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 (19)

Search Parameters:
Keywords = nonlinear neutral dynamic equation

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
22 pages, 342 KiB  
Article
Hyers–Ulam and Hyers–Ulam–Rassias Stability for a Class of Fractional Evolution Differential Equations with Neutral Time Delay
by Kholoud N. Alharbi
Symmetry 2025, 17(1), 83; https://doi.org/10.3390/sym17010083 - 7 Jan 2025
Abstract
In this paper, we demonstrate that neutral fractional evolution equations with finite delay possess a stable mild solution. Our model incorporates a mixed fractional derivative that combines the Riemann–Liouville and Caputo fractional derivatives with orders 0<α<1 and [...] Read more.
In this paper, we demonstrate that neutral fractional evolution equations with finite delay possess a stable mild solution. Our model incorporates a mixed fractional derivative that combines the Riemann–Liouville and Caputo fractional derivatives with orders 0<α<1 and 1<β<2. We identify the infinitesimal generator of the cosine family and analyze the stability of the mild solution using both Hyers–Ulam–Rassias and Hyers–Ulam stability methodologies, ensuring robust and reliable results for fractional dynamic systems with delay. In order to guarantee that the features of invariance under transformations, such as rotations or reflections, result in the presence of fixed points that remain unchanging and represent the consistency and balance of the underlying system, fixed-point theorems employ the symmetry idea. Lastly, the results obtained are applied to a fractional order nonlinear wave equation with finite delay with respect to time. Full article
16 pages, 14452 KiB  
Article
Disconnected Stationary Solutions in 3D Kolmogorov Flow and Their Relation to Chaotic Dynamics
by Nikolay M. Evstigneev, Taisia V. Karamysheva, Nikolai A. Magnitskii and Oleg I. Ryabkov
Mathematics 2024, 12(21), 3389; https://doi.org/10.3390/math12213389 - 30 Oct 2024
Viewed by 565
Abstract
This paper aims to investigate the nonlinear transition to turbulence in generalized 3D Kolmogorov flow. The difference between this and classical Kolmogorov flow is that the forcing term in the x direction sin(y) is replaced with [...] Read more.
This paper aims to investigate the nonlinear transition to turbulence in generalized 3D Kolmogorov flow. The difference between this and classical Kolmogorov flow is that the forcing term in the x direction sin(y) is replaced with sin(y)cos(z). This drastically complicates the problem. First, a stability analysis is performed by deriving the analog of the Orr–Sommerfeld equation. It is shown that for infinite stretching, the flow is stable, contrary to classical forcing. Next, a neutral curve is constructed, and the stability of the main solution is analyzed. It is shown that for the cubic domain, the main solution is linearly stable, at least for 0<R100. Next, we turn our attention to the numerical investigation of the solutions in the cubic domain. The main feature of this problem is that it is spatially periodic, allowing one to apply a relatively simple pseudo-spectral numerical method for its investigation. We apply the method of deflation to find distinct solutions in the discrete system and the method of arc length continuation to trace the bifurcation solution branches. Such solutions are called disconnected solutions if these are solutions not connected to the branch of the main solution. We investigate the influence of disconnected solutions on the dynamics of the system. It is demonstrated that when disconnected solutions are formed, the nonlinear transition to turbulence is possible, and dangerous initial conditions are these disconnected solutions. Full article
(This article belongs to the Special Issue Applied Mathematics in Nonlinear Dynamics and Chaos)
Show Figures

Figure 1

Figure 1
<p>Neutral curve for the generalized Kolmogorov flow problem for <math display="inline"><semantics> <mrow> <mn>0.1</mn> <mo>≤</mo> <mi>α</mi> <mo>≤</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>2</mn> <mo>≤</mo> <mi>R</mi> <mo>≤</mo> <mn>30</mn> </mrow> </semantics></math>. Numbers near gray curves correspond to different values of wavenumbers <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>x</mi> </msub> <mo>∈</mo> <mrow> <mo>{</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> <mo>}</mo> </mrow> </mrow> </semantics></math>, and the red curve is the hull of all wavenumbers such that the base solution is linearly stable above it and linearly unstable below it (as indicated in the figure).</p>
Full article ">Figure 2
<p>Verification of the bifurcation diagram of the base branch and disconnected branch that was found in [<a href="#B12-mathematics-12-03389" class="html-bibr">12</a>] for the 3D Kolmogorov flow problem with classical forcing and visualization of the branches in physical space by the absolute velocities isosurfaces (blue color represents lower magnitude, red color represents greater magnitude). Red dots represent solutions that were found by the process of deflation, black dots represent continuation of the solution trajectories, and <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> is the <math display="inline"><semantics> <msub> <mi>L</mi> <mn>2</mn> </msub> </semantics></math> norm in the solution in physical space. The saddle-point bifurcation at <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>∼</mo> <mn>55.588</mn> </mrow> </semantics></math> forms a disconnected solution branch.</p>
Full article ">Figure 3
<p>Bifurcation diagram of stationary solutions and visualization of the branches in physical space by the absolute velocities isosurfaces (blue color represents lower magnitude, red color represents greater magnitude). Red dots represent solutions that were found by the process of deflation, black dots represent continuation of the solution trajectories, and <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> is the <math display="inline"><semantics> <msub> <mi>L</mi> <mn>2</mn> </msub> </semantics></math> norm in the solution in physical space. Zoomed area near the maximum parameter value in the neighborhood of the main solution branch.</p>
Full article ">Figure 4
<p>Bifurcation diagram and visualization of chaotic attractors. Attractor trajectories are represented in same norm <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> on the <span class="html-italic">y</span> axis and the value of the solution at the <math display="inline"><semantics> <msub> <mi>L</mi> <mn>1</mn> </msub> </semantics></math> norm on the <span class="html-italic">x</span> axis. Vertical green segments represent the magnitude of the trajectories in <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> norm and arrows represent the position of the illustration on the bifurcation diagram.</p>
Full article ">Figure 5
<p>Bifurcation diagram and visualization of solutions at some points for <math display="inline"><semantics> <mrow> <mn>15.76</mn> <mo>≤</mo> <mi>R</mi> <mo>≤</mo> <mn>20</mn> </mrow> </semantics></math>. Vertical green segments represent the magnitude of the trajectories in <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> norm and arrows represent the position of the solution visualization on the bifurcation diagram.</p>
Full article ">Figure 6
<p>Bifurcation diagram and visualization of the distance between disconnected branches and chaotic trajectory. The distance is measured in the Euclidean norm in the Fourier space by (<a href="#FD30-mathematics-12-03389" class="html-disp-formula">30</a>). Vertical green segments represent the magnitude of the trajectories in <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> norm and arrows represent the position of the visualization on the bifurcation diagram.</p>
Full article ">Figure 7
<p>Bifurcation diagram and visualization of fifteen leading Lupunov exponents for some solutions. If a graph is not provided, then there are more than fifteen positive leading exponents for <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>&gt;</mo> <mn>21</mn> </mrow> </semantics></math>. Vertical green segments represent the magnitude of the trajectories in <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> norm and arrows represent the position of the illustration on the bifurcation diagram. Each subgraph displays evolution of leading Lyapunov exponents as a function of reorthogonalization iterations.</p>
Full article ">Figure 8
<p>Bifurcation diagram and visualization of solutions at some points for <math display="inline"><semantics> <mrow> <mn>21</mn> <mo>≤</mo> <mi>R</mi> <mo>≤</mo> <mn>22</mn> </mrow> </semantics></math>. Vertical green segments represent the magnitude of the trajectories in <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> norm and arrows represent the position of the solution visualization on the bifurcation diagram.</p>
Full article ">Figure 9
<p>Bifurcation diagram and visualization of solutions at some points for <math display="inline"><semantics> <mrow> <mn>22</mn> <mo>≤</mo> <mi>R</mi> <mo>≤</mo> <mn>24</mn> </mrow> </semantics></math>. Vertical green segments represent the magnitude of the trajectories in <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> norm and arrows represent the position of the solution visualization on the bifurcation diagram.</p>
Full article ">Figure 10
<p>Bifurcation diagram and visualization of solutions at some points for <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>=</mo> <mn>26</mn> </mrow> </semantics></math>. Vertical green segments represent the magnitude of the trajectories in <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> norm and arrows represent the position of the solution visualization on the bifurcation diagram.</p>
Full article ">Figure 11
<p>Bifurcation diagram and visualization of solutions at some points for <math display="inline"><semantics> <mrow> <mn>29</mn> <mo>≤</mo> <mi>R</mi> <mo>≤</mo> <mn>30</mn> </mrow> </semantics></math>. Vertical green segments represent the magnitude of the trajectories in <math display="inline"><semantics> <mrow> <mo>∥</mo> <mi>u</mi> <mo>∥</mo> </mrow> </semantics></math> norm and arrows represent the position of the solution visualization on the bifurcation diagram.</p>
Full article ">
21 pages, 1940 KiB  
Article
Oscillation Criteria for Nonlinear Third-Order Delay Dynamic Equations on Time Scales Involving a Super-Linear Neutral Term
by Qinghua Feng and Bin Zheng
Fractal Fract. 2024, 8(2), 115; https://doi.org/10.3390/fractalfract8020115 - 14 Feb 2024
Cited by 5 | Viewed by 1304
Abstract
In the sense of an arbitrary time scale, some new sufficient conditions on oscillation are presented in this paper for a class of nonlinear third-order delay dynamic equations involving a local fractional derivative with a super-linear neutral term. The established oscillation results include [...] Read more.
In the sense of an arbitrary time scale, some new sufficient conditions on oscillation are presented in this paper for a class of nonlinear third-order delay dynamic equations involving a local fractional derivative with a super-linear neutral term. The established oscillation results include known Kamenev and Philos-type oscillation criteria and are new oscillation results so far in the literature. Some inequalities, the Riccati transformation, the integral technique, and the theory of time scale are used in the establishment of these oscillation criteria. The proposed results unify continuous and discrete analysis, and the process of deduction is further extended to another class of nonlinear third-order delay dynamic equations involving a local fractional derivative with a super-linear neutral term and a damping term. As applications for the established oscillation criteria, some examples are given. Full article
(This article belongs to the Special Issue Mathematical and Physical Analysis of Fractional Dynamical Systems)
Show Figures

Figure 1

Figure 1
<p>Numerical demonstration of the oscillatory behavior of the solution of (43) without a damping term when <span class="html-italic">x</span> is large enough.</p>
Full article ">Figure 2
<p>Numerical demonstration of the oscillatory behavior of the solutions of (44) with a damping term when <span class="html-italic">x</span> is large enough.</p>
Full article ">
29 pages, 10330 KiB  
Article
Enhanced Integrator with Drift Elimination for Accurate Flux Estimation in Sensorless Controlled Interior PMSM for High-Performance Full Speed Range Hybrid Electric Vehicles Applications
by Sadiq Ur Rahman and Chaoying Xia
Machines 2023, 11(7), 769; https://doi.org/10.3390/machines11070769 - 24 Jul 2023
Viewed by 1930
Abstract
Interior Permanent Magnet Synchronous Motor (IPMSM) motion-sensorless speed control necessitates precise knowledge of rotor flux, speed, and position. Due to numerous non-ideal aspects, such as converter nonlinearities, detection errors, integral initial value, and parameter mismatches, the conventional first-order integrator’s estimated rotor flux experiences [...] Read more.
Interior Permanent Magnet Synchronous Motor (IPMSM) motion-sensorless speed control necessitates precise knowledge of rotor flux, speed, and position. Due to numerous non-ideal aspects, such as converter nonlinearities, detection errors, integral initial value, and parameter mismatches, the conventional first-order integrator’s estimated rotor flux experiences a DC offset (Doff). Low-pass filters (LPF) with a constant cut-off frequency yield accurate estimates only in the medium- and high-speed range; however, at the low-speed area, both magnitude and phase estimates are inaccurate. The presented technique resolves the aforementioned issue for a broad speed range. In order to achieve precise flux estimation, this article presents an improved technique of flux estimator with two distinct drift mitigation strategies for the motion-sensorless field-oriented control (FOC) system of IPMSM. Using the orthogonality of the α- and β-axes, the proposed drift elimination system can estimate drift in different situations while maintaining a high level of dynamic performance. The stator flux linkage (SFL) computation in the synchronous coordinate is established from the estimation of the rotating shaft’s permanent magnetic flux linkage orientation and the statistical equations model of the SFL. By comparing the calculated SFL vector to the SFL vector derived from the stator winding voltage and currents integral model with a drift PI compensation loop, a feedback loop is formed to neutralize integral drift, and the rotational speed and position of an IPMSM is estimated utilizing the vector product of the two flux linkages in a phase-locked loop. Theoretical interpretation is presented, and Matlab Simulink simulations, as well as experimental outcomes, consistently demonstrate that the suggested estimation techniques can eliminate the phenomenon of flux drift. Full article
(This article belongs to the Special Issue Advanced Data Analytics in Intelligent Industry: Theory and Practice)
Show Figures

Figure 1

Figure 1
<p>IPMSM sensorless control using rotor flux estimation.</p>
Full article ">Figure 2
<p>Diagram of IPMSM voltage model for rotor speed and position estimation.</p>
Full article ">Figure 3
<p>Design of flux estimator combined with PLL for the motion-sensorless system of IPMSM: (<b>a</b>) drift elimination strategy 1; (<b>b</b>) drift elimination strategy 2.</p>
Full article ">Figure 4
<p>Schematic diagram of the PI controller’s estimation error for <math display="inline"><semantics><mrow><mover><mi>e</mi><mo stretchy="false">^</mo></mover><msub><mrow/><mrow><mi>d</mi><mi>c</mi></mrow></msub></mrow></semantics></math>.</p>
Full article ">Figure 5
<p>Schematic diagram of the PI controller’s estimation for <math display="inline"><semantics><mrow><mover><mi>e</mi><mo stretchy="false">^</mo></mover><msub><mrow/><mrow><mi>d</mi><mi>c</mi></mrow></msub></mrow></semantics></math>.</p>
Full article ">Figure 6
<p>Schematic representation of overall motion-sensorless control system.</p>
Full article ">Figure 7
<p>Matlab Simulation results with constant speed (300 r/min) under the <math display="inline"><semantics><mrow><mi mathvariant="normal">α</mi></mrow></semantics></math>-axis voltage disturbance (0 V and 0.6 V). (<b>a</b>) Artificially injected Doff and DC-drift estimation error, DM1, and DM2. (<b>b</b>) Speed and position estimation error, DM1, and DM2. (<b>c</b>) Observed speed estimation, DM1, and DM2. (<b>d</b>) Estimation fluxes and the errors of estimated fluxes, DM1, and DM2.</p>
Full article ">Figure 7 Cont.
<p>Matlab Simulation results with constant speed (300 r/min) under the <math display="inline"><semantics><mrow><mi mathvariant="normal">α</mi></mrow></semantics></math>-axis voltage disturbance (0 V and 0.6 V). (<b>a</b>) Artificially injected Doff and DC-drift estimation error, DM1, and DM2. (<b>b</b>) Speed and position estimation error, DM1, and DM2. (<b>c</b>) Observed speed estimation, DM1, and DM2. (<b>d</b>) Estimation fluxes and the errors of estimated fluxes, DM1, and DM2.</p>
Full article ">Figure 8
<p>Simulation results with constant speed (300 r/min) under the <math display="inline"><semantics><mrow><mi>α</mi></mrow></semantics></math>-axis voltage disturbance (1 V and 1.5 V). (<b>a</b>) Artificially injected Doff and DC-drift estimation error, DM1, and DM2. (<b>b</b>) Speed and position estimation error, DM1, and DM2. (<b>c</b>) Observed speed estimation, DM1, and DM2. (<b>d</b>) Estimation fluxes and the errors of estimated fluxes, DM1, and DM2.</p>
Full article ">Figure 8 Cont.
<p>Simulation results with constant speed (300 r/min) under the <math display="inline"><semantics><mrow><mi>α</mi></mrow></semantics></math>-axis voltage disturbance (1 V and 1.5 V). (<b>a</b>) Artificially injected Doff and DC-drift estimation error, DM1, and DM2. (<b>b</b>) Speed and position estimation error, DM1, and DM2. (<b>c</b>) Observed speed estimation, DM1, and DM2. (<b>d</b>) Estimation fluxes and the errors of estimated fluxes, DM1, and DM2.</p>
Full article ">Figure 9
<p>Simulation results with constant speed (300 r/min) under the <math display="inline"><semantics><mrow><mi>α</mi><mi>β</mi></mrow></semantics></math>-axis voltage disturbance. (<b>a</b>) Artificially injected Doff and DC-drift estimation error, DM1, and DM2. (<b>b</b>) Speed and position estimation error, DM1, and DM2. (<b>c</b>) Observed speed estimation, DM1, and DM2. (<b>d</b>) Estimation fluxes and the errors of estimated fluxes, DM1, and DM2.</p>
Full article ">Figure 9 Cont.
<p>Simulation results with constant speed (300 r/min) under the <math display="inline"><semantics><mrow><mi>α</mi><mi>β</mi></mrow></semantics></math>-axis voltage disturbance. (<b>a</b>) Artificially injected Doff and DC-drift estimation error, DM1, and DM2. (<b>b</b>) Speed and position estimation error, DM1, and DM2. (<b>c</b>) Observed speed estimation, DM1, and DM2. (<b>d</b>) Estimation fluxes and the errors of estimated fluxes, DM1, and DM2.</p>
Full article ">Figure 10
<p>Structure diagram of real-time simulation experimental system for IPMSM based on dSPACE.</p>
Full article ">Figure 11
<p>IPMSM experimental platform based on dSPACE.</p>
Full article ">Figure 12
<p>Experimental results of drift elimination strategy 1 with 300 RPM speed under the <math display="inline"><semantics><mrow><mi>α</mi></mrow></semantics></math>-axis voltage disturbance (0 V and 0.6 V). (<b>a</b>) Artificially injected Doff and speed estimation; (<b>b</b>) flux estimation and its enlarged view; (<b>c</b>) angle estimation and its enlarger view; (<b>d</b>) angle and speed estimation errors.</p>
Full article ">Figure 13
<p>Experimental results of drift elimination strategy 2 with 300 RPM speed under the <math display="inline"><semantics><mrow><mi>α</mi></mrow></semantics></math>-axis voltage disturbance (0.6 V and 0 V). (<b>a</b>) Artificially injected Doff and speed estimation; (<b>b</b>) flux estimation and its enlarged view; (<b>c</b>) angle estimation and its enlarger view; (<b>d</b>) angle and speed estimation errors.</p>
Full article ">Figure 14
<p>Experimental results of drift elimination strategy 1 with 300 RPM speed under the <math display="inline"><semantics><mrow><mi>α</mi></mrow></semantics></math>-axis voltage disturbance (1.0 V to 1.5 V). (<b>a</b>) Artificially injected Doff and speed estimation; (<b>b</b>) flux estimation and its enlarged view; (<b>c</b>) angle estimation and its enlarger view; (<b>d</b>) angle and speed estimation errors.</p>
Full article ">Figure 15
<p>Experimental results of drift elimination strategy 2 with 300 RPM speed under the <math display="inline"><semantics><mrow><mi>α</mi></mrow></semantics></math>-axis voltage disturbance. (<b>a</b>) Artificially injected Doff and speed estimation; (<b>b</b>) flux estimation and its enlarged view; (<b>c</b>) angle estimation and its enlarger view; (<b>d</b>) angle and speed estimation errors.</p>
Full article ">Figure 16
<p>Experimental results of drift elimination strategy 1 with 300 RPM speed under the <math display="inline"><semantics><mrow><mi>α</mi><mi>β</mi></mrow></semantics></math>-axis voltage disturbance. (<b>a</b>) Artificially injected Doff and speed estimation; (<b>b</b>) flux estimation and its enlarged view; (<b>c</b>) angle estimation and its enlarger view; (<b>d</b>) angle and speed estimation errors.</p>
Full article ">Figure 17
<p>Experimental results of drift elimination strategy 2 with 300 RPM speed under the <math display="inline"><semantics><mrow><mi>α</mi><mi>β</mi></mrow></semantics></math>-axis voltage disturbance. (<b>a</b>) Artificially injected Doff and speed estimation; (<b>b</b>) flux estimation and its enlarged view; (<b>c</b>) angle estimation and its enlarger view; (<b>d</b>) angle and speed estimation errors.</p>
Full article ">
19 pages, 1578 KiB  
Article
An Improved Method of Model-Free Adaptive Predictive Control: A Case of pH Neutralization in WWTP
by Jufeng Li, Zhihe Tang, Hui Luan, Zhongyao Liu, Baochang Xu, Zhongjun Wang and Wei He
Processes 2023, 11(5), 1448; https://doi.org/10.3390/pr11051448 - 10 May 2023
Cited by 2 | Viewed by 2389
Abstract
pH neutralization reaction process plays a crucial role in Waste Water Treatment Process (WWTP). Traditional PID Proportion Integral Differential, (or even advanced PID control) algorithms have poor performance on WWTP due to the strong non-linearity, large time lag, and large inertia characteristics of [...] Read more.
pH neutralization reaction process plays a crucial role in Waste Water Treatment Process (WWTP). Traditional PID Proportion Integral Differential, (or even advanced PID control) algorithms have poor performance on WWTP due to the strong non-linearity, large time lag, and large inertia characteristics of pH neutralization. Therefore, finding a superior control method to maintain the pH value of wastewater within the normal range will greatly help to improve the efficiency and effectiveness of wastewater treatment. The chemical reaction mechanism of pH neutralization reaction process is first analyzed, and a mechanistic model of pH neutralization reaction process is developed based on the reaction of ions during acid-alkali neutralization and the electric balance equation. Then, combining the characteristics of generalized predictive control and Model-Free Adaptive Control (MFAC), a Model-Free Adaptive Predictive Control (MFAPC) method based on compact format dynamic linearization is introduced. An Improved Model Free Adaptive PI Predictive Control algorithm (IMFAPC) with proportional (P) and integral (I) algorithms is proposed to further improve the control performance. IMFAPC is proposed on the basis of MFAPC, combining the advantages of generalized predictive control, introducing a PI module consisting of error and error sum, and predicting the PI module, making it possible to produce more accurate constraints on the control inputs, avoiding increasing errors, and improving the control effect of delayed systems at the same time. pH neutralization process simulation experimental results show that compared with the ordinary Model-Free Adaptive Control (MFAC) and MFAPC, the IMFAPC control algorithms has the best performance in terms of accuracy, overshoot, and the robustness. Full article
(This article belongs to the Special Issue Industrial Wastewater Treatment)
Show Figures

Figure 1

Figure 1
<p>pH neutralization titration curve.</p>
Full article ">Figure 2
<p>pH neutralization reaction schematic.</p>
Full article ">Figure 3
<p>Object Structure Diagram. <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>(</mo> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math>—output of time lag inertia link; <math display="inline"><semantics> <msub> <mi>T</mi> <mi>p</mi> </msub> </semantics></math>—inertia time; <span class="html-italic">d</span>—time lag steps.</p>
Full article ">Figure 4
<p>Control system structure diagram.</p>
Full article ">Figure 5
<p>The setting value is between 6 and 8.</p>
Full article ">Figure 6
<p>The setting value is 5 and 10.</p>
Full article ">Figure 7
<p>When the lag steps are 3 and 5, respectively.</p>
Full article ">Figure 8
<p>System response when the <math display="inline"><semantics> <msub> <mi>T</mi> <mi>P</mi> </msub> </semantics></math> = 40 min, <span class="html-italic">d</span> = 1, <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math> = 2, <math display="inline"><semantics> <msub> <mi>ω</mi> <mi>k</mi> </msub> </semantics></math> is 0.05 and 0.5.</p>
Full article ">Figure 9
<p>System response when the inertia time of the system (time constant), <math display="inline"><semantics> <msub> <mi>T</mi> <mi>P</mi> </msub> </semantics></math> is 50, 100 min, <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math> = 2, d = 1.</p>
Full article ">Figure 10
<p>The control effect of IMFAPC under different <math display="inline"><semantics> <mi>λ</mi> </semantics></math>.</p>
Full article ">Figure 11
<p>The rest of the conditions are the same as the simulations of IMFAPC with different <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math>.</p>
Full article ">
35 pages, 4688 KiB  
Article
Estimating Mean Wind Profiles Inside Realistic Urban Canopies
by Huanhuan Wang, Eden Furtak-Cole and Keith Ngan
Atmosphere 2023, 14(1), 50; https://doi.org/10.3390/atmos14010050 - 27 Dec 2022
Cited by 7 | Viewed by 2068
Abstract
Mean wind profiles within a unit-aspect-ratio street canyon have been estimated by solving the three-dimensional Poisson equation for a set of discrete vortex sheets. The validity of this approach, which assumes inviscid vortex dynamics away from boundaries and a small nonlinear contribution to [...] Read more.
Mean wind profiles within a unit-aspect-ratio street canyon have been estimated by solving the three-dimensional Poisson equation for a set of discrete vortex sheets. The validity of this approach, which assumes inviscid vortex dynamics away from boundaries and a small nonlinear contribution to the growth of turbulent fluctuations, is tested for a series of idealised and realistic flows. In this paper, the effects of urban geometry on accuracy are examined with neutral flow over shallow, deep, asymmetric and realistic canyons, while thermal effects are investigated for a single street canyon and both bottom cooling and heating. The estimated mean profiles of the streamwise and spanwise velocity components show good agreement with reference profiles obtained from the large-eddy simulation: the canyon-averaged errors (e.g., normalised absolute errors around 1%) are of the same order of magnitude as those for the unit-aspect-ratio street canyon. It is argued that the approach generalises to more realistic flows because strong spatial localisation of the vorticity field is preserved. This work may be applied to high-resolution modelling of winds and pollutants, for which mean wind profiles are required, and fast statistical modelling, for which physically-based estimates can serve as initial guesses or substitutes for analytical models. Full article
(This article belongs to the Special Issue Multiscale Aspects of Mesoscale and Microscale Flows)
Show Figures

Figure 1

Figure 1
<p>Urban geometries and vortex sheets (<b>a</b>) shallow or deep canyon; (<b>b</b>) step-up canyon; (<b>c</b>) step-down canyon; (<b>d</b>) homogeneous neighbourhood (Whampoa, Hong Kong); (<b>e</b>) heterogeneous neighbourhood (Central, Hong Kong). The vortex sheets are indicated in colour: roof level, <math display="inline"><semantics> <msup> <mo>ℜ</mo> <mi>t</mi> </msup> </semantics></math> (violet); ground level, <math display="inline"><semantics> <msup> <mo>ℜ</mo> <mi>b</mi> </msup> </semantics></math> (yellow); intermediate, <math display="inline"><semantics> <msup> <mo>ℜ</mo> <mi>i</mi> </msup> </semantics></math> (red); sidewall, <math display="inline"><semantics> <msup> <mo>ℜ</mo> <mi>s</mi> </msup> </semantics></math> (green). There are several intermediate and sidewall vortex sheets, denoted as <math display="inline"><semantics> <msup> <mo>ℜ</mo> <msub> <mi>i</mi> <mi>k</mi> </msub> </msup> </semantics></math> and <math display="inline"><semantics> <msup> <mo>ℜ</mo> <msub> <mi>s</mi> <mi>k</mi> </msub> </msup> </semantics></math>, for the asymmetric and realistic canyons. Computational domain parameters are described in <a href="#atmosphere-14-00050-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 2
<p>Validation of the IDDES model (OpenFOAM) for neutral flow over a 2-D street canyon. Normalised mean streamwise velocity and TKE validation of IDDES (solid line) against wind tunnel experiment [<a href="#B54-atmosphere-14-00050" class="html-bibr">54</a>] (filled circles) at (<b>a</b>,<b>f</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mo>−</mo> <mn>0.4</mn> </mrow> </semantics></math>; (<b>b</b>,<b>g</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mo>−</mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>c</b>,<b>h</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> </mrow> </semantics></math>; (<b>d</b>,<b>i</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>e</b>,<b>j</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.4</mn> </mrow> </semantics></math>. Reference values for the streamwise velocity and TKE, <math display="inline"><semantics> <mrow> <mo>〈</mo> <mover> <msub> <mi>U</mi> <mi>s</mi> </msub> <mo>¯</mo> </mover> <mo>〉</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>〈</mo> <mover> <mrow> <mi>T</mi> <mi>K</mi> <msub> <mi>E</mi> <mi>s</mi> </msub> </mrow> <mo>¯</mo> </mover> <mo>〉</mo> </mrow> </semantics></math>, represent averages of the IDDES data over the shear layer, <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mo>≤</mo> <mo> </mo> <mi>z</mi> <mo>/</mo> <mi>H</mi> <mo> </mo> <mo>≤</mo> <mo> </mo> <mn>1.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Validation of the LES model for stratified flow over a regular array against the wind-tunnel data of Uehara et al. [<a href="#B58-atmosphere-14-00050" class="html-bibr">58</a>]: (<b>a</b>) streamwise velocity; (<b>b</b>) temperature. The bulk Richardson number Rb is defined in Equation (14).</p>
Full article ">Figure 4
<p>Flow structures in the <span class="html-italic">x</span>–<span class="html-italic">z</span> plane for uniform street canyons with a mean flow perpendicular to the canyon axis (<math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> </mrow> </semantics></math>°): (<b>left</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>middle</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.5</mn> </mrow> </semantics></math>; (<b>right</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math>. (<b>a</b>–<b>c</b>) Temporally and spatially averaged streamlines; vertical velocities are plotted in colour; (<b>d</b>–<b>f</b>) spatially averaged vorticity magnitude normalised by the maximum value, <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mo>|</mo> </mrow> <mover accent="true"> <mi>ω</mi> <mo stretchy="false">→</mo> </mover> <msup> <mrow> <mo>|</mo> <mo>|</mo> </mrow> <mo>*</mo> </msup> </mrow> </semantics></math>; (<b>g</b>–<b>i</b>) vertical profiles of the TKE and <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mo>|</mo> </mrow> <mover accent="true"> <mi>ω</mi> <mo stretchy="false">→</mo> </mover> <msup> <mrow> <mo>|</mo> <mo>|</mo> </mrow> <mo>*</mo> </msup> </mrow> </semantics></math>. The TKE is normalised using <math display="inline"><semantics> <mrow> <mo>〈</mo> <msub> <mover accent="true"> <mi>U</mi> <mo stretchy="false">¯</mo> </mover> <mi>ref</mi> </msub> <mo>〉</mo> </mrow> </semantics></math>, the IDDES streamwise velocity at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Vertical profiles of normalised velocities for <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>≤</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> </mrow> </semantics></math>°. Estimated (solid lines) and IDDES (dashed lines) profiles are compared for: (<b>a</b>,<b>d</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>b</b>,<b>e</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.5</mn> </mrow> </semantics></math>; (<b>c</b>,<b>f</b>) <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math>. Streamwise (<b>top</b>) and spanwise (<b>bottom</b>) velocities are normalised by <math display="inline"><semantics> <mrow> <mo stretchy="false">〈</mo> <msub> <mover accent="true"> <mi>U</mi> <mo stretchy="false">¯</mo> </mover> <mi>ref</mi> </msub> <mo stretchy="false">〉</mo> </mrow> </semantics></math>, the IDDES streamwise velocity at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> </mrow> </semantics></math> = 1. Henceforth this definition of <math display="inline"><semantics> <mrow> <mo stretchy="false">〈</mo> <msub> <mover accent="true"> <mi>U</mi> <mo stretchy="false">¯</mo> </mover> <mi>ref</mi> </msub> <mo stretchy="false">〉</mo> </mrow> </semantics></math> is used for all vertical profiles.</p>
Full article ">Figure 6
<p>Flow structures for a deep canyon with <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> </mrow> </semantics></math>°. (<b>a</b>) Temporally and spatially averaged streamlines; (<b>b</b>) spanwise vorticity; (<b>c</b>) vertical profiles of the TKE and spanwise vorticity. As in <a href="#atmosphere-14-00050-f004" class="html-fig">Figure 4</a>, the streamlines and vorticity field are averaged in the spanwise direction.</p>
Full article ">Figure 7
<p>Vertical profiles of normalised velocities for the deep canyon (<math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>3</mn> </mrow> </semantics></math>) and various <math display="inline"><semantics> <mi>θ</mi> </semantics></math>. Mean streamwise (<b>top</b>) and spanwise (<b>bottom</b>) profiles are shown for the vortex method (dotted black line) and IDDES (dashed blue line). (<b>a</b>,<b>f</b>) <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> </mrow> </semantics></math>°; (<b>b</b>,<b>g</b>) <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>30</mn> </mrow> </semantics></math>°; (<b>c</b>,<b>h</b>) <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>45</mn> </mrow> </semantics></math>°; (<b>d</b>,<b>i</b>) <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>60</mn> </mrow> </semantics></math>°; (<b>e</b>,<b>j</b>) <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>90</mn> </mrow> </semantics></math>°. The mean streamwise velocity at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math> is used for the normalisation.</p>
Full article ">Figure 8
<p>Variation of the canyon-averaged relative errors with <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> </mrow> </semantics></math> at (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>45</mn> </mrow> </semantics></math>°; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>90</mn> </mrow> </semantics></math>°.</p>
Full article ">Figure 9
<p>As in <a href="#atmosphere-14-00050-f004" class="html-fig">Figure 4</a> but for asymmetric canyons. (<b>Top</b>) step-up canyon (<math display="inline"><semantics> <mrow> <mi>B</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>2</mn> </mrow> </semantics></math>); (<b>bottom</b>) step-down canyon (<math display="inline"><semantics> <mrow> <mi>B</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.5</mn> </mrow> </semantics></math>). (<b>a</b>,<b>d</b>) Temporally and spatially averaged streamlines; (<b>b</b>,<b>e</b>) spatially averaged vorticity magnitude; (<b>c</b>,<b>f</b>) vertical profiles of the normalised TKE and vorticity magnitude.</p>
Full article ">Figure 10
<p>Vertical profiles of normalised velocities for step-up (<math display="inline"><semantics> <mrow> <mi>B</mi> <mi>R</mi> <mo> </mo> <mo>&gt;</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math>) and step-down (<math display="inline"><semantics> <mrow> <mi>B</mi> <mi>R</mi> <mo> </mo> <mo>&lt;</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math>) canyons and <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>45</mn> </mrow> </semantics></math>°. (<b>Top</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>x</mi> </msub> </semantics></math>; (<b>bottom</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>y</mi> </msub> </semantics></math>. Predictions with (solid red line) and without (dotted black line) the intermediate vortex sheet located on top of the lower building are compared to IDDES (dashed blue line).</p>
Full article ">Figure 11
<p>Time-averaged normalised vorticity magnitude for the homogeneous neighbourhood and an external wind at 45 (indicated by the arrow): (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>–</mo> <mi>y</mi> </mrow> </semantics></math> plane at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>H</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.5</mn> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>–</mo> <mi>z</mi> </mrow> </semantics></math> plane at <math display="inline"><semantics> <mrow> <mi>y</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>300</mn> </mrow> </semantics></math> m (indicated by the white dashed line in panel <b>a</b>).</p>
Full article ">Figure 12
<p>Mean absolute (<math display="inline"><semantics> <msub> <mi>ε</mi> <mi>a</mi> </msub> </semantics></math>) and relative (<math display="inline"><semantics> <msub> <mi>ε</mi> <mi>r</mi> </msub> </semantics></math>) errors against the number of sidewall vortex sheets, <math display="inline"><semantics> <msub> <mi>N</mi> <mi>s</mi> </msub> </semantics></math>, for <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>45</mn> </mrow> </semantics></math>°: (<b>a</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>x</mi> </msub> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>y</mi> </msub> </semantics></math>. The error bars represent the minimum and maximum errors for each ensemble of <math display="inline"><semantics> <msub> <mi>N</mi> <mi>s</mi> </msub> </semantics></math> sidewall vortex sheets. absolute errors are represented by blue lines; relative ones are denoted by red lines.</p>
Full article ">Figure 13
<p>Vertical profiles of mean normalised velocities within the homogeneous domain. (<b>Top</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>x</mi> </msub> </semantics></math> estimated by <math display="inline"><semantics> <msup> <mi mathvariant="normal">Ω</mi> <mrow> <mi>b</mi> <mi>e</mi> <mi>s</mi> <mi>t</mi> <mspace width="0.166667em"/> <mi>x</mi> </mrow> </msup> </semantics></math> (solid red line) and <math display="inline"><semantics> <msup> <mi mathvariant="normal">Ω</mi> <mrow> <mi>h</mi> <mi>o</mi> <mi>m</mi> <mo>,</mo> <mi>b</mi> <mi>a</mi> <mi>s</mi> <mi>e</mi> </mrow> </msup> </semantics></math> (dash-dotted green line); (<b>bottom</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>y</mi> </msub> </semantics></math> estimated by <math display="inline"><semantics> <msup> <mi mathvariant="normal">Ω</mi> <mrow> <mi>b</mi> <mi>e</mi> <mi>s</mi> <mi>t</mi> <mspace width="0.166667em"/> <mi>y</mi> </mrow> </msup> </semantics></math> (solid red line) and a single vortex sheet (<math display="inline"><semantics> <msup> <mi mathvariant="normal">Ω</mi> <mrow> <mi>w</mi> <mi>o</mi> <mi>r</mi> <mi>s</mi> <mi>t</mi> <mspace width="0.166667em"/> <mi>y</mi> </mrow> </msup> </semantics></math>; dash-dotted green line). The IDDES profiles are plotted with a dashed blue line. The reference value, <math display="inline"><semantics> <mrow> <mo stretchy="false">〈</mo> <msub> <mover accent="true"> <mi>U</mi> <mo stretchy="false">¯</mo> </mover> <mi>ref</mi> </msub> <mo stretchy="false">〉</mo> </mrow> </semantics></math>, was obtained by averaging the IDDES streamwise velocity at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <msub> <mi>H</mi> <mi>avg</mi> </msub> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math>. Geometric constants were calculated at 45°.</p>
Full article ">Figure 14
<p>Time-averaged normalised vorticity magnitude for the Central neighbourhood and an external easterly wind (indicated by the arrow): (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> <mi>y</mi> </mrow> </semantics></math> plane at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> m; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>–</mo> <mi>z</mi> </mrow> </semantics></math> plane at <math display="inline"><semantics> <mrow> <mi>y</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>90</mn> </mrow> </semantics></math> m (indicated by the white dashed line in panel <b>a</b>).</p>
Full article ">Figure 15
<p>Vertical profiles of normalised mean velocities for the heterogeneous neighbourhood: (<b>top</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>x</mi> </msub> </semantics></math>; (<b>bottom</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>y</mi> </msub> </semantics></math>. The vertical profiles represent horizontal averages over: (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>) the entire domain (excluding lateral buffer); (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>) <math display="inline"><semantics> <msup> <mo>ℜ</mo> <mrow> <mi>i</mi> <mn>1</mn> </mrow> </msup> </semantics></math> (shown in red in <a href="#atmosphere-14-00050-f001" class="html-fig">Figure 1</a>e); (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>) <math display="inline"><semantics> <msup> <mo>ℜ</mo> <mrow> <mi>i</mi> <mn>2</mn> </mrow> </msup> </semantics></math> (shown in violet in <a href="#atmosphere-14-00050-f001" class="html-fig">Figure 1</a>e). Westerly results are shown at the left (<b>a</b>–<b>c</b>,<b>g</b>–<b>i</b>), easterly results at the right (<b>d</b>–<b>f</b>,<b>j</b>–<b>l</b>). The reference value, <math display="inline"><semantics> <mrow> <mo stretchy="false">〈</mo> <msub> <mover accent="true"> <mi>U</mi> <mo stretchy="false">¯</mo> </mover> <mi>ref</mi> </msub> <mo stretchy="false">〉</mo> </mrow> </semantics></math>, was obtained by averaging the IDDES streamwise velocity at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <msub> <mi>H</mi> <mi>avg</mi> </msub> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math>. The geometric constants correspond to the westerly case and the average over the entire domain, <math display="inline"><semantics> <msub> <mi>D</mi> <mrow> <mi>w</mi> <mi>h</mi> <mi>o</mi> <mi>l</mi> <mi>e</mi> </mrow> </msub> </semantics></math>. Agreement is degraded for the averages over the subdomains, <math display="inline"><semantics> <msub> <mi>D</mi> <msup> <mo>ℜ</mo> <mrow> <mi>i</mi> <mn>1</mn> </mrow> </msup> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>D</mi> <msup> <mo>ℜ</mo> <mrow> <mi>i</mi> <mn>2</mn> </mrow> </msup> </msub> </semantics></math>.</p>
Full article ">Figure 16
<p>Flow structures for stratified flow within an <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math> street canyon: (<b>top</b>) stable conditions (<math display="inline"><semantics> <mrow> <mi>Rb</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.38</mn> </mrow> </semantics></math>); (<b>bottom</b>) unstable conditions (<math display="inline"><semantics> <mrow> <mi>Rb</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mo>−</mo> <mn>0.39</mn> </mrow> </semantics></math>). (<b>a</b>,<b>d</b>) Spatially averaged streamlines; (<b>b</b>,<b>e</b>) spatially averaged vorticity magnitude; (<b>c</b>,<b>f</b>) vertical profiles of normalised TKE and vorticity magnitude. See <a href="#atmosphere-14-00050-f004" class="html-fig">Figure 4</a> for additional figure details.</p>
Full article ">Figure 17
<p>Vertical profiles of normalised streamwise velocities for <span class="html-italic">AR</span> = 1 and different Rb. (<b>Top</b>) Stable stratification; (<b>bottom</b>) unstable stratification. Predicted (solid lines) and IDDES (dashed lines) results are compared at <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>45</mn> </mrow> </semantics></math>°; the geometric constants were calculated at 0°.</p>
Full article ">Figure 18
<p>Canyon-averaged relative error in <math display="inline"><semantics> <msub> <mi>u</mi> <mi>x</mi> </msub> </semantics></math> versus Rb.</p>
Full article ">Figure A1
<p>(<b>a</b>–<b>j</b>) Vertical profiles of normalised velocities within the deep canyon (<math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>3</mn> </mrow> </semantics></math>) for the complete set of vortex sheets <math display="inline"><semantics> <msup> <mi mathvariant="normal">Ω</mi> <mrow> <mi>c</mi> <mi>o</mi> <mi>m</mi> <mi>p</mi> <mi>l</mi> <mi>e</mi> <mi>t</mi> <mi>e</mi> </mrow> </msup> </semantics></math> (black dotted line), the reduced set <math display="inline"><semantics> <msup> <mi mathvariant="normal">Ω</mi> <mrow> <mi>d</mi> <mi>e</mi> <mi>e</mi> <mi>p</mi> </mrow> </msup> </semantics></math> (red line), and the IDDES baseline (blue line).</p>
Full article ">Figure A2
<p>Vertical profiles of the normalised vorticity magnitude for the numerical configurations of <a href="#atmosphere-14-00050-t0A3" class="html-table">Table A3</a> and an external wind at 0°.</p>
Full article ">Figure A3
<p>Vertical profiles of normalised streamwise velocities for <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>R</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> </mrow> </semantics></math>°. Predicted profiles (solid lines) were obtained from the vortex sheet strengths of <a href="#atmosphere-14-00050-t0A4" class="html-table">Table A4</a>; they all show similarly good agreement with the IDDES reference (dashed lines).</p>
Full article ">Figure A4
<p>As in <a href="#atmosphere-14-00050-f002" class="html-fig">Figure 2</a>, but for the runs of <a href="#atmosphere-14-00050-t0A3" class="html-table">Table A3</a>. Validation of the IDDES model (OpenFOAM) for neutral flow over a 2-D street canyon. Normalised mean streamwise velocity and TKE validation of IDDES (solid line) against wind tunnel experiment [<a href="#B54-atmosphere-14-00050" class="html-bibr">54</a>] (filled circles) at (<b>a</b>,<b>f</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mo>−</mo> <mn>0.4</mn> </mrow> </semantics></math>; (<b>b</b>,<b>g</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mo>−</mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>c</b>,<b>h</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> </mrow> </semantics></math>; (<b>d</b>,<b>i</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>e</b>,<b>j</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>W</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.4</mn> </mrow> </semantics></math>. Reference values for the streamwise velocity and TKE, <math display="inline"><semantics> <mrow> <mo>〈</mo> <mover> <msub> <mi>U</mi> <mi>s</mi> </msub> <mo>¯</mo> </mover> <mo>〉</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>〈</mo> <mover> <mrow> <mi>T</mi> <mi>K</mi> <msub> <mi>E</mi> <mi>s</mi> </msub> </mrow> <mo>¯</mo> </mover> <mo>〉</mo> </mrow> </semantics></math>, represent averages of the IDDES data over the shear layer, <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mo>≤</mo> <mo> </mo> <mi>z</mi> <mo>/</mo> <mi>H</mi> <mo> </mo> <mo>≤</mo> <mo> </mo> <mn>1.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure A5
<p>Vertical profiles of temporally and spatially averaged normalised velocity for two different resolutions. (<b>a</b>) a 2-D street canyon (<span class="html-italic">AR</span> = 1); (<b>b</b>) a realistic canyon (<a href="#atmosphere-14-00050-f001" class="html-fig">Figure 1</a>d). Finer grid sizes (<math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.5</mn> </mrow> </semantics></math> m) are represented by green dotted lines; coarser grid sizes (<math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> </mrow> </semantics></math> m) are indicated by red dash-dotted lines.</p>
Full article ">Figure A6
<p>Absolute (<math display="inline"><semantics> <msub> <mi>ε</mi> <mi>a</mi> </msub> </semantics></math>) and relative (<math display="inline"><semantics> <msub> <mi>ε</mi> <mi>r</mi> </msub> </semantics></math>) errors versus <math display="inline"><semantics> <mrow> <mi>B</mi> <mi>R</mi> </mrow> </semantics></math> for the asymmetric canyons. <math display="inline"><semantics> <msub> <mi>ε</mi> <mi>a</mi> </msub> </semantics></math> is normalised by <math display="inline"><semantics> <msub> <mi>U</mi> <mi>ref</mi> </msub> </semantics></math> (see Equation (4)).</p>
Full article ">Figure A7
<p>As in <a href="#atmosphere-14-00050-f010" class="html-fig">Figure 10</a>, but for the streamwise velocity component over the subregions. (<b>a</b>,<b>b</b>) step-up canyon; (<b>c</b>,<b>d</b>) step-down canyon.</p>
Full article ">Figure A8
<p>As in <a href="#atmosphere-14-00050-f012" class="html-fig">Figure 12</a>, but for the heterogeneous neighbourhood and an easterly wind. (<b>a</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>x</mi> </msub> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <msub> <mi>u</mi> <mi>y</mi> </msub> </semantics></math>. <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>i</mi> </msub> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> <mo>,</mo> <mo> </mo> <msub> <mi>ε</mi> <mi>r</mi> </msub> </mrow> </semantics></math> are represented by yellow dotted lines; <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>i</mi> </msub> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0</mn> <mo>,</mo> <mo> </mo> <msub> <mi>ε</mi> <mi>a</mi> </msub> </mrow> </semantics></math> are represented by pink dashed lines; <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>i</mi> </msub> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> <mo>,</mo> <mo> </mo> <msub> <mi>ε</mi> <mi>r</mi> </msub> </mrow> </semantics></math> are represented by greed dotted lines; <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>i</mi> </msub> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>1</mn> <mo>,</mo> <mo> </mo> <msub> <mi>ε</mi> <mi>a</mi> </msub> </mrow> </semantics></math> are represented by black dashed lines; <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>i</mi> </msub> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>2</mn> <mo>,</mo> <mo> </mo> <msub> <mi>ε</mi> <mi>r</mi> </msub> </mrow> </semantics></math> are represented by red dotted lines; <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>i</mi> </msub> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>2</mn> <mo>,</mo> <mo> </mo> <msub> <mi>ε</mi> <mi>a</mi> </msub> </mrow> </semantics></math> are represented by blue dashed lines.</p>
Full article ">
14 pages, 537 KiB  
Article
Simulation Studies on the Dissipative Modified Kawahara Solitons in a Complex Plasma
by Sherif M. E. Ismaeel, Abdul-Majid Wazwaz, Elsayed Tag-Eldin and Samir A. El-Tantawy
Symmetry 2023, 15(1), 57; https://doi.org/10.3390/sym15010057 - 26 Dec 2022
Cited by 36 | Viewed by 1812
Abstract
In this work, a damped modified Kawahara equation (mKE) with cubic nonlinearity and two dispersion terms including the third- and fifth-order derivatives is analyzed. We employ an effective semi-analytical method to achieve the goal set for this study. For this purpose, the ansatz [...] Read more.
In this work, a damped modified Kawahara equation (mKE) with cubic nonlinearity and two dispersion terms including the third- and fifth-order derivatives is analyzed. We employ an effective semi-analytical method to achieve the goal set for this study. For this purpose, the ansatz method is implemented to find some approximate solutions to the damped mKE. Based on the proposed method, two different formulas for the analytical symmetric approximations are formally obtained. The derived formulas could be utilized for studying all traveling waves described by the damped mKE, such as symmetric solitary waves (SWs), shock waves, cnoidal waves, etc. Moreover, the energy of the damped dressed solitons is derived. Furthermore, the obtained approximations are used for studying the dynamics of the dissipative dressed (modified Kawahara (mK)) dust-ion acoustic (DIA) solitons in an unmagnetized collisional superthermal plasma consisting of inertia-less superthermal electrons and inertial cold ions as well as immobile negative dust grains. Numerically, the impact of the collisional parameter that arises as a result of taking the ion-neutral collisions into account and the electron spectral index on the profile of the dissipative structures are examined. Finally, the analytical and numerical approximations using the finite difference method (FDM) are compared in order to confirm the high accuracy of the obtained approximations. The achieved results contribute to explaining the mystery of several nonlinear phenomena that arise in different plasma physics, nonlinear optics, shallow water waves, oceans, and seas, and so on. Full article
(This article belongs to the Section Mathematics)
Show Figures

Figure 1

Figure 1
<p>The damped modified Kawahara soliton energy <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mi>R</mi> </msub> <mfenced open="(" close=")"> <mi>τ</mi> </mfenced> </mrow> </semantics></math> is plotted against the physical parameters <math display="inline"><semantics> <mfenced separators="" open="(" close=")"> <mi>κ</mi> <mo>,</mo> <mi>R</mi> </mfenced> </semantics></math>.</p>
Full article ">Figure 2
<p>The profile of the dissipative dressed solitons according to (<b>a</b>) the approximation (<a href="#FD27-symmetry-15-00057" class="html-disp-formula">27</a>) and (<b>b</b>) the approximation (<a href="#FD28-symmetry-15-00057" class="html-disp-formula">28</a>) is plotted against the collisional frequency parameter <span class="html-italic">R</span>.</p>
Full article ">Figure 3
<p>The profile of the dissipative dressed solitons according to (<b>a</b>) the approximation (<a href="#FD27-symmetry-15-00057" class="html-disp-formula">27</a>) and (<b>b</b>) the approximation (<a href="#FD28-symmetry-15-00057" class="html-disp-formula">28</a>) is plotted against the time of propagation <math display="inline"><semantics> <mi>τ</mi> </semantics></math>.</p>
Full article ">Figure 4
<p>The profile of the dissipative dressed solitons according to (<b>a</b>) the approximation (<a href="#FD27-symmetry-15-00057" class="html-disp-formula">27</a>) and (<b>b</b>) the approximation (<a href="#FD28-symmetry-15-00057" class="html-disp-formula">28</a>) is plotted against the electron spectral index <math display="inline"><semantics> <mi>κ</mi> </semantics></math>.</p>
Full article ">Figure 5
<p>A comparison between (<b>a</b>) the approximations (<a href="#FD27-symmetry-15-00057" class="html-disp-formula">27</a>) and (<a href="#FD28-symmetry-15-00057" class="html-disp-formula">28</a>) as well as FDM approximation and (<b>b</b>) the approximations (<a href="#FD27-symmetry-15-00057" class="html-disp-formula">27</a>) and (<a href="#FD28-symmetry-15-00057" class="html-disp-formula">28</a>) as well as the numerical solution using Wolfram Mathematica.</p>
Full article ">
22 pages, 4038 KiB  
Article
Dynamic Analysis of Sigmoid Bidirectional FG Microbeams under Moving Load and Thermal Load: Analytical Laplace Solution
by Mohamed A. Attia, Ammar Melaibari, Rabab A. Shanab and Mohamed A. Eltaher
Mathematics 2022, 10(24), 4797; https://doi.org/10.3390/math10244797 - 16 Dec 2022
Cited by 14 | Viewed by 1531
Abstract
This paper presents for the first time a closed-form solution of the dynamic response of sigmoid bidirectional functionally graded (SBDFG) microbeams under moving harmonic load and thermal environmental conditions. The formulation is established in the context of the modified couple stress theory to [...] Read more.
This paper presents for the first time a closed-form solution of the dynamic response of sigmoid bidirectional functionally graded (SBDFG) microbeams under moving harmonic load and thermal environmental conditions. The formulation is established in the context of the modified couple stress theory to integrate the effects of microstructure. On the basis of the elasticity theory, nonclassical governing equations are derived by using Hamilton’s principle in combination with the parabolic higher-order shear deformation theory considering the physical neutral plane concept. Sigmoid distribution functions are used to describe the temperature-dependent thermomechanical material of bulk continuums of the beam in both the axial and thickness directions, and the gradation of the material length scale parameter is also considered. Linear and nonlinear temperature profiles are considered to present the environmental thermal loads. The Laplace transform is exploited for the first time to evaluate the closed-form solution of the proposed model for a simply supported (SS) boundary condition. The solution is verified by comparing the predicted fundamental frequency and dynamic response with the previously published results. A parametric study is conducted to explore the impacts of gradient indices in both directions, graded material length scale parameters, thermal loads, and moving speed of the acted load on the dynamic response of microbeams. The results can serve as a principle for evaluating the multi-functional and optimal design of microbeams acted upon by a moving load. Full article
(This article belongs to the Section Engineering Mathematics)
Show Figures

Figure 1

Figure 1
<p>Illustration of a 2D-FG microbeam exposed to a moving load and thermal environment.</p>
Full article ">Figure 2
<p>Comparison of the dependency of the dimensionless fundamental frequency of the beam at different temperature differences <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>T</mi> </mrow> </semantics></math> under LTR for <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>x</mi> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mrow> <mn>0.5</mn> <mo>,</mo> <mo> </mo> <mn>2</mn> </mrow> <mo>)</mo> </mrow> <mo>,</mo> <mo> </mo> <msub> <mi>k</mi> <mi>z</mi> </msub> <mo>=</mo> <mn>0.5</mn> <mo>,</mo> <mo> </mo> <mi>l</mi> <mi>c</mi> <mi>s</mi> <mo>=</mo> <mn>0.25</mn> <mo> </mo> <mi>h</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>The variation in the dynamic deflection at the center of the beam vs. moving load velocity.</p>
Full article ">Figure 4
<p>Influence of the transverse gradient index on the dimensionless fundamental frequency at different temperature differences <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>T</mi> </mrow> </semantics></math> under LTR and based on classical analysis <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>l</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mo> </mo> <msub> <mi>k</mi> <mi>x</mi> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Influence of the axial gradient index on the dimensionless fundamental frequency at different temperature differences <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>T</mi> </mrow> </semantics></math> under LTR and based on classical analysis <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi>l</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mo> </mo> <msub> <mi>k</mi> <mi>z</mi> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Influence of the transverse gradient index on the variation in the maximum normalized central dynamic deflection with the velocity <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>k</mi> <mi>x</mi> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>; (__) temperature-independent, (- -) temperature-dependent <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>T</mi> <mo>=</mo> <mn>80</mn> <mo>,</mo> <mo> </mo> <mi>L</mi> <mi>T</mi> <mi>R</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Influence of the axial gradient index on the variation in the maximum normalized central dynamic deflection with the velocity <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>k</mi> <mi>z</mi> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>; (__) temperature-independent, (- -) temperature-dependent <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>T</mi> <mo>=</mo> <mn>80</mn> <mo>,</mo> <mo> </mo> <mi>L</mi> <mi>T</mi> <mi>R</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Influence of temperature distribution (temperature-independent, LTR and NTR with <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>) on the dimensionless central deflection vs. time under a uniform moving load (<math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>x</mi> </msub> <mo>=</mo> <msub> <mi>k</mi> <mi>z</mi> </msub> </mrow> </semantics></math> = 0.2) at <math display="inline"><semantics> <mover accent="true"> <mi>v</mi> <mo>¯</mo> </mover> </semantics></math> = 0.1, 0.4, and 0.8 and <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>T</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Influence of the dimensionless velocity on the variation in the dimensionless central deflection vs. the dimensionless time under a uniform moving load based on CL and CS formulations (<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>T</mi> <mo>=</mo> <mn>80</mn> <mo>,</mo> <mo> </mo> <mi>L</mi> <mi>T</mi> <mi>R</mi> <mo>,</mo> <mo> </mo> <msub> <mi>k</mi> <mi>x</mi> </msub> <mo>=</mo> <msub> <mi>k</mi> <mi>z</mi> </msub> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math>.</p>
Full article ">
11 pages, 266 KiB  
Article
Existence Results of Periodic Solutions to First-Order Neutral Differential Equations on Time Scales
by Qiaoling Zhang, Bo Du and Xueping Hu
Symmetry 2022, 14(11), 2405; https://doi.org/10.3390/sym14112405 - 14 Nov 2022
Viewed by 1324
Abstract
The purpose of this paper is to study the existence of periodic solutions for the first-order nonlinear neutral differential equation on time scales. Burton–Krasnoselskii’s fixed point theorem will be sufficiently general for application to the considered equation. An example has been carried out [...] Read more.
The purpose of this paper is to study the existence of periodic solutions for the first-order nonlinear neutral differential equation on time scales. Burton–Krasnoselskii’s fixed point theorem will be sufficiently general for application to the considered equation. An example has been carried out to show our results. It should be pointed out that the problem of periodic solutions is one of the current hot topics in the study of dynamic equations, which contains rich symmetry ideas and methods. Full article
(This article belongs to the Section Mathematics)
15 pages, 3944 KiB  
Article
Dynamic Model for Caragana korshinskii Shrub Aboveground Biomass Based on Theoretical and Allometric Growth Equations
by Xuejuan Jin, Hao Xu, Bo Wang and Xiaohua Wang
Forests 2022, 13(9), 1444; https://doi.org/10.3390/f13091444 - 8 Sep 2022
Cited by 1 | Viewed by 1431
Abstract
As one of the ways to achieve carbon neutralization, shrub biomass plays an important role for natural resource management decision making in arid regions. To investigate biomass dynamic variations of Caragana korshinskii, a typical shrub found in the arid desert area of [...] Read more.
As one of the ways to achieve carbon neutralization, shrub biomass plays an important role for natural resource management decision making in arid regions. To investigate biomass dynamic variations of Caragana korshinskii, a typical shrub found in the arid desert area of Ningxia, northwest China, we combined a nonlinear simultaneous (NLS) equation system with theoretical growth (TG) and allometric growth (AG) equations. On the basis of a large biomass survey dataset and analytical data of shrub stems, four methods (NOLS, NSUR, 2SLS, and 3SLS) of the NLS equations system were combined with the TG and AG equations. A model was subsequently established to predict the AGB growth of C. korshinskii. The absolute mean residual (AMR), root mean system error (RMSE), and adjusted determination coefficient (adj-R2) were used to evaluate the performance of the equations. Results revealed that the NSUR method of the NLS equations had better performance than other methods and the independent equations for BD and H growth and AGB. Additionally, the NSUR method exhibited extremely significant differences (p < 0.0001) when compared with the equations without heteroscedasticity on the basis of the likelihood ratio (LR) test, which used the power function (PF) as the variance function. The NSUR method of the NLS equations was an efficient method for predicting the dynamic growth of AGB by combining the TG and AG equations and could estimate the carbon storage for shrubs accurately, which was important for stand productivity and carbon sequestration capacity. Full article
(This article belongs to the Special Issue Advances in Monitoring and Assessment of Forest Carbon Storage)
Show Figures

Figure 1

Figure 1
<p>Map (ArcGIS v10.4.1) of the study area located in Ningxia Province, China.</p>
Full article ">Figure 2
<p>Relationships between the observed and fitted values of BD growth, H growth, and AGB models of <span class="html-italic">C. korshinskii</span> shrubs according to the NSUR method of the NLS equation system.</p>
Full article ">Figure 3
<p>Plots of standardized residuals against fitted values of the BD growth, H growth, and AGB models with and without heteroscedasticity of <span class="html-italic">C. korshinskii</span> shrubs according to the NSUR method of the NLS equation system.</p>
Full article ">Figure 4
<p>Predicted BD, H, and AGB values of <span class="html-italic">C. korshinskii</span> shrubs using the NLS equation and predicted AGB versus age and number of stems by the Gompertz equation using GNL regression. Black dots indicate the observed values.</p>
Full article ">
21 pages, 456 KiB  
Article
Exponential Stability of Highly Nonlinear Hybrid Differently Structured Neutral Stochastic Differential Equations with Unbounded Delays
by Boliang Lu, Quanxin Zhu and Ping He
Fractal Fract. 2022, 6(7), 385; https://doi.org/10.3390/fractalfract6070385 - 9 Jul 2022
Cited by 3 | Viewed by 1613
Abstract
This paper mainly studies the exponential stability of the highly nonlinear hybrid neutral stochastic differential equations (NSDEs) with multiple unbounded time-dependent delays and different structures. We prove the existence and uniqueness of the exact global solution of the new stochastic system, and then [...] Read more.
This paper mainly studies the exponential stability of the highly nonlinear hybrid neutral stochastic differential equations (NSDEs) with multiple unbounded time-dependent delays and different structures. We prove the existence and uniqueness of the exact global solution of the new stochastic system, and then give several criteria of the exponential stability, including the q1th moment and almost surely exponential stability. Additionally, some numerical examples are given to illustrate the main results. Such systems are widely applied in physics and other fields. For example, a specific case is pantograph dynamics, in which the delay term is a proportional function. These are widely used to determine the motion of a pantograph head on an electric locomotive collecting current from an overhead trolley wire. Compared with the existing works, our results extend the single constant delay of coefficients to multiple unbounded time-dependent delays, which is more general and applicable. Full article
(This article belongs to the Special Issue Fractional Processes and Multidisciplinary Applications)
Show Figures

Figure 1

Figure 1
<p>Computer simulation of the solution <span class="html-italic">x</span>(<span class="html-italic">t</span>) of Equation (<a href="#FD48-fractalfract-06-00385" class="html-disp-formula">48</a>).</p>
Full article ">Figure 2
<p>Computer simulation of the <math display="inline"><semantics> <mrow> <mo form="prefix">ln</mo> <mrow> <mo>|</mo> <mi>x</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> <mo>|</mo> </mrow> <mo>/</mo> <mi>t</mi> </mrow> </semantics></math> of the solution <span class="html-italic">x</span>(<span class="html-italic">t</span>) of Equation (<a href="#FD48-fractalfract-06-00385" class="html-disp-formula">48</a>).</p>
Full article ">Figure 3
<p>Computer simulation of the solution <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> of Equation (<a href="#FD51-fractalfract-06-00385" class="html-disp-formula">51</a>).</p>
Full article ">Figure 4
<p>Computer simulation of the solution <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> of Equation (<a href="#FD52-fractalfract-06-00385" class="html-disp-formula">52</a>).</p>
Full article ">Figure 5
<p>Computer simulation of the solution <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> of the Equation (<a href="#FD53-fractalfract-06-00385" class="html-disp-formula">53</a>).</p>
Full article ">
12 pages, 786 KiB  
Article
Oscillation of Solutions to Third-Order Nonlinear Neutral Dynamic Equations on Time Scales
by Yang-Cong Qiu, Kuo-Shou Chiu, Said R. Grace, Qingmin Liu and Irena Jadlovská
Mathematics 2022, 10(1), 86; https://doi.org/10.3390/math10010086 - 27 Dec 2021
Cited by 4 | Viewed by 2248
Abstract
In this paper, we are concerned with the oscillation of solutions to a class of third-order nonlinear neutral dynamic equations on time scales. New oscillation criteria are presented by employing the Riccati transformation and integral averaging technique. Two examples are shown to illustrate [...] Read more.
In this paper, we are concerned with the oscillation of solutions to a class of third-order nonlinear neutral dynamic equations on time scales. New oscillation criteria are presented by employing the Riccati transformation and integral averaging technique. Two examples are shown to illustrate the conclusions. Full article
18 pages, 340 KiB  
Article
Oscillation Criteria for Third-Order Nonlinear Neutral Dynamic Equations with Mixed Deviating Arguments on Time Scales
by Zhiyu Zhang, Ruihua Feng, Irena Jadlovská and Qingmin Liu
Mathematics 2021, 9(5), 552; https://doi.org/10.3390/math9050552 - 5 Mar 2021
Cited by 4 | Viewed by 1638
Abstract
Under a couple of canonical and mixed canonical-noncanonical conditions, we investigate the oscillation and asymptotic behavior of solutions to a class of third-order nonlinear neutral dynamic equations with mixed deviating arguments on time scales. By means of the double Riccati transformation and the [...] Read more.
Under a couple of canonical and mixed canonical-noncanonical conditions, we investigate the oscillation and asymptotic behavior of solutions to a class of third-order nonlinear neutral dynamic equations with mixed deviating arguments on time scales. By means of the double Riccati transformation and the inequality technique, new oscillation criteria are established, which improve and generalize related results in the literature. Several examples are given to illustrate the main results. Full article
(This article belongs to the Special Issue Oscillation Theory for Differential Equations)
11 pages, 273 KiB  
Article
New Criteria on Oscillatory and Asymptotic Behavior of Third-Order Nonlinear Dynamic Equations with Nonlinear Neutral Terms
by Said R. Grace, Jehad Alzabut and Abdullah Özbekler
Entropy 2021, 23(2), 227; https://doi.org/10.3390/e23020227 - 15 Feb 2021
Cited by 4 | Viewed by 2024
Abstract
In the paper, we provide sufficient conditions for the oscillatory and asymptotic behavior of a new type of third-order nonlinear dynamic equations with mixed nonlinear neutral terms. Our theorems not only improve and extend existing theorems in the literature but also provide a [...] Read more.
In the paper, we provide sufficient conditions for the oscillatory and asymptotic behavior of a new type of third-order nonlinear dynamic equations with mixed nonlinear neutral terms. Our theorems not only improve and extend existing theorems in the literature but also provide a new approach as far as the nonlinear neutral terms are concerned. The main results are illustrated by some particular examples. Full article
(This article belongs to the Special Issue Nonlinear Dynamics and Analysis)
16 pages, 948 KiB  
Article
Venture Capital Contracting with Ambiguity Sharing and Effort Complementarity Effect
by Jiajia Chang, Zhijun Hu and Hui Yang
Mathematics 2020, 8(1), 140; https://doi.org/10.3390/math8010140 - 19 Jan 2020
Cited by 4 | Viewed by 2655
Abstract
In this paper, we established a continuous-time agency model in which an ambiguity-averse venture capitalist (VC) employs an ambiguity-neutral entrepreneur (EN) to manage an innovative project. We analyzed the connection between ambiguity sharing and incentives under double moral hazard. Applying a stochastic dynamic [...] Read more.
In this paper, we established a continuous-time agency model in which an ambiguity-averse venture capitalist (VC) employs an ambiguity-neutral entrepreneur (EN) to manage an innovative project. We analyzed the connection between ambiguity sharing and incentives under double moral hazard. Applying a stochastic dynamic programming approach, we solved the VC’s maximization problem and obtained the Hamilton–Jacobi–Bellman (HJB) equation under a special form of the value function. We showed that the optimal pay-performance sensitivity was a fixed point of a nonlinear equation. The model ambiguity on the probability measure induced a tradeoff between ambiguity sharing and the incentive compensation that improved the EN’s pay-performance sensitivity level. Besides, we simulated the model and showed that when two efforts were complementary, the VC’s effort did not monotonically decrease with respect to the pay-performance sensitivity, while the EN’s effort did not monotonically increase in the pay-performance sensitivity level. More importantly, we found that as efforts tended to be more complementary, the optimal pay-performance sensitivity tended to approach those that maximized the efforts exerted by the EN and the VC. Full article
(This article belongs to the Section Financial Mathematics)
Show Figures

Figure 1

Figure 1
<p>Changes of <math display="inline"><semantics> <mrow> <msubsup> <mi>β</mi> <mi>t</mi> <mo>∗</mo> </msubsup> </mrow> </semantics></math> in ambiguity aversion: (<b>a</b>) <math display="inline"><semantics> <mi>θ</mi> </semantics></math> for <math display="inline"><semantics> <mrow> <msup> <mi>ρ</mi> <mn>2</mn> </msup> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mtext> </mtext> <mn>0.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mn>1</mn> </semantics></math>. (Note: <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> ); (<b>b</b>) <math display="inline"><semantics> <mi>θ</mi> </semantics></math> for <math display="inline"><semantics> <mrow> <msup> <mi>ρ</mi> <mn>2</mn> </msup> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mtext> </mtext> <mn>0.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mn>1</mn> </semantics></math>. (Note: <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 2
<p>Efforts dynamics for different complementarity levels <math display="inline"><semantics> <mi>α</mi> </semantics></math>.</p>
Full article ">Figure 3
<p>Optimal pay-performance sensitivity level for a different degree of complementarity <math display="inline"><semantics> <mi>α</mi> </semantics></math>.</p>
Full article ">
Back to TopTop