics></math> (red lines). (<b>a</b>) Time evolution of the scalar field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>0</mn></mrow></semantics></math>. (<b>b</b>) Time evolution of the electromagnetic field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>1</mn></mrow></semantics></math>. (<b>c</b>) Time evolution of the gravitational field with <math display="inline"><semantics><mrow><mi>l</mi><mo>=</mo><mn>2</mn></mrow></semantics></math>.</p>
Full article ">
20 pages, 602 KiB  
Article
Quantum-Field Multiloop Calculations in Critical Dynamics
by Ella Ivanova, Georgii Kalagov, Marina Komarova and Mikhail Nalimov
Symmetry 2023, 15(5), 1026; https://doi.org/10.3390/sym15051026 - 6 May 2023
Viewed by 1442
Abstract
The quantum-field renormalization group method is one of the most efficient and powerful tools for studying critical and scaling phenomena in interacting many-particle systems. The multiloop Feynman diagrams underpin the specific implementation of the renormalization group program. In recent years, multiloop computation has [...] Read more.
The quantum-field renormalization group method is one of the most efficient and powerful tools for studying critical and scaling phenomena in interacting many-particle systems. The multiloop Feynman diagrams underpin the specific implementation of the renormalization group program. In recent years, multiloop computation has had a significant breakthrough in both static and dynamic models of critical behavior. In the paper, we focus on the state-of-the-art computational techniques for critical dynamic diagrams and the results obtained with their help. The generic nature of the evaluated physical observables in a wide class of field models is manifested in the asymptotic character of perturbation expansions. Thus, the Borel resummation of series is required to process multiloop results. Such a procedure also enables one to take high-order contributions into consideration properly. The paper outlines the resummation framework in dynamic models and the circumstances in which it can be useful. An important resummation criterion is the properties of the higher-order asymptotics of the perturbation theory. In static theories, these properties are determined by the method of instanton analysis. A similar approach is applicable in critical dynamics models. We describe the calculation of these asymptotics in dynamical models and present the results of the corresponding resummation. Full article
(This article belongs to the Special Issue Review on Quantum Field Theory)
Show Figures

Figure 1

Figure 1
<p>The Schwinger–Keldysh contour in the complex <span class="html-italic">t</span>-plane.</p>
Full article ">Figure 2
<p>The second-order contribution to the four-point Green function. Here, the retarded propagator <math display="inline"><semantics> <mrow> <mo>〈</mo> <mi>ϕ</mi> <mspace width="0.166667em"/> <msup> <mi>ϕ</mi> <mo>′</mo> </msup> <mo>〉</mo> </mrow> </semantics></math> is indicated by a solid arrow, while a plain line depicts the propagator <math display="inline"><semantics> <mrow> <mo>〈</mo> <mi>ϕ</mi> <mspace width="0.166667em"/> <mi>ϕ</mi> <mo>〉</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>The example of two-loop contribution to the four-point Green function.</p>
Full article ">
Back to TopTop