[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (90)

Search Parameters:
Keywords = center manifold

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 2759 KiB  
Article
Stability and Hopf Bifurcation Analysis of a Predator–Prey Model with Weak Allee Effect Delay and Competition Delay
by Yurong Dong, Hua Liu, Yumei Wei, Qibin Zhang and Gang Ma
Mathematics 2024, 12(18), 2853; https://doi.org/10.3390/math12182853 - 13 Sep 2024
Abstract
The purpose of this paper is to study a predator–prey model with Allee effect and double time delays. This research examines the dynamics of the model, with a focus on positivity, existence, stability and Hopf bifurcations. The stability of the periodic solution and [...] Read more.
The purpose of this paper is to study a predator–prey model with Allee effect and double time delays. This research examines the dynamics of the model, with a focus on positivity, existence, stability and Hopf bifurcations. The stability of the periodic solution and the direction of the Hopf bifurcation are elucidated by applying the normal form theory and the center manifold theorem. To validate the correctness of the theoretical analysis, numerical simulations were conducted. The results suggest that a weak Allee effect delay can promote stability within the model, transitioning it from instability to stability. Nevertheless, the competition delay induces periodic oscillations and chaotic dynamics, ultimately resulting in the population’s collapse. Full article
(This article belongs to the Section Mathematical Biology)
17 pages, 4909 KiB  
Article
Stability of Breathers for a Periodic Klein–Gordon Equation
by Martina Chirilus-Bruckner, Jesús Cuevas-Maraver and Panayotis G. Kevrekidis
Entropy 2024, 26(9), 756; https://doi.org/10.3390/e26090756 - 4 Sep 2024
Viewed by 261
Abstract
The existence of breather-type solutions, i.e., solutions that are periodic in time and exponentially localized in space, is a very unusual feature for continuum, nonlinear wave-type equations. Following an earlier work establishing a theorem for the existence of such structures, we bring to [...] Read more.
The existence of breather-type solutions, i.e., solutions that are periodic in time and exponentially localized in space, is a very unusual feature for continuum, nonlinear wave-type equations. Following an earlier work establishing a theorem for the existence of such structures, we bring to bear a combination of analysis-inspired numerical tools that permit the construction of such waveforms to a desired numerical accuracy. In addition, this enables us to explore their numerical stability. Our computations show that for the spatially heterogeneous form of the ϕ4 model considered herein, the breather solutions are generically unstable. Their instability seems to generically favor the motion of the relevant structures. We expect that these results may inspire further studies towards the identification of stable continuous breathers in spatially heterogeneous, continuum nonlinear wave equation models. Full article
(This article belongs to the Special Issue Recent Advances in the Theory of Nonlinear Lattices)
Show Figures

Figure 1

Figure 1
<p>Breather spatial profiles in the upper-left and -right panels are at different times (<math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mo> </mo> <mi>t</mi> <mo>≈</mo> <mn>25</mn> <mo>·</mo> <mn>2</mn> <mi>π</mi> <mo>/</mo> <msub> <mi>ω</mi> <mo>∗</mo> </msub> </mrow> </semantics></math>). The lower panel shows the time evolution of the center value (red dot). Coefficients are chosen to be a <span class="html-italic">resonance-free triplet</span> <math display="inline"><semantics> <mrow> <msub> <mi>s</mi> <mrow> <mi>s</mi> <mi>t</mi> <mi>e</mi> <mi>p</mi> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mi>q</mi> <mrow> <mi>s</mi> <mi>t</mi> <mi>e</mi> <mi>p</mi> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mi>ω</mi> <mo>∗</mo> </msub> </mrow> </semantics></math> as in Definition 2, with <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.43</mn> <mo>,</mo> <mi>ε</mi> <mo>=</mo> <mn>0.15</mn> </mrow> </semantics></math>. The initial condition is based on Proposition 1 (numerical simulation with <tt>pdepe</tt> from Matlab, R 2020a).</p>
Full article ">Figure 2
<p>Coefficients <math display="inline"><semantics> <msub> <mi>s</mi> <mrow> <mi>s</mi> <mi>t</mi> <mi>e</mi> <mi>p</mi> </mrow> </msub> </semantics></math> as defined in (<a href="#FD7-entropy-26-00756" class="html-disp-formula">7</a>) for different values of <span class="html-italic">p</span> (left: <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>6</mn> <mo>/</mo> <mn>13</mn> <mo>≈</mo> <mn>0.4615</mn> </mrow> </semantics></math>, right: <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.41</mn> </mrow> </semantics></math>). The closer <span class="html-italic">p</span> is to <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math>, the steeper and thinner the region unequal 1 is. In the limit of <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>→</mo> <mn>3</mn> <mo>/</mo> <mn>8</mn> </mrow> </semantics></math>, the step flattens and widens approaching the function identical 1. The steeper the step, the wider the band gaps. Numerically, it is more convenient to smooth out the steps using scaled versions of <math display="inline"><semantics> <mrow> <mi>tanh</mi> <mo>(</mo> <mi mathvariant="normal">x</mi> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Comparison of exact computation and numerical approximation of the step-function heterogeneity and of the discriminant of Equation (<a href="#FD9-entropy-26-00756" class="html-disp-formula">9</a>) (for <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math>) using Matlab (ODE15s). (<b>a</b>) Exact step potential (<a href="#FD7-entropy-26-00756" class="html-disp-formula">7</a>) <span class="html-italic">(solid red line)</span> and discretized version of the step function on the computational grid based on Equation (<a href="#FD19-entropy-26-00756" class="html-disp-formula">19</a>) <span class="html-italic">(black crosses)</span>. (<b>b</b>) Exact discriminant (<a href="#FD9-entropy-26-00756" class="html-disp-formula">9</a>) <span class="html-italic">(solid red line)</span> and numerical approximation of (<a href="#FD5-entropy-26-00756" class="html-disp-formula">5</a>) using (<a href="#FD19-entropy-26-00756" class="html-disp-formula">19</a>) <span class="html-italic">(black crosses).</span></p>
Full article ">Figure 4
<p>Exact band structure computed from the exact discriminant (<a href="#FD9-entropy-26-00756" class="html-disp-formula">9</a>) (red) for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <msub> <mi>s</mi> <mi>step</mi> </msub> <mo>,</mo> <mo> </mo> <mi>q</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> vs. the numerically computed band structure for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <msub> <mi>s</mi> <mi>step</mi> </msub> <mo>,</mo> <mo> </mo> <mi>q</mi> <mo>=</mo> <msub> <mi>q</mi> <mn>0</mn> </msub> <msub> <mi>s</mi> <mi>step</mi> </msub> </mrow> </semantics></math> with a smoothed step function as in (<a href="#FD19-entropy-26-00756" class="html-disp-formula">19</a>) using Matlab (black), along with odd multiples of <math display="inline"><semantics> <msub> <mi>ω</mi> <mo>∗</mo> </msub> </semantics></math> from (<a href="#FD8-entropy-26-00756" class="html-disp-formula">8</a>) (blue dashed lines). Parameter setting: <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math>. Observe how the choice of <math display="inline"><semantics> <msub> <mi>q</mi> <mn>0</mn> </msub> </semantics></math> puts the first band edge at <math display="inline"><semantics> <msub> <mi>ω</mi> <mo>∗</mo> </msub> </semantics></math> and how the band structure eventually approaches the exact expression for <math display="inline"><semantics> <mrow> <mi>q</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> for higher bands, just as predicted by theory. For <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <msub> <mi>s</mi> <mi>step</mi> </msub> <mo>,</mo> <mi>q</mi> <mo>=</mo> <mi>s</mi> <mrow> <mo>(</mo> <msub> <mi>q</mi> <mn>0</mn> </msub> <mo>−</mo> <msup> <mi>ε</mi> <mn>2</mn> </msup> <mo>)</mo> </mrow> </mrow> </semantics></math> with small <math display="inline"><semantics> <mi>ε</mi> </semantics></math>, the frequency (<math display="inline"><semantics> <msub> <mi>ω</mi> <mo>∗</mo> </msub> </semantics></math>) moves slightly into a spectral gap.</p>
Full article ">Figure 5
<p>Numerical linear dispersion for <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math> obtained from (<a href="#FD20-entropy-26-00756" class="html-disp-formula">20</a>) (green dots) and numerically computed dispersion relation via Floquet discriminant (black dots) for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> </semantics></math> as in (<a href="#FD19-entropy-26-00756" class="html-disp-formula">19</a>). The frequency (<math display="inline"><semantics> <msup> <mi>ω</mi> <mo>∗</mo> </msup> </semantics></math>) and its odd multiples up to 15 are indicated by blue dashed lines.</p>
Full article ">Figure 6
<p>Breather at <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>ω</mi> <mo>=</mo> <msup> <mi>ω</mi> <mo>∗</mo> </msup> </mrow> </semantics></math>. The left panel shows the profile at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, while the right panel shows the odd-<span class="html-italic">k</span> Fourier coefficients (<math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> </mrow> <msub> <mi>z</mi> <mi>k</mi> </msub> <mrow> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>|</mo> </mrow> </mrow> </semantics></math>) on a semilogarithmic scale.</p>
Full article ">Figure 7
<p>Dependence of energy as a function of the breather frequency for <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>(<b>Top left</b>) Non-zero component of the phase mode. (<b>Top right</b>) Translational mode defined as <math display="inline"><semantics> <mrow> <msub> <mo>∂</mo> <mi>x</mi> </msub> <mi>u</mi> <mrow> <mo>(</mo> <mi>x</mi> <mo>,</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. (<b>Bottom</b>) Components of the localized mode. In every panel, <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>ω</mi> <mo>=</mo> <msup> <mi>ω</mi> <mo>∗</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Dependence of the modulus of the multiplier associated with the localized mode with respect to the frequency for <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Evolution of the breather wavefunction (<math display="inline"><semantics> <mrow> <mi>u</mi> <mo>(</mo> <mi>x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>) (<b>left</b>) and the logarithm (base 10) of the energy density (<math display="inline"><semantics> <mrow> <mi>h</mi> <mo>(</mo> <mi>x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>) (<b>right</b>) with respect to time (<b>left</b>) for a moving breather with <math display="inline"><semantics> <mrow> <mi>ω</mi> <mo>=</mo> <mn>2.64</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math>, generated by a perturbation of <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Evolution of the energy center for moving breathers with <math display="inline"><semantics> <mrow> <mi>ω</mi> <mo>=</mo> <mn>2.64</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>ω</mi> <mo>=</mo> <mn>2.66</mn> </mrow> </semantics></math> obtained by adding a perturbation with amplitudes of <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> (<b>left</b>) and <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math> (<b>right</b>).</p>
Full article ">
18 pages, 3650 KiB  
Article
Ensuring Topological Data-Structure Preservation under Autoencoder Compression Due to Latent Space Regularization in Gauss–Legendre Nodes
by Chethan Krishnamurthy Ramanaik, Anna Willmann, Juan-Esteban Suarez Cardona, Pia Hanfeld, Nico Hoffmann and Michael Hecht
Axioms 2024, 13(8), 535; https://doi.org/10.3390/axioms13080535 - 7 Aug 2024
Viewed by 646
Abstract
We formulate a data-independent latent space regularization constraint for general unsupervised autoencoders. The regularization relies on sampling the autoencoder Jacobian at Legendre nodes, which are the centers of the Gauss–Legendre quadrature. Revisiting this classic allows us to prove that regularized autoencoders ensure a [...] Read more.
We formulate a data-independent latent space regularization constraint for general unsupervised autoencoders. The regularization relies on sampling the autoencoder Jacobian at Legendre nodes, which are the centers of the Gauss–Legendre quadrature. Revisiting this classic allows us to prove that regularized autoencoders ensure a one-to-one re-embedding of the initial data manifold into its latent representation. Demonstrations show that previously proposed regularization strategies, such as contractive autoencoding, cause topological defects even in simple examples, as do convolutional-based (variational) autoencoders. In contrast, topological preservation is ensured by standard multilayer perceptron neural networks when regularized using our approach. This observation extends from the classic FashionMNIST dataset to (low-resolution) MRI brain scans, suggesting that reliable low-dimensional representations of complex high-dimensional datasets can be achieved using this regularization technique. Full article
(This article belongs to the Special Issue Differential Geometry and Its Application, 2nd edition)
Show Figures

Figure 1

Figure 1
<p>Illustration of the latent representation <math display="inline"><semantics> <mrow> <msup> <mi mathvariant="script">D</mi> <mo>′</mo> </msup> <mo>=</mo> <mi>φ</mi> <mrow> <mo>(</mo> <mi mathvariant="script">D</mi> <mo>)</mo> </mrow> <mo>⊆</mo> <msub> <mo>Ω</mo> <msub> <mi>m</mi> <mn>1</mn> </msub> </msub> </mrow> </semantics></math> of the data manifold <math display="inline"><semantics> <mrow> <mi mathvariant="script">D</mi> <mo>⊆</mo> <msub> <mo>Ω</mo> <msub> <mi>m</mi> <mn>2</mn> </msub> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>dim</mi> <mi mathvariant="script">D</mi> <mo>=</mo> <msub> <mi>m</mi> <mn>0</mn> </msub> <mo>&lt;</mo> <msub> <mi>m</mi> <mn>1</mn> </msub> <mo>&lt;</mo> <msub> <mi>m</mi> <mn>2</mn> </msub> <mo>∈</mo> <mi mathvariant="double-struck">N</mi> </mrow> </semantics></math> given by the autoencoder <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>φ</mi> <mo>,</mo> <mi>ν</mi> <mo>)</mo> </mrow> </semantics></math>. The decoder is a one-to-one mapping of the hypercube <math display="inline"><semantics> <msub> <mo>Ω</mo> <msub> <mi>m</mi> <mn>1</mn> </msub> </msub> </semantics></math> to its image <math display="inline"><semantics> <mrow> <mi>ν</mi> <mo>(</mo> <msub> <mo>Ω</mo> <msub> <mi>m</mi> <mn>1</mn> </msub> </msub> <mo>)</mo> <mo>⊃</mo> <mi mathvariant="script">D</mi> </mrow> </semantics></math>, including <math display="inline"><semantics> <mi mathvariant="script">D</mi> </semantics></math> in its interior and consequently guaranteeing Equation (<a href="#FD1-axioms-13-00535" class="html-disp-formula">1</a>).</p>
Full article ">Figure 2
<p>Circle reconstruction using various autoencoder models.</p>
Full article ">Figure 3
<p>Torus reconstruction using various autoencoder models, <math display="inline"><semantics> <mrow> <mi>dim</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>FashionMNIST reconstruction with varying levels of Gaussian noise, latent dimension <math display="inline"><semantics> <mrow> <mi>dim</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Two show cases of FashionMNIST reconstruction for latent dimension <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>. First row shows the input image with vertical, horizontal flips, and <math display="inline"><semantics> <mrow> <mn>0</mn> <mo>%</mo> <mo>,</mo> <mn>10</mn> <mo>%</mo> <mo>,</mo> <mn>20</mn> <mo>%</mo> <mo>,</mo> <mn>50</mn> <mo>%</mo> <mo>,</mo> <mn>70</mn> <mo>%</mo> </mrow> </semantics></math> of Gaussian noise. Rows beneath show the results of (2) MLAP-AE, (3) CNN-AE, (4) MLP-VAE, (5) CNN-VAE, (6) ContraAE, (7) AE-REG, and (8) Hybrid AE-REG.</p>
Full article ">Figure 6
<p>FashionMNIST geodesics in latent dimension <math display="inline"><semantics> <mrow> <mi>dim</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>MRI reconstruction, latent dimension <math display="inline"><semantics> <mrow> <mi>dim</mi> <mo>=</mo> <mn>40</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>MRI show case. First row shows the input image with vertical, horizontal flips, and <math display="inline"><semantics> <mrow> <mn>0</mn> <mo>%</mo> <mo>,</mo> <mn>10</mn> <mo>%</mo> <mo>,</mo> <mn>20</mn> <mo>%</mo> <mo>,</mo> <mn>50</mn> <mo>%</mo> <mo>,</mo> <mn>70</mn> <mo>%</mo> </mrow> </semantics></math> of Gaussian noise. Rows beneath show the results of (2) MLAP-AE, (3) CNN-AE, (4) MLP-VAE, (5) CNN-VAE, (6) ContraAE, (7) AE-REG, and (8) Hybrid AE-REG.</p>
Full article ">Figure 9
<p>MRI geodesics for latent dimension <math display="inline"><semantics> <mrow> <mi>dim</mi> <mo>=</mo> <mn>40</mn> </mrow> </semantics></math> with various levels of Gaussian noise.</p>
Full article ">
20 pages, 3200 KiB  
Article
Research on Stability and Bifurcation for Two-Dimensional Two-Parameter Squared Discrete Dynamical Systems
by Limei Liu and Xitong Zhong
Mathematics 2024, 12(15), 2423; https://doi.org/10.3390/math12152423 - 4 Aug 2024
Viewed by 453
Abstract
This study investigates a class of two-dimensional, two-parameter squared discrete dynamical systems. It determines the conditions for local stability at the fixed points for these proposed systems. Theoretical and numerical analyses are conducted to examine the bifurcation behavior of the proposed systems. Conditions [...] Read more.
This study investigates a class of two-dimensional, two-parameter squared discrete dynamical systems. It determines the conditions for local stability at the fixed points for these proposed systems. Theoretical and numerical analyses are conducted to examine the bifurcation behavior of the proposed systems. Conditions for the existence of Naimark–Sacker bifurcation, transcritical bifurcation, and flip bifurcation are derived using center manifold theorem and bifurcation theory. Results of the theoretical analyses are validated by numerical simulation studies. Numerical simulations also reveal the complex bifurcation behaviors exhibited by the proposed systems and their advantage in image encryption. Full article
Show Figures

Figure 1

Figure 1
<p>The output of <math display="inline"><semantics> <mi>x</mi> </semantics></math> component with respect to <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mo>−</mo> <mn>0.2</mn> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> <mo>≤</mo> <mi>b</mi> <mo>=</mo> <mo>≤</mo> <mn>0.01</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Phase portrait of system (1) for <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <msub> <mi>x</mi> <mn>0</mn> </msub> <mo>,</mo> <msub> <mi>y</mi> <mn>0</mn> </msub> <mo stretchy="false">)</mo> <mo>=</mo> <mo stretchy="false">(</mo> <mn>0.02</mn> <mo>,</mo> <mn>0.01</mn> <mo stretchy="false">)</mo> <mo>,</mo> <mi>a</mi> <mo>=</mo> <mo>−</mo> <mn>0.2</mn> <mo>,</mo> <mi>b</mi> <mo>=</mo> <mo>−</mo> <mn>0.6717</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Maximum modulus of the eigenvalues of Jacobian matrix at <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mn>0</mn> <mo>,</mo> <mn>0</mn> <mo stretchy="false">)</mo> </mrow> </semantics></math> with respect to <span class="html-italic">b</span>.</p>
Full article ">Figure 4
<p>Output of <math display="inline"><semantics> <mi>x</mi> </semantics></math> component with respect to <math display="inline"><semantics> <mi>b</mi> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Phase portrait of system (1) for <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <msub> <mi>x</mi> <mn>0</mn> </msub> <mo>,</mo> <msub> <mi>y</mi> <mn>0</mn> </msub> <mo stretchy="false">)</mo> <mo>=</mo> <mo stretchy="false">(</mo> <mn>0.02</mn> <mo>,</mo> <mn>0.01</mn> <mo stretchy="false">)</mo> <mo>,</mo> <mi>a</mi> <mo>=</mo> <mo>−</mo> <mn>0.2</mn> <mo>,</mo> <mi>b</mi> <mo>=</mo> <mn>0.5333</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Phase portrait of the proposed system (1) with <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <msub> <mi>x</mi> <mn>0</mn> </msub> <mo>,</mo> <msub> <mi>y</mi> <mn>0</mn> </msub> <mo stretchy="false">)</mo> <mo>=</mo> <mo stretchy="false">(</mo> <mn>0.01</mn> <mo>,</mo> <mn>0.01</mn> <mo stretchy="false">)</mo> <mo>,</mo> <mi>a</mi> <mo>=</mo> <mn>0.3</mn> <mo>,</mo> <mi>b</mi> <mo>=</mo> <mn>1.93</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Encryption and decryption results with the proposed system (1).</p>
Full article ">
11 pages, 248 KiB  
Article
“Except for This Hysteria, She Is the Perfect Woman”: Women and Hysteria in An Inconvenient Wife
by Nina Marie Voigt
Humanities 2024, 13(4), 100; https://doi.org/10.3390/h13040100 - 25 Jul 2024
Viewed by 427
Abstract
Historical fiction can be understood as a hybrid space: it represents the past and simultaneously allows a consideration of the culture it is written in. Under the assumption that novels help address cultural shifts and attitudes, this paper aims to investigate how, why, [...] Read more.
Historical fiction can be understood as a hybrid space: it represents the past and simultaneously allows a consideration of the culture it is written in. Under the assumption that novels help address cultural shifts and attitudes, this paper aims to investigate how, why, and with what implications medical discourses surrounding women are depicted in fiction. This paper explores the manifold conceptualizations of hysteria in An Inconvenient Wife written by Megan Chance in 1998, arguing that the novel presents a complex view of discourses of medicalization. Its central claim is that the novel constructs hysteria not only as a tool of oppression but also as a tool with which to escape social constraints and patriarchal control. Through understanding historical fiction as not merely commenting on the past, but as addressing contemporary issues, the text adds to discussions centering on intersections of medicine and literature. Full article
(This article belongs to the Special Issue Literature and Medicine)
20 pages, 10101 KiB  
Article
An Invariant Filtering Method Based on Frame Transformed for Underwater INS/DVL/PS Navigation
by Can Wang, Chensheng Cheng, Chun Cao, Xinyu Guo, Guang Pan and Feihu Zhang
J. Mar. Sci. Eng. 2024, 12(7), 1178; https://doi.org/10.3390/jmse12071178 - 13 Jul 2024
Viewed by 782
Abstract
Underwater vehicles heavily depend on the integration of inertial navigation with Doppler Velocity Log (DVL) for fusion-based localization. Given the constraints imposed by sensor costs, ensuring the optimization ability and robustness of fusion algorithms is of paramount importance. While filtering-based techniques such as [...] Read more.
Underwater vehicles heavily depend on the integration of inertial navigation with Doppler Velocity Log (DVL) for fusion-based localization. Given the constraints imposed by sensor costs, ensuring the optimization ability and robustness of fusion algorithms is of paramount importance. While filtering-based techniques such as Extended Kalman Filter (EKF) offer mature solutions to nonlinear problems, their reliance on linearization approximation may compromise final accuracy. Recently, Invariant EKF (IEKF) methods based on the concept of smooth manifolds have emerged to address this limitation. However, the optimization by matrix Lie groups must satisfy the “group affine” property to ensure state independence, which constrains the applicability of IEKF to high-precision positioning of underwater multi-sensor fusion. In this study, an alternative state-independent underwater fusion invariant filtering approach based on a two-frame group utilizing DVL, Inertial Measurement Unit (IMU), and Earth-Centered Earth-Fixed (ECEF) configuration is proposed. This methodology circumvents the necessity for group affine in the presence of biases. We account for inertial biases and DVL pole-arm effects, achieving convergence in an imperfect IEKF by either fixed observation or body observation information. Through simulations and real datasets that are time-synchronized, we demonstrate the effectiveness and robustness of the proposed algorithm. Full article
(This article belongs to the Special Issue Autonomous Marine Vehicle Operations—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>AUV is localized by multi-source data from underwater TFG-IEKF.</p>
Full article ">Figure 2
<p>HoloOcean simulation environment and vehicles, where (<b>a</b>) is the simulation environment of HoloOcean in open water, and (<b>b</b>) is the slewing body-type vehicle in HoloOcean [<a href="#B53-jmse-12-01178" class="html-bibr">53</a>].</p>
Full article ">Figure 3
<p>Comparison of localization results for four sets of trajectories in the simulation environment. The original data represent the online data after INS/DVL fusion by the AUV in a specific way (integrated within the Inertial Unit), with (<b>a</b>) trajectory 1, (<b>b</b>) trajectory 2, (<b>c</b>) trajectory 3, and (<b>d</b>) trajectory 4.</p>
Full article ">Figure 4
<p>Vehicle used for data acquisition and data acquisition environment, where (<b>a</b>) shows the AUV for data acquisition in a lake, and (<b>b</b>) shows the data acquisition in a marine environment.</p>
Full article ">Figure 5
<p>Comparison of REs between the proposed algorithm and the state-of-the-art Underwater IEKF method for each set of trajectories at different time scales. with (<b>a</b>) scenario 1, (<b>b</b>) scenario 2, (<b>c</b>) scenario 3, and (<b>d</b>) scenario 4.</p>
Full article ">Figure 5 Cont.
<p>Comparison of REs between the proposed algorithm and the state-of-the-art Underwater IEKF method for each set of trajectories at different time scales. with (<b>a</b>) scenario 1, (<b>b</b>) scenario 2, (<b>c</b>) scenario 3, and (<b>d</b>) scenario 4.</p>
Full article ">Figure 6
<p>Various types of AUV tracks in different regions, with (<b>a</b>) RE of Traj. 1, (<b>b</b>) RE of Traj. 2, (<b>c</b>) RE of Traj. 3, and (<b>d</b>) RE of Traj. 4. In the figure, circles indicate outliers, triangles indicate means, and the horizontal line in the box indicates the median.</p>
Full article ">Figure 6 Cont.
<p>Various types of AUV tracks in different regions, with (<b>a</b>) RE of Traj. 1, (<b>b</b>) RE of Traj. 2, (<b>c</b>) RE of Traj. 3, and (<b>d</b>) RE of Traj. 4. In the figure, circles indicate outliers, triangles indicate means, and the horizontal line in the box indicates the median.</p>
Full article ">
19 pages, 3576 KiB  
Article
Understanding Researchers’ AI Readiness in a Higher Education Context: Q Methodology Research
by Youmen Chaaban, Saba Qadhi, Juebei Chen and Xiangyun Du
Educ. Sci. 2024, 14(7), 709; https://doi.org/10.3390/educsci14070709 - 29 Jun 2024
Viewed by 886
Abstract
Taking a human-centered socio-cultural perspective, this study explored the manifold individual and structural processes that contribute to researchers’ AI readiness. Forty-three graduate students and faculty at one university in Qatar took part in this Q methodology study. The results represented participants’ collective perspectives [...] Read more.
Taking a human-centered socio-cultural perspective, this study explored the manifold individual and structural processes that contribute to researchers’ AI readiness. Forty-three graduate students and faculty at one university in Qatar took part in this Q methodology study. The results represented participants’ collective perspectives on what they considered relevant to their AI readiness. A 5 + 1-factor solution was accepted, illustrating diverse perspectives and no consensus. The factors were termed based on their main foci, as follows, (F-1) how technical skills are acquired, (F-2) when it is all about ethics, (F-3) when technical skills meet ethical considerations, (F-4a and F-4b) when opposites concede, and (F-5) how collaborations reflect AI readiness. The results revealed the diversity of viewpoints among participants, and the interrelations among some factors. This study recommended a holistic approach to enhance AI readiness. It suggested integrating targeted educational initiatives and developing localized ethical frameworks to promote responsible AI use across various research disciplines. Full article
(This article belongs to the Section Technology Enhanced Education)
Show Figures

Figure 1

Figure 1
<p>Grid.</p>
Full article ">Figure 2
<p>Factor array for F-1.</p>
Full article ">Figure 3
<p>Factor array for F-2.</p>
Full article ">Figure 4
<p>Factor array for F-3.</p>
Full article ">Figure 5
<p>Factor array for F-4a.</p>
Full article ">Figure 6
<p>Factor array for F-4b.</p>
Full article ">Figure 7
<p>Factor array for F-5.</p>
Full article ">
21 pages, 1263 KiB  
Article
The Kinematic Models of the SINS and Its Errors on the SE(3) Group in the Earth-Centered Inertial Coordinate System
by Ke Fang, Tijing Cai and Bo Wang
Sensors 2024, 24(12), 3864; https://doi.org/10.3390/s24123864 - 14 Jun 2024
Viewed by 455
Abstract
In this paper, the kinematic models of the Strapdown Inertial Navigation System (SINS) and its errors on the SE(3) group in the Earth-Centered Inertial frame (ECI) are established. On the one hand, with the ECI frame being regarded as the [...] Read more.
In this paper, the kinematic models of the Strapdown Inertial Navigation System (SINS) and its errors on the SE(3) group in the Earth-Centered Inertial frame (ECI) are established. On the one hand, with the ECI frame being regarded as the reference, based on the joint representation of attitude and velocity on the SE(3) group, the dynamic of the local geographic coordinate system (n-frame) and the body coordinate system (b-frame) evolve on the differentiable manifold, respectively, and the high-order expansion of the Baker–Campbell–Haussdorff equation compensates for the non-commutative motion errors stimulated by strong maneuverability. On the other hand, the kinematics of the left- and right-invariant errors of the n-frame and the b-frame on the SE(3) group are separately derived, where the errors of the b-frame completely depend on inertial sensor errors, while the errors of the n-frame rely on position errors and velocity errors. In this way, the errors brought by the inconsistency of the reference coordinate system are tackled, and a novel attitude error definition is introduced to separate and decouple the factors affecting the dynamic of the n-frame errors and the b-frame errors for better attitude estimation. Through a turntable experiment and a car-mounted field experiment, the effectiveness of the proposed kinematic models in estimating attitude has been verified, with a remarkable improvement in yaw angle accuracy in the case of large initial misalignment angles, and the models developed have better robustness compared to the traditional SE(3) group-based model. Full article
(This article belongs to the Section Navigation and Positioning)
Show Figures

Figure 1

Figure 1
<p>The attitude angle error curves at the initial misalignment angles of [<math display="inline"><semantics> <msup> <mn>0</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>0</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>0</mn> <mo>∘</mo> </msup> </semantics></math>].</p>
Full article ">Figure 2
<p>The attitude angle error curves at the initial misalignment angles of [<math display="inline"><semantics> <msup> <mn>2</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>2</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>].</p>
Full article ">Figure 3
<p>The attitude angle error curves at the initial misalignment angles of [<math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>50</mn> <mo>∘</mo> </msup> </semantics></math>].</p>
Full article ">Figure 4
<p>The variance of the yaw angle error at the initial misalignment angles of [<math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>50</mn> <mo>∘</mo> </msup> </semantics></math>].</p>
Full article ">Figure 5
<p>The attitude angle error curves at the initial misalignment angles of <math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>50</mn> <mo>∘</mo> </msup> </semantics></math> after the observation variance adjustment.</p>
Full article ">Figure 6
<p>The trajectory of the car-mounted field experiment.</p>
Full article ">Figure 7
<p>The attitude angle error curves of the four models with 0–100 s range in the car-mounted field experiment.</p>
Full article ">Figure 8
<p>The attitude angle error curves of the four models after 100 s in the car-mounted field experiment.</p>
Full article ">
124 pages, 17855 KiB  
Review
Atomic Quantum Technologies for Quantum Matter and Fundamental Physics Applications
by Jorge Yago Malo, Luca Lepori, Laura Gentini and Maria Luisa (Marilù) Chiofalo
Technologies 2024, 12(5), 64; https://doi.org/10.3390/technologies12050064 - 7 May 2024
Viewed by 3598
Abstract
Physics is living an era of unprecedented cross-fertilization among the different areas of science. In this perspective review, we discuss the manifold impact that state-of-the-art cold and ultracold-atomic platforms can have in fundamental and applied science through the development of platforms for quantum [...] Read more.
Physics is living an era of unprecedented cross-fertilization among the different areas of science. In this perspective review, we discuss the manifold impact that state-of-the-art cold and ultracold-atomic platforms can have in fundamental and applied science through the development of platforms for quantum simulation, computation, metrology and sensing. We illustrate how the engineering of table-top experiments with atom technologies is engendering applications to understand problems in condensed matter and fundamental physics, cosmology and astrophysics, unveil foundational aspects of quantum mechanics, and advance quantum chemistry and the emerging field of quantum biology. In this journey, we take the perspective of two main approaches, i.e., creating quantum analogues and building quantum simulators, highlighting that independently of the ultimate goal of a universal quantum computer to be met, the remarkable transformative effects of these achievements remain unchanged. We wish to convey three main messages. First, this atom-based quantum technology enterprise is signing a new era in the way quantum technologies are used for fundamental science, even beyond the advancement of knowledge, which is characterised by truly cross-disciplinary research, extended interplay between theoretical and experimental thinking, and intersectoral approach. Second, quantum many-body physics is unavoidably taking center stage in frontier’s science. Third, quantum science and technology progress will have capillary impact on society, meaning this effect is not confined to isolated or highly specialized areas of knowledge, but is expected to reach and have a pervasive influence on a broad range of society aspects: while this happens, the adoption of a responsible research and innovation approach to quantum technologies is mandatory, to accompany citizens in building awareness and future scaffolding. Following on all the above reflections, this perspective review is thus aimed at scientists active or interested in interdisciplinary research, providing the reader with an overview of the current status of these wide fields of research where cold and ultracold-atomic platforms play a vital role in their description and simulation. Full article
(This article belongs to the Section Quantum Technologies)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of the different atom−technology platforms considered in the <xref ref-type="sec" rid="sec2dot1-technologies-12-00064">Section 2.1</xref>. (<bold>a</bold>) Atoms with tunable interactions. Top: physical meaning of the scattering length <italic>a</italic> illustrated through the simplest case of an attractive square−well potential. <inline-formula><mml:math id="mm334"><mml:semantics><mml:mrow><mml:mi>a</mml:mi><mml:mo>&lt;</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (left): the wave function has no zeros in the physical <inline-formula><mml:math id="mm335"><mml:semantics><mml:mrow><mml:mi>r</mml:mi><mml:mo>&gt;</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> plane. <inline-formula><mml:math id="mm336"><mml:semantics><mml:mrow><mml:mi>a</mml:mi><mml:mo>&lt;</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (right): a bound state appears. Bottom left: Fano−Feshbach magnetic resonance mechanism. A static magnetic field tunes the energy difference between the threshold of the open channel and the bound state in the closed channel. Bottom right: resulting scattering length vs. magnetic field. (<bold>b</bold>) Optical lattices. From top to bottom: 1D, 2D, and 3D optical lattices leading to 2D, 1D, and 0D−like confinement, respectively. Center and bottom images are from [<xref ref-type="bibr" rid="B109-technologies-12-00064">109</xref>]. (<bold>c</bold>) Trapped ions. Linear Paul trap. Ions are trapped using static electric control fields combined with time-dependent radio−frequency oscillating electric fields to stabilize the confinement. In tight radial confinement, laser−cooled ions form a linear string (inset image for eight ions) with a spacing determined by the trade−off between external confining fields and ion−ion Coulomb repulsion. Image from [<xref ref-type="bibr" rid="B2-technologies-12-00064">2</xref>]. (<bold>d</bold>) Rydberg atoms. Left: Rydbergs are atoms excited to very high energy states. Right: Rydberg blockade: the shift due to their strong mutual interaction displaces a doubly excited state out of resonance. Center: therefore, it is unlikely to have two Rydbergs closer than the corresponding blockade radius <inline-formula><mml:math id="mm337"><mml:semantics><mml:msub><mml:mi>r</mml:mi><mml:mi>b</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula>. (<bold>e</bold>) Dipolar atoms. Atoms can be represented as dipoles, the interaction being thus necessarily anisotropic: in this 1D configuration, e.g., a rotation of the angle <inline-formula><mml:math id="mm338"><mml:semantics><mml:mi>θ</mml:mi></mml:semantics></mml:math></inline-formula> between the dipole direction and the direction of atoms line up, can turn the interaction from maximally repulsive (head to head) to maximally attractive (head−to−tail). (<bold>f</bold>) Atoms in optical cavities: the free–space cooperativity <inline-formula><mml:math id="mm339"><mml:semantics><mml:msub><mml:mi>η</mml:mi><mml:mrow><mml:mi>f</mml:mi><mml:mi>s</mml:mi></mml:mrow></mml:msub></mml:semantics></mml:math></inline-formula> related to the atom-photon strength <inline-formula><mml:math id="mm3"><mml:semantics><mml:msubsup><mml:mi>g</mml:mi><mml:mn>0</mml:mn><mml:mn>2</mml:mn></mml:msubsup></mml:semantics></mml:math></inline-formula> is amplified into <inline-formula><mml:math id="mm4"><mml:semantics><mml:mrow><mml:mi>η</mml:mi><mml:mo>=</mml:mo><mml:mn>4</mml:mn><mml:mi mathvariant="script">F</mml:mi><mml:msub><mml:mi>η</mml:mi><mml:mrow><mml:mi>f</mml:mi><mml:mi>s</mml:mi></mml:mrow></mml:msub><mml:mo>/</mml:mo><mml:mi>π</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> by the number of round trips of the photon in the cavity, and measured by the free-spectral range <inline-formula><mml:math id="mm5"><mml:semantics><mml:mi mathvariant="script">F</mml:mi></mml:semantics></mml:math></inline-formula>, a geometric quantity. The system is open and out-of-equilibrium, due to unavoidable photon leakage and atomic spontaneous emission at rates <inline-formula><mml:math id="mm6"><mml:semantics><mml:mi>κ</mml:mi></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm7"><mml:semantics><mml:mi>γ</mml:mi></mml:semantics></mml:math></inline-formula>, respectively. In fact, it can be shown that <inline-formula><mml:math id="mm8"><mml:semantics><mml:mrow><mml:mi>η</mml:mi><mml:mo>=</mml:mo><mml:mn>4</mml:mn><mml:msubsup><mml:mi>g</mml:mi><mml:mn>0</mml:mn><mml:mn>2</mml:mn></mml:msubsup><mml:mo>/</mml:mo><mml:mrow><mml:mo>(</mml:mo><mml:mi>κ</mml:mi><mml:mi>γ</mml:mi><mml:mo>)</mml:mo></mml:mrow></mml:mrow></mml:semantics></mml:math></inline-formula>, so <inline-formula><mml:math id="mm9"><mml:semantics><mml:msubsup><mml:mi>g</mml:mi><mml:mn>0</mml:mn><mml:mn>2</mml:mn></mml:msubsup></mml:semantics></mml:math></inline-formula> can be tuned by geometrical means. (<bold>g</bold>) Summary of energy scale for the interaction strengths from dilute (also with Fano-Feshbach resonances), to dipolar, to hetero-nuclear, to Rydberg atoms.</p>
Full article ">Figure 2
<p>Diagram of an open quantum system. The physical system of interest (<italic>S</italic>) is immersed in a typically larger environment (<italic>E</italic>) also denoted as the bath. The <inline-formula><mml:math id="mm340"><mml:semantics><mml:mrow><mml:mi>S</mml:mi><mml:mo>−</mml:mo><mml:mi>E</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> environment coupling can be described by a discrete set of dissipative channels <inline-formula><mml:math id="mm341"><mml:semantics><mml:msub><mml:mi>c</mml:mi><mml:mi>i</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> under the suitable approximations.</p>
Full article ">Figure 3
<p>Diagrams for the stochastic unraveling of an open system: (<bold>a</bold>) A physical system <italic>S</italic> emits photons into the environment with a characteristic rate <inline-formula><mml:math id="mm342"><mml:semantics><mml:mi>γ</mml:mi></mml:semantics></mml:math></inline-formula>; when these are recorded via our photodetector, we gain the information that the system has undergone a quantum jump; (<bold>b</bold>) The output photons of our system <italic>S</italic> is instead coupled to a strong oscillatory signal <inline-formula><mml:math id="mm343"><mml:semantics><mml:mi mathvariant="sans-serif">Ω</mml:mi></mml:semantics></mml:math></inline-formula> through a homodyne detection scheme of intensities <inline-formula><mml:math id="mm344"><mml:semantics><mml:mrow><mml:msub><mml:mi>I</mml:mi><mml:mn>1</mml:mn></mml:msub><mml:mo>,</mml:mo><mml:msub><mml:mi>I</mml:mi><mml:mn>2</mml:mn></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula> at the output of a beam splitter. The system is then weakly and continuously monitored leading to a noisy diffusive behavior instead of projective collapse.</p>
Full article ">Figure 4
<p>Construction of an MPS representation of a generic state: A quantum state representing a <italic>M</italic>-unit system, e.g., <italic>M</italic> particles or lattice sites, requires a set of complex coefficients given by the product of the local dimensions <inline-formula><mml:math id="mm345"><mml:semantics><mml:msub><mml:mi>σ</mml:mi><mml:mi>i</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula>, with <inline-formula><mml:math id="mm346"><mml:semantics><mml:mrow><mml:mi>dim</mml:mi><mml:mo>(</mml:mo><mml:msub><mml:mi>σ</mml:mi><mml:mi>i</mml:mi></mml:msub><mml:mo>)</mml:mo><mml:mo>=</mml:mo><mml:mi>d</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>, generating a <italic>M</italic>-legged tensor of dimension <inline-formula><mml:math id="mm347"><mml:semantics><mml:msup><mml:mi>d</mml:mi><mml:mi>M</mml:mi></mml:msup></mml:semantics></mml:math></inline-formula>. By performing iterative local SVDs for each dimension <inline-formula><mml:math id="mm348"><mml:semantics><mml:msub><mml:mi>σ</mml:mi><mml:mi>i</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> we can write the state as a set of product of local matrices <inline-formula><mml:math id="mm349"><mml:semantics><mml:msup><mml:mrow><mml:mi mathvariant="normal">M</mml:mi></mml:mrow><mml:mrow><mml:mo>[</mml:mo><mml:mi>i</mml:mi><mml:mo>]</mml:mo></mml:mrow></mml:msup></mml:semantics></mml:math></inline-formula> for each combination of <inline-formula><mml:math id="mm350"><mml:semantics><mml:mrow><mml:mo>{</mml:mo><mml:msub><mml:mi>σ</mml:mi><mml:mn>1</mml:mn></mml:msub><mml:mo>,</mml:mo><mml:mo>.</mml:mo><mml:mo>.</mml:mo><mml:mo>.</mml:mo><mml:mo>,</mml:mo><mml:msub><mml:mi>σ</mml:mi><mml:mi>M</mml:mi></mml:msub><mml:mo>}</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula>. These matrices can be compressed in a controlled way, according to their bipartite entanglement with the rest of the system. Figure taken from [<xref ref-type="bibr" rid="B381-technologies-12-00064">381</xref>].</p>
Full article ">Figure 5
<p>Basic scheme of variational hybrid quantum algorithms. A <italic>N</italic>-qubit quantum register is initialized in the variational state <inline-formula><mml:math id="mm351"><mml:semantics><mml:mrow><mml:mo stretchy="true" maxsize="120%">|</mml:mo><mml:msubsup><mml:mi>ψ</mml:mi><mml:mn>0</mml:mn><mml:mrow><mml:mo>(</mml:mo><mml:mn>0</mml:mn><mml:mo>)</mml:mo></mml:mrow></mml:msubsup><mml:mo>,</mml:mo><mml:msubsup><mml:mi>ψ</mml:mi><mml:mn>0</mml:mn><mml:mrow><mml:mo>(</mml:mo><mml:mn>1</mml:mn><mml:mo>)</mml:mo></mml:mrow></mml:msubsup><mml:mo>,</mml:mo><mml:mo>.</mml:mo><mml:mo>.</mml:mo><mml:mo>.</mml:mo><mml:msubsup><mml:mi>ψ</mml:mi><mml:mn>0</mml:mn><mml:mrow><mml:mo>(</mml:mo><mml:mi>N</mml:mi><mml:mo>)</mml:mo></mml:mrow></mml:msubsup><mml:mo stretchy="true" maxsize="120%">⟩</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula> at the step <inline-formula><mml:math id="mm352"><mml:semantics><mml:mrow><mml:mi>i</mml:mi><mml:mo>=</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>. The state evolves under the application of several parametric quantum gates (purple boxes) until a quantum measurement is performed (red boxes). The mean value of the relevant observables, e.g., the Hamiltonian of the quantum system, is then computed using a classical processor (yellow boxes) that feeds back to the quantum processor a set of new variational parameters, and the entire procedure restarts. New parameters should be predicted using a classical iterative minimization sub-routine so that when convergence is reached at <inline-formula><mml:math id="mm353"><mml:semantics><mml:mrow><mml:mi>i</mml:mi><mml:mo>=</mml:mo><mml:mi>I</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>, the solution to the original problem is found.</p>
Full article ">Figure 6
<p>Conceptual operation scheme of an atom interferometer. (<bold>a</bold>) Concept for the case of beam splitters and mirrors. The probability of detecting an excited two-level atom fluctuates as the interaction time is manipulated by laser pulses. If the energy of the laser photons matches the energy difference between a ground and an excited state, the population undergoes Rabi oscillations characterised by a frequency <inline-formula><mml:math id="mm354"><mml:semantics><mml:mi mathvariant="sans-serif">Ω</mml:mi></mml:semantics></mml:math></inline-formula>. A pulse of duration <inline-formula><mml:math id="mm355"><mml:semantics><mml:mfrac><mml:mi>π</mml:mi><mml:mrow><mml:mn>2</mml:mn><mml:mi mathvariant="sans-serif">Ω</mml:mi></mml:mrow></mml:mfrac></mml:semantics></mml:math></inline-formula> creates a balanced 50:50 superposition state, whereas a pulse of duration <inline-formula><mml:math id="mm356"><mml:semantics><mml:mfrac><mml:mi>π</mml:mi><mml:mi mathvariant="sans-serif">Ω</mml:mi></mml:mfrac></mml:semantics></mml:math></inline-formula> results into a population inversion. (<bold>b</bold>) Concept for an atom-interferometry Mach-Zender protocol. A sequence of three laser pulses is applied to a cluster of atoms in their ground state. First, the initial pulse splits the atomic beam by generating an equally-weighted superposition state as described in (<bold>a</bold>), inducing both momentum transfer and spatial separation. After a duration of time <italic>T</italic>, a <inline-formula><mml:math id="mm357"><mml:semantics><mml:mi>π</mml:mi></mml:semantics></mml:math></inline-formula> pulse mirrors the atomic beam by inverting the populations of the ground and excited states as in (<bold>a</bold>), subsequently making their trajectories converge. Finally, one more <inline-formula><mml:math id="mm358"><mml:semantics><mml:mfrac><mml:mi>π</mml:mi><mml:mn>2</mml:mn></mml:mfrac></mml:semantics></mml:math></inline-formula> beam-splitting pulse mixes the populations, leading to interference patterns in the output port populations. The interferometric phase, which is influenced by the external fields such as the gravitational one <inline-formula><mml:math id="mm359"><mml:semantics><mml:mover accent="true"><mml:mi>g</mml:mi><mml:mo>→</mml:mo></mml:mover></mml:semantics></mml:math></inline-formula>, can be assessed by quantifying the number of atoms in each port. The sensitivity of the system relies on the size of the enclosed space-time area encompassed by the atom trajectories.</p>
Full article ">Figure 7
<p>Conceptual map of the achieved gain <inline-formula><mml:math id="mm360"><mml:semantics><mml:mrow><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:msub><mml:mi>θ</mml:mi><mml:mi>SQL</mml:mi></mml:msub><mml:mo>=</mml:mo><mml:mn>1</mml:mn><mml:mo>/</mml:mo><mml:msqrt><mml:mi>N</mml:mi></mml:msqrt></mml:mrow></mml:semantics></mml:math></inline-formula> in phase sensitivity over the standard quantum limit, vs. the total number of atoms <italic>N</italic> or, when fluctuations are present, its average. Figure adapted from [<xref ref-type="bibr" rid="B2-technologies-12-00064">2</xref>], to provide an idea of the current technological landscape. The logarithmic scale [left, in dB, <inline-formula><mml:math id="mm361"><mml:semantics><mml:mrow><mml:mn>10</mml:mn><mml:msub><mml:mo form="prefix">log</mml:mo><mml:mn>10</mml:mn></mml:msub><mml:msup><mml:mrow><mml:mo>(</mml:mo><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:msub><mml:mi>θ</mml:mi><mml:mi>SQL</mml:mi></mml:msub><mml:mo>/</mml:mo><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:mi>θ</mml:mi><mml:mo>)</mml:mo></mml:mrow><mml:mn>2</mml:mn></mml:msup></mml:mrow></mml:semantics></mml:math></inline-formula>] and the linear scale [right, <inline-formula><mml:math id="mm362"><mml:semantics><mml:msup><mml:mrow><mml:mo>(</mml:mo><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:msub><mml:mi>θ</mml:mi><mml:mi>SQL</mml:mi></mml:msub><mml:mo>/</mml:mo><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:mi>θ</mml:mi><mml:mo>)</mml:mo></mml:mrow><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>] are employed to display the gain. As a reference, the solid thick line represents the Heisenberg limit with <inline-formula><mml:math id="mm363"><mml:semantics><mml:mrow><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:msub><mml:mi>θ</mml:mi><mml:mi>HL</mml:mi></mml:msub><mml:mo>=</mml:mo><mml:mn>1</mml:mn><mml:mo>/</mml:mo><mml:mi>N</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>. Different colors refer to different experimental platforms as follows. Black: trapped ions. Red: Bose-Einstein condensates. Blue: cold thermal ensembles. Different symbols refer to different ways of estimating sensitivity, as follows. Stars: gains directly measured in phase sensitivity, acquired through full phase estimation experiments. Circles: expected gains based on the characterization of the quantum state, as computed e.g., from <inline-formula><mml:math id="mm364"><mml:semantics><mml:mrow><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:mi>θ</mml:mi><mml:mo>=</mml:mo><mml:msub><mml:mi>ξ</mml:mi><mml:mi>R</mml:mi></mml:msub><mml:mo>/</mml:mo><mml:msqrt><mml:mi>N</mml:mi></mml:msqrt></mml:mrow></mml:semantics></mml:math></inline-formula> using the spin-squeezing parameter <inline-formula><mml:math id="mm365"><mml:semantics><mml:msub><mml:mi>ξ</mml:mi><mml:mi>R</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula>, or as <inline-formula><mml:math id="mm366"><mml:semantics><mml:mrow><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:mi>θ</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn><mml:mo>/</mml:mo><mml:msqrt><mml:msub><mml:mi>F</mml:mi><mml:mi mathvariant="normal">Q</mml:mi></mml:msub></mml:msqrt></mml:mrow></mml:semantics></mml:math></inline-formula> using the quantum Fisher information <inline-formula><mml:math id="mm367"><mml:semantics><mml:msub><mml:mi>F</mml:mi><mml:mi mathvariant="normal">Q</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula>. Filled (open) circles: results obtained without (with) the subtraction of technical and/or imaging noise. We refer to [<xref ref-type="bibr" rid="B2-technologies-12-00064">2</xref>] for the correspondence of each symbol to a different experiment.</p>
Full article ">Figure 8
<p>Showcase of phase sensitivity for ion Schrödinger cat states. Figure adapted from [<xref ref-type="bibr" rid="B2-technologies-12-00064">2</xref>]. (<bold>a</bold>) Typical parity oscillations achieved using cat states, characterized by a distinctive period of <inline-formula><mml:math id="mm368"><mml:semantics><mml:mrow><mml:mn>2</mml:mn><mml:mi>π</mml:mi><mml:mo>/</mml:mo><mml:mi>N</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> (here illustrated for <inline-formula><mml:math id="mm369"><mml:semantics><mml:mrow><mml:mi>N</mml:mi><mml:mo>=</mml:mo><mml:mn>8</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>). (<bold>b</bold>) Overview of experimental realisations. The Fisher information <italic>F</italic> obtained from experimentally extracted visibilities <italic>V</italic>, is presented as a function of the number of qubits <italic>N</italic>, specifically expressed as <inline-formula><mml:math id="mm370"><mml:semantics><mml:mrow><mml:mi>F</mml:mi><mml:mo>=</mml:mo><mml:msup><mml:mi>V</mml:mi><mml:mn>2</mml:mn></mml:msup><mml:msup><mml:mi>N</mml:mi><mml:mn>2</mml:mn></mml:msup></mml:mrow></mml:semantics></mml:math></inline-formula>. Upper thick line: Heisenberg limit (HL) <inline-formula><mml:math id="mm371"><mml:semantics><mml:mrow><mml:mi>F</mml:mi><mml:mo>=</mml:mo><mml:msup><mml:mi>N</mml:mi><mml:mn>2</mml:mn></mml:msup></mml:mrow></mml:semantics></mml:math></inline-formula>. Lower thick line: standard quantum limit (SQL) <inline-formula><mml:math id="mm372"><mml:semantics><mml:mrow><mml:mi>F</mml:mi><mml:mo>=</mml:mo><mml:mi>N</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>. Thin lines: bounds for valuable <inline-formula><mml:math id="mm373"><mml:semantics><mml:mrow><mml:mi>k</mml:mi><mml:mo>−</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula>particle entanglement, employed to perimetrise the shaded region corresponding to <inline-formula><mml:math id="mm374"><mml:semantics><mml:mrow><mml:mo>(</mml:mo><mml:mi>k</mml:mi><mml:mo>+</mml:mo><mml:mn>1</mml:mn><mml:mo>)</mml:mo><mml:mo>−</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula>particle entanglement. Notice that the darker red region indicates useful genuine <inline-formula><mml:math id="mm375"><mml:semantics><mml:mrow><mml:mi>N</mml:mi><mml:mo>−</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula>particle entanglement, while the lighter red region signifies useful <inline-formula><mml:math id="mm376"><mml:semantics><mml:mrow><mml:mo>(</mml:mo><mml:mi>N</mml:mi><mml:mo>−</mml:mo><mml:mn>1</mml:mn><mml:mo>)</mml:mo><mml:mo>−</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula>particle entanglement, and so forth. Inset: zoomed−in view for the specific case with <inline-formula><mml:math id="mm377"><mml:semantics><mml:mrow><mml:mi>N</mml:mi><mml:mo>=</mml:mo><mml:mn>2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> ions.</p>
Full article ">Figure 9
<p>BEC-BCS crossover. Schematic phase diagram within the parameter space defined by temperature <italic>T</italic> (in units of the Fermi temperature <inline-formula><mml:math id="mm378"><mml:semantics><mml:mrow><mml:msub><mml:mi>T</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mo>=</mml:mo><mml:msub><mml:mi>E</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mo>/</mml:mo><mml:msub><mml:mi>k</mml:mi><mml:mi>B</mml:mi></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula>, where <inline-formula><mml:math id="mm379"><mml:semantics><mml:msub><mml:mi>E</mml:mi><mml:mi>F</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> represents the Fermi energy) and inverse scattering length <inline-formula><mml:math id="mm380"><mml:semantics><mml:mrow><mml:mn>1</mml:mn><mml:mo>/</mml:mo><mml:mo>(</mml:mo><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>a</mml:mi><mml:mo>)</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula> in units of <inline-formula><mml:math id="mm381"><mml:semantics><mml:msubsup><mml:mi>k</mml:mi><mml:mi>F</mml:mi><mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:msubsup></mml:semantics></mml:math></inline-formula>. The green lower line separates the superfluid phase with broken symmetry from the normal phases (i.e., non−superfluid phases). The critical temperature <inline-formula><mml:math id="mm382"><mml:semantics><mml:msub><mml:mi>T</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> exhibits an increasing trend with the inverse scattering length, starting from the BCS limit and approaching the BEC value <inline-formula><mml:math id="mm383"><mml:semantics><mml:mrow><mml:msub><mml:mi>T</mml:mi><mml:mrow><mml:mi>B</mml:mi><mml:mi>E</mml:mi><mml:mi>C</mml:mi></mml:mrow></mml:msub><mml:mo>≃</mml:mo><mml:mn>0.218</mml:mn><mml:msub><mml:mi>T</mml:mi><mml:mi>F</mml:mi></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula> of a gas composed of non−interacting bosons. Notably, <inline-formula><mml:math id="mm384"><mml:semantics><mml:msub><mml:mi>T</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> passes through an optimal value around the resonance point <inline-formula><mml:math id="mm385"><mml:semantics><mml:mrow><mml:msup><mml:mrow><mml:mo>(</mml:mo><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>a</mml:mi><mml:mo>)</mml:mo></mml:mrow><mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:msup><mml:mo>=</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>. The predicted value of <inline-formula><mml:math id="mm386"><mml:semantics><mml:msub><mml:mi>T</mml:mi><mml:mrow><mml:mi>B</mml:mi><mml:mi>E</mml:mi><mml:mi>C</mml:mi></mml:mrow></mml:msub></mml:semantics></mml:math></inline-formula> can only be obtained by considering pairing fluctuations, as first emphasized by Nozières and Schmitt−Rink [<xref ref-type="bibr" rid="B555-technologies-12-00064">555</xref>]. Otherwise, without accounting for these fluctuations, <inline-formula><mml:math id="mm387"><mml:semantics><mml:msub><mml:mi>T</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> would indefinitely increase with the strength of pairing. The red dashed line represents the dissociation temperature <inline-formula><mml:math id="mm388"><mml:semantics><mml:msup><mml:mi>T</mml:mi><mml:mo>*</mml:mo></mml:msup></mml:semantics></mml:math></inline-formula> at which composite bosons, formed by two fermions, are disrupted by thermal fluctuations. This temperature can be viewed as indicative of the opening of a pseudogap in the single−particle spectral function. On the BCS side of the phase diagram, the critical and dissociation temperatures coincide, while in the BEC limit, they differ.</p>
Full article ">Figure 10
<p>Universal behavior of superfluid systems. Log-log plot of the transition temperature <inline-formula><mml:math id="mm389"><mml:semantics><mml:msub><mml:mi>T</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> as a function of the normalized gap energy <inline-formula><mml:math id="mm390"><mml:semantics><mml:mrow><mml:mn>2</mml:mn><mml:mi mathvariant="sans-serif">Δ</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> relative to the effective Fermi temperature <inline-formula><mml:math id="mm391"><mml:semantics><mml:msubsup><mml:mi>T</mml:mi><mml:mi>F</mml:mi><mml:mo>*</mml:mo></mml:msubsup></mml:semantics></mml:math></inline-formula>. The different regions correspond to BCS systems (<bold>a</bold>), the crossover regime (<bold>b</bold>), strongly bound composite bosons exhibiting BEC-like behavior (<bold>c</bold>). In (<bold>a</bold>,<bold>b</bold>), <inline-formula><mml:math id="mm392"><mml:semantics><mml:mrow><mml:mn>2</mml:mn><mml:mi mathvariant="sans-serif">Δ</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> signifies the the energy required to break apart a fermion pair and <inline-formula><mml:math id="mm393"><mml:semantics><mml:msubsup><mml:mi>T</mml:mi><mml:mi>F</mml:mi><mml:mo>*</mml:mo></mml:msubsup></mml:semantics></mml:math></inline-formula> the effective Fermi temperature built from the system density. In (<bold>c</bold>), <inline-formula><mml:math id="mm394"><mml:semantics><mml:mrow><mml:mn>2</mml:mn><mml:mi mathvariant="sans-serif">Δ</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> represents the minimum energy necessary to separate the composite boson into two fermions, that is the ionization energy leading to a charged atomic core and an electron, and <inline-formula><mml:math id="mm395"><mml:semantics><mml:msubsup><mml:mi>T</mml:mi><mml:mi>F</mml:mi><mml:mo>*</mml:mo></mml:msubsup></mml:semantics></mml:math></inline-formula> is the corresponding ionic Fermi temperature. Image from [<xref ref-type="bibr" rid="B549-technologies-12-00064">549</xref>].</p>
Full article ">Figure 11
<p>Conceptual map of BEC−BCS crossover theories within the parameter space defined by <inline-formula><mml:math id="mm396"><mml:semantics><mml:mrow><mml:mo>−</mml:mo><mml:msup><mml:mrow><mml:mo>(</mml:mo><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>a</mml:mi><mml:mo>)</mml:mo></mml:mrow><mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:msup></mml:mrow></mml:semantics></mml:math></inline-formula>, which govern the transition between the BEC and BCS limits, and by <inline-formula><mml:math id="mm397"><mml:semantics><mml:msup><mml:mrow><mml:mo>(</mml:mo><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:msub><mml:mi>r</mml:mi><mml:mn>0</mml:mn></mml:msub><mml:mo>)</mml:mo></mml:mrow><mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:msup></mml:semantics></mml:math></inline-formula>, characterizing the width of the resonance from narrow to broad, respectively. On the left side are general model−frameworks, including one or two−channel models. On the right side, the corresponding theoretical or Quantum Monte Carlo (QMC) methods employed to investigate the crossover are illustrated, covering the narrow (indicated by a red stripe), intermediate (represented by orange), and broad (depicted by green) regions. The boson-fermion local-field theory (BFLF) serves as a bridge for intermediate−to−large values of <inline-formula><mml:math id="mm398"><mml:semantics><mml:msup><mml:mrow><mml:mo>(</mml:mo><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:msub><mml:mi>r</mml:mi><mml:mn>0</mml:mn></mml:msub><mml:mo>)</mml:mo></mml:mrow><mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:msup></mml:semantics></mml:math></inline-formula>, encompassing fluctuations through a comprehensive local−field theory of the boson−fermion Hamiltonian. For an account of the different theories and corresponding references, see the main text. The conceptual map is adapted from [<xref ref-type="bibr" rid="B579-technologies-12-00064">579</xref>].</p>
Full article ">Figure 12
<p>Universal behavior of fermionic superfluids in the BCS-BEC crossover. Fermions interacting via a shape-resonance in the form of a well-barrier potential, within one-channel Hamiltonian treated in mean field. (<bold>a</bold>) Chemical potential <inline-formula><mml:math id="mm399"><mml:semantics><mml:mi>μ</mml:mi></mml:semantics></mml:math></inline-formula> at temperature <inline-formula><mml:math id="mm400"><mml:semantics><mml:mrow><mml:mi>T</mml:mi><mml:mo>=</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> vs. correlation length <inline-formula><mml:math id="mm401"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>ξ</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> in units of the Fermi wavevector <inline-formula><mml:math id="mm402"><mml:semantics><mml:msubsup><mml:mi>k</mml:mi><mml:mi>F</mml:mi><mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:msubsup></mml:semantics></mml:math></inline-formula>. <inline-formula><mml:math id="mm403"><mml:semantics><mml:mi>μ</mml:mi></mml:semantics></mml:math></inline-formula> is normalized to the Fermi energy <inline-formula><mml:math id="mm404"><mml:semantics><mml:msub><mml:mi>E</mml:mi><mml:mi>F</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> when <inline-formula><mml:math id="mm405"><mml:semantics><mml:mrow><mml:mi>μ</mml:mi><mml:mo>&gt;</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (BCS side) and to half the binding energy <inline-formula><mml:math id="mm406"><mml:semantics><mml:mrow><mml:msub><mml:mi>E</mml:mi><mml:mi>b</mml:mi></mml:msub><mml:mo>/</mml:mo><mml:mn>2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> of the composite bosons when <inline-formula><mml:math id="mm407"><mml:semantics><mml:mrow><mml:mi>μ</mml:mi><mml:mo>&lt;</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (BEC side). The model parameters are the scattering length <italic>a</italic>, effective range <inline-formula><mml:math id="mm408"><mml:semantics><mml:msup><mml:mi>R</mml:mi><mml:mo>*</mml:mo></mml:msup></mml:semantics></mml:math></inline-formula> in <inline-formula><mml:math id="mm409"><mml:semantics><mml:msubsup><mml:mi>k</mml:mi><mml:mi>F</mml:mi><mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:msubsup></mml:semantics></mml:math></inline-formula> units, and the diluteness parameter <inline-formula><mml:math id="mm410"><mml:semantics><mml:mrow><mml:mi>n</mml:mi><mml:msubsup><mml:mi>r</mml:mi><mml:mn>0</mml:mn><mml:mn>3</mml:mn></mml:msubsup></mml:mrow></mml:semantics></mml:math></inline-formula> in terms of the density <italic>n</italic> and the width of the well <inline-formula><mml:math id="mm411"><mml:semantics><mml:msub><mml:mi>r</mml:mi><mml:mn>0</mml:mn></mml:msub></mml:semantics></mml:math></inline-formula>. Each symbol refers to a different realisation of the set of model parameters, as in the legend. (<bold>b</bold>) Same as in (<bold>a</bold>), but for the condensate fraction <inline-formula><mml:math id="mm412"><mml:semantics><mml:mrow><mml:mn>2</mml:mn><mml:msub><mml:mi>N</mml:mi><mml:mn>0</mml:mn></mml:msub><mml:mo>/</mml:mo><mml:mi>N</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> at <inline-formula><mml:math id="mm413"><mml:semantics><mml:mrow><mml:mi>T</mml:mi><mml:mo>=</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> in units of the number <italic>N</italic> of fermions. Notice how both quantities show universal behavior, irrespective of the microscopic details. From [<xref ref-type="bibr" rid="B607-technologies-12-00064">607</xref>].</p>
Full article ">Figure 13
<p>Universal behavior of fermionic superfluids in the BCS-BEC crossover. Fermions interacting via a model separable potential characterized by a strength <italic>g</italic> and a range <inline-formula><mml:math id="mm414"><mml:semantics><mml:msubsup><mml:mi>k</mml:mi><mml:mn>0</mml:mn><mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:msubsup></mml:semantics></mml:math></inline-formula> like that introduced by NSR [<xref ref-type="bibr" rid="B555-technologies-12-00064">555</xref>], in the additional presence of a spin-orbit coupling with strength <inline-formula><mml:math id="mm415"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula>, and treated within the conserving <inline-formula><mml:math id="mm416"><mml:semantics><mml:mrow><mml:mi>G</mml:mi><mml:msub><mml:mi>G</mml:mi><mml:mn>0</mml:mn></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula> approximation [<xref ref-type="bibr" rid="B599-technologies-12-00064">599</xref>]. (<bold>a</bold>) Effective chemical potential <inline-formula><mml:math id="mm417"><mml:semantics><mml:mover accent="true"><mml:mi>μ</mml:mi><mml:mo>˜</mml:mo></mml:mover></mml:semantics></mml:math></inline-formula> at <inline-formula><mml:math id="mm418"><mml:semantics><mml:msub><mml:mi>T</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> vs. the pair correlation length <inline-formula><mml:math id="mm419"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>ξ</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>. <inline-formula><mml:math id="mm420"><mml:semantics><mml:mover accent="true"><mml:mi>μ</mml:mi><mml:mo>˜</mml:mo></mml:mover></mml:semantics></mml:math></inline-formula> is normalized as in <xref ref-type="fig" rid="technologies-12-00064-f012">Figure 12</xref>a. Each symbol refers to a different realisation of the set of model parameters <italic>g</italic>, <inline-formula><mml:math id="mm421"><mml:semantics><mml:msub><mml:mi>k</mml:mi><mml:mn>0</mml:mn></mml:msub></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm422"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula>, as in the legend. In the inset, a specific region is examined, highlighting the difference between <inline-formula><mml:math id="mm423"><mml:semantics><mml:mover accent="true"><mml:mi>μ</mml:mi><mml:mo>˜</mml:mo></mml:mover></mml:semantics></mml:math></inline-formula> and its non-interacting value <inline-formula><mml:math id="mm424"><mml:semantics><mml:msub><mml:mover accent="true"><mml:mi>μ</mml:mi><mml:mo>˜</mml:mo></mml:mover><mml:mrow><mml:mi>N</mml:mi><mml:mi>I</mml:mi></mml:mrow></mml:msub></mml:semantics></mml:math></inline-formula> as a function of <inline-formula><mml:math id="mm425"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula>. For larger <inline-formula><mml:math id="mm426"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula> values, pair correlation lengths <inline-formula><mml:math id="mm427"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>ξ</mml:mi><mml:mo>&gt;</mml:mo><mml:mn>2</mml:mn><mml:mi>π</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> no longer correspond to a weakly interacting BCS regime, regardless of the interaction range (represented by <inline-formula><mml:math id="mm428"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mn>0</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>4</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>). (<bold>b</bold>) Log-plot of the critical temperature <inline-formula><mml:math id="mm429"><mml:semantics><mml:msub><mml:mi>T</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> multiplied by the effective mass <inline-formula><mml:math id="mm430"><mml:semantics><mml:msub><mml:mi>m</mml:mi><mml:mi>b</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> of the non-condensed resonant pairs, vs. <inline-formula><mml:math id="mm431"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>ξ</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>. Different values of <inline-formula><mml:math id="mm432"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula> are represented by symbols in the legend. Squares correspond to <inline-formula><mml:math id="mm433"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mn>0</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, triangles to <inline-formula><mml:math id="mm434"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mn>0</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>4</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>. The coupling <italic>g</italic> ranges from <inline-formula><mml:math id="mm435"><mml:semantics><mml:mrow><mml:mn>1.25</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> to 30. In both (<bold>a</bold>,<bold>b</bold>), the vertical dashed lines at <inline-formula><mml:math id="mm436"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>ξ</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn><mml:mo>/</mml:mo><mml:mi>π</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm437"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mi>F</mml:mi></mml:msub><mml:mi>ξ</mml:mi><mml:mo>=</mml:mo><mml:mn>2</mml:mn><mml:mi>π</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> indicate the thresholds for the deep BEC and BCS regimes, respectively. Notice how both quantities show universal behavior, irrespective of the microscopic details, provided that <inline-formula><mml:math id="mm438"><mml:semantics><mml:mrow><mml:mi>λ</mml:mi><mml:mo>&lt;</mml:mo><mml:msub><mml:mi>λ</mml:mi><mml:mi>B</mml:mi></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula> be below the threshold for the transition to a topological state and the nature of the fluid changes. From [<xref ref-type="bibr" rid="B608-technologies-12-00064">608</xref>] (Permission to use this content granted by Creative Commons Licence <uri>https://creativecommons.org/licenses/by/4.0/</uri>).</p>
Full article ">Figure 14
<p>Aubry-like transition in ion platforms. Concept of the experiment in [<xref ref-type="bibr" rid="B28-technologies-12-00064">28</xref>]. Laser-cooled trapped ions with charge <inline-formula><mml:math id="mm439"><mml:semantics><mml:mrow><mml:mo>+</mml:mo><mml:mo>|</mml:mo><mml:mi>e</mml:mi><mml:mo>|</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula> are arranged at an average distance <italic>d</italic>, determined by the interplay between Coulomb repulsion and external (harmonic) confinement. Superimposed is a periodic optical lattice with period <italic>a</italic> and potential height <italic>V</italic>. The lengths <italic>a</italic> and <italic>d</italic> are chosen to have an irrational, incommensurate, ratio. In the commensurate phase, whenever <inline-formula><mml:math id="mm440"><mml:semantics><mml:mrow><mml:mi>V</mml:mi><mml:mo>&gt;</mml:mo><mml:msub><mml:mi>V</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula> exceeds a critical value <inline-formula><mml:math id="mm441"><mml:semantics><mml:msub><mml:mi>V</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula>, ions pin into the bottom of the lattice wells forming commensurate segments separated by incommensurate regions. With increasing the level of incommensuration, pinning occurs at progressively larger <inline-formula><mml:math id="mm442"><mml:semantics><mml:msub><mml:mi>V</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> values. Quantum effects arise from tunneling between lattice sites. When this happens, the quantum probability density <inline-formula><mml:math id="mm443"><mml:semantics><mml:msup><mml:mrow><mml:mo>|</mml:mo><mml:mi>ψ</mml:mi><mml:mrow><mml:mo>(</mml:mo><mml:mi>x</mml:mi><mml:mo>)</mml:mo></mml:mrow><mml:mo>|</mml:mo></mml:mrow><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (here depicted for the central particle) becomes bimodal, and can be detected. Image inspired from [<xref ref-type="bibr" rid="B643-technologies-12-00064">643</xref>].</p>
Full article ">Figure 15
<p>Commensurate-incommensurate physics. Phase diagram of the Aubry transition for the experimental realisation in <xref ref-type="fig" rid="technologies-12-00064-f014">Figure 14</xref>. The variance of the central-particle density probability is used as an indicator for the transition in the space of the governing parameters, that are the incommensuration degree <inline-formula><mml:math id="mm444"><mml:semantics><mml:mi mathvariant="sans-serif">Δ</mml:mi></mml:semantics></mml:math></inline-formula> and the normalised height <italic>K</italic> of the lattice potential. Red region: localized phase. Blue region: delocalized phase. Yellow region: intermediate phase. Figure inspired from [<xref ref-type="bibr" rid="B643-technologies-12-00064">643</xref>] (Permission to use this content granted by Creative Commons Licence <uri>https://creativecommons.org/licenses/by/4.0/</uri>).</p>
Full article ">Figure 16
<p>Commensurate-incommensurate physics in the Meissner−to−vortex transition. DMRG prediction of the phase diagram for experiments of the type of Atala et al. [<xref ref-type="bibr" rid="B27-technologies-12-00064">27</xref>] on a two−leg ladder (see also Figure 18c), but in the case of hard−core spinless bosons case, i.e., infinitely repulsive on-site bosons. Left panel: phase diagram in the space of the governing parameters, that are the flux <inline-formula><mml:math id="mm445"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula> of the synthetic gauge field per plaquette, and the coupling <inline-formula><mml:math id="mm446"><mml:semantics><mml:mrow><mml:mi mathvariant="sans-serif">Ω</mml:mi><mml:mo>/</mml:mo><mml:mi>t</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> between across the rungs in units of the tunneling <italic>t</italic> along the legs. Black solid line: phase boundary between the Meissner and vortex phases. Compared to the non−interacting case (red dashed line), the hard-core nature of the bosons favors the persistence of the Meissner phase above the threshold <inline-formula><mml:math id="mm447"><mml:semantics><mml:mrow><mml:mi mathvariant="sans-serif">Ω</mml:mi><mml:mo>&gt;</mml:mo><mml:msub><mml:mi mathvariant="sans-serif">Ω</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula> for all fluxes <inline-formula><mml:math id="mm448"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula>, except at <inline-formula><mml:math id="mm449"><mml:semantics><mml:mrow><mml:mi>λ</mml:mi><mml:mo>=</mml:mo><mml:mi>π</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>. Shaded area: a second incommensuration appears. In the green region (region II), additional peaks at <inline-formula><mml:math id="mm450"><mml:semantics><mml:mrow><mml:mi>k</mml:mi><mml:mo>=</mml:mo><mml:mi>π</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> emerge in the Fourier transform of the rung current correlations, which dominate in the blue region (region III). The double line (green vs dark red) at <inline-formula><mml:math id="mm451"><mml:semantics><mml:mrow><mml:mi>λ</mml:mi><mml:mo>=</mml:mo><mml:mi>π</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> represents the transition to a localized phase. Right panel: Intensity plots of the momentum distribution <inline-formula><mml:math id="mm452"><mml:semantics><mml:mrow><mml:mi>n</mml:mi><mml:mo>(</mml:mo><mml:mi>k</mml:mi><mml:mo>)</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula> versus <inline-formula><mml:math id="mm453"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula> and <italic>k</italic>. (<bold>a</bold>) <inline-formula><mml:math id="mm454"><mml:semantics><mml:mrow><mml:mi mathvariant="sans-serif">Ω</mml:mi><mml:mo>/</mml:mo><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1.75</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> in the Meissner phase, characterised by one single maximum at <inline-formula><mml:math id="mm455"><mml:semantics><mml:mrow><mml:mi>k</mml:mi><mml:mo>=</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for all <inline-formula><mml:math id="mm456"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula>. At <inline-formula><mml:math id="mm457"><mml:semantics><mml:mrow><mml:mi>λ</mml:mi><mml:mo>=</mml:mo><mml:mi>π</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm458"><mml:semantics><mml:mrow><mml:mi>n</mml:mi><mml:mo>(</mml:mo><mml:mi>k</mml:mi><mml:mo>)</mml:mo><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, corresponding to the formation of a fully localized state (dark-red solid line). (<bold>b</bold>,<bold>c</bold>) for <inline-formula><mml:math id="mm459"><mml:semantics><mml:mrow><mml:mi mathvariant="sans-serif">Ω</mml:mi><mml:mo>/</mml:mo><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm460"><mml:semantics><mml:mrow><mml:mn>0.25</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, respectively, showing the transition from the Meissner phase to the vortex phase, characterised by two maxima of <inline-formula><mml:math id="mm461"><mml:semantics><mml:mrow><mml:mi>n</mml:mi><mml:mo>(</mml:mo><mml:mi>k</mml:mi><mml:mo>)</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula> symmetrically located around <inline-formula><mml:math id="mm462"><mml:semantics><mml:mrow><mml:mi>k</mml:mi><mml:mo>=</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>. Figure from [<xref ref-type="bibr" rid="B655-technologies-12-00064">655</xref>].</p>
Full article ">Figure 17
<p>Schematic experimental detection protocol of Many-Body Localization (MBL). (<bold>A</bold>) The system is initially prepared in a charge-density state with half-filling, where a particle/doublon is present in every second lattice site. Over time, particles can tunnel with rate <italic>J</italic> to neighbouring sites and doublons experience a local energy offset <italic>U</italic> due to particle interactions and local random disorder of order <inline-formula><mml:math id="mm463"><mml:semantics><mml:mrow><mml:mi>D</mml:mi><mml:mi>e</mml:mi><mml:mi>l</mml:mi><mml:mi>t</mml:mi><mml:mi>a</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula>. (<bold>B</bold>) A phenomenological phase diagram illustrates the MBL transition, reflecting the ergodicity properties of the system. In the localization phase (yellow region), the initial configuration persists for long periods of time, in contrast to the ergodic delocalized phase (white region), where the particles evenly distribute across the lattice. The stripped region represents the variation in the transition point based on the initial number of doublons, highlighting the interplay between disorder <inline-formula><mml:math id="mm464"><mml:semantics><mml:mi mathvariant="sans-serif">Δ</mml:mi></mml:semantics></mml:math></inline-formula> and interaction <italic>U</italic>, which are factors included in the Hamiltonian contributions schematically depicted in (<bold>C</bold>). Figure from [<xref ref-type="bibr" rid="B33-technologies-12-00064">33</xref>].</p>
Full article ">Figure 18
<p>Floquet engineering: summary of the main protocols. (<bold>a</bold>) The behavior of atoms in a shaken optical lattice can be described by an effective tunneling constant <inline-formula><mml:math id="mm465"><mml:semantics><mml:msub><mml:mi>J</mml:mi><mml:mi>eff</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula>, which can take negative values within certain parameter ranges. When <inline-formula><mml:math id="mm466"><mml:semantics><mml:mrow><mml:msub><mml:mi>J</mml:mi><mml:mi>eff</mml:mi></mml:msub><mml:mo>&lt;</mml:mo><mml:mn>0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, condensation occurs at new minima (shown in red) in the distribution instead of the previous minima centered at zero quasi-momentum (shown in blue). The effective tunneling is experimentally estimated in a Bose-Einstein condensate (BEC) by measuring the suppression of diffusion. (<bold>b</bold>) If the driving force is asymmetric in time, this lattice distortion leads to a change in the energy bands, which amounts to the introduction of Peierls phases [<xref ref-type="bibr" rid="B717-technologies-12-00064">717</xref>] to the effective tunneling elements. By controlling the amplitude of the drive, the positions of the minima in the quasi-momentum distribution can be adjusted, allowing the creation of condensates with finite momenta, as observed in experimental observations. (<bold>c</bold>) By combining field gradients with the imprinting of Peierls phases, it becomes possible to engineer intriguing plaquette models with controlled effective flux. Figure from [<xref ref-type="bibr" rid="B718-technologies-12-00064">718</xref>].</p>
Full article ">Figure 19
<p>Diagram of a quantum measurement and feedback protocol. The AMO platform, in this case atoms in an optical cavity, is weakly probed via some homodyne-like measurement. The outcome is classically processed by a computer that modifies the feedback controller accordingly in order to steer the state to the desired phase.</p>
Full article ">Figure 20
<p>Cross-section of a neutron star. The neutron superfluid in the crust and the proton superfluid in the core are expected to be paired in <sup>1</sup><italic>S</italic><sub>0</sub> states, and the neutron superfluid in the core in <sup>3</sup><italic>P</italic><sub>2</sub> states. The electrons are instead expected to be in a normal state. A quark superfluid in the core is also expected to be BCS paired in <sup>1</sup><italic>S</italic><sub>0</sub> states. Figure re-drawn from [<xref ref-type="bibr" rid="B843-technologies-12-00064">843</xref>].</p>
Full article ">Figure 21
<p>Qualitative QCD phase diagram for varying temperature <italic>T</italic>, baryonic chemical potential <inline-formula><mml:math id="mm470"><mml:semantics><mml:mi>μ</mml:mi></mml:semantics></mml:math></inline-formula>, and isospin chemical potential <inline-formula><mml:math id="mm471"><mml:semantics><mml:msub><mml:mi>μ</mml:mi><mml:mi>I</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> (see text). Reported are the order parameters identifying the different phases. They are expressed as vacuum expectation values of bilinears of the field <inline-formula><mml:math id="mm472"><mml:semantics><mml:mrow><mml:mo>Ψ</mml:mo><mml:mo>=</mml:mo><mml:mo>(</mml:mo><mml:msub><mml:mi>ψ</mml:mi><mml:mi>u</mml:mi></mml:msub><mml:mo>,</mml:mo><mml:msub><mml:mi>ψ</mml:mi><mml:mi>d</mml:mi></mml:msub><mml:mo>)</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula>, where <inline-formula><mml:math id="mm473"><mml:semantics><mml:msub><mml:mi>ψ</mml:mi><mml:mi>u</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm474"><mml:semantics><mml:msub><mml:mi>ψ</mml:mi><mml:mi>d</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> are fields for the up and down quarks, <italic>C</italic> denotes the charge-conjugation unitary matrix acting on each of them, <inline-formula><mml:math id="mm475"><mml:semantics><mml:msub><mml:mi>σ</mml:mi><mml:mn>2</mml:mn></mml:msub></mml:semantics></mml:math></inline-formula> is the second Pauli matrix acting on the color (u, d) indexes, and <inline-formula><mml:math id="mm476"><mml:semantics><mml:msub><mml:mi>γ</mml:mi><mml:mn>5</mml:mn></mml:msub></mml:semantics></mml:math></inline-formula> denotes the fifth Dirac matrix. The bar over the field operators indicates the usual Lorentz inverse conjugation. For a detailed description of the order parameters, see [<xref ref-type="bibr" rid="B807-technologies-12-00064">807</xref>,<xref ref-type="bibr" rid="B809-technologies-12-00064">809</xref>,<xref ref-type="bibr" rid="B810-technologies-12-00064">810</xref>,<xref ref-type="bibr" rid="B819-technologies-12-00064">819</xref>].</p>
Full article ">Figure 22
<p>Gravitational Waves detection with atom technologies. Comparative analysis of strain measurements proposed by AEDGE with those proposed or realised by light-interferometry experiments as in the legend. Sensitivities to black hole (BH) mergers at different redshifts <italic>z</italic> and with varying total masses are showcased. Lines: predicted strain signals for binary BH mergers, with equal mass (solid lines) and significantly different masses (dashed lines), specifically 3000 <inline-formula><mml:math id="mm477"><mml:semantics><mml:msub><mml:mi>M</mml:mi><mml:mo>⊙</mml:mo></mml:msub></mml:semantics></mml:math></inline-formula> and 30 <inline-formula><mml:math id="mm478"><mml:semantics><mml:msub><mml:mi>M</mml:mi><mml:mo>⊙</mml:mo></mml:msub></mml:semantics></mml:math></inline-formula>. Shown is also the estimated level of gravitational gradient noise (GGN), that could potentially arise in a terrestrial detector of kilometer scale: this emphasizes the need for effective mitigation strategies. Notice that potential synergistic collaborations are possible between AEDGE and other detectors investigating different stages and histories of BH mergers. Image from [<xref ref-type="bibr" rid="B50-technologies-12-00064">50</xref>] (Permission to use this content is granted by Creative Commons Licence <uri>https://creativecommons.org/licenses/by/4.0/</uri>).</p>
Full article ">Figure 23
<p>Quest for Dark Matter with atom technologies. Comparison between current experiment sensitivity to a linear coupling of scalar ULDM to quarks (shaded region) and the sensitivity proposed by STE−QUEST mission (dashed line). Experiments include atomic clocks [<xref ref-type="bibr" rid="B927-technologies-12-00064">927</xref>], the MICROSCOPE experiment obtained from [<xref ref-type="bibr" rid="B904-technologies-12-00064">904</xref>] and torsion balances [<xref ref-type="bibr" rid="B928-technologies-12-00064">928</xref>], as in the legend. Image from [<xref ref-type="bibr" rid="B52-technologies-12-00064">52</xref>] (Permission to use this content granted by Creative Commons Licence <uri>https://creativecommons.org/licenses/by/4.0/</uri>).</p>
Full article ">Figure 24
<p>Analogue gravity simulation with atom technologies of Hawking radiation in acoustic holes. (<bold>a</bold>) predicted density−density correlation function <inline-formula><mml:math id="mm479"><mml:semantics><mml:mrow><mml:mrow><mml:mo>(</mml:mo><mml:mi>n</mml:mi><mml:mi>ξ</mml:mi><mml:mn>1</mml:mn><mml:mo>)</mml:mo></mml:mrow><mml:mo>×</mml:mo><mml:mo>[</mml:mo><mml:msup><mml:mi>G</mml:mi><mml:mrow><mml:mo>(</mml:mo><mml:mn>2</mml:mn><mml:mo>)</mml:mo></mml:mrow></mml:msup><mml:mrow><mml:mo>(</mml:mo><mml:mi>x</mml:mi><mml:mo>,</mml:mo><mml:msup><mml:mi>x</mml:mi><mml:mo>′</mml:mo></mml:msup><mml:mo>)</mml:mo></mml:mrow><mml:mo>−</mml:mo><mml:mn>1</mml:mn><mml:mo>]</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula> between Hawking particles and their partners, rescaled to the diluteness parameter <inline-formula><mml:math id="mm480"><mml:semantics><mml:mrow><mml:mi>n</mml:mi><mml:msub><mml:mi>ξ</mml:mi><mml:mn>1</mml:mn></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula>, with <italic>n</italic> the density and <inline-formula><mml:math id="mm481"><mml:semantics><mml:mrow><mml:msub><mml:mi>ξ</mml:mi><mml:mn>1</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:msqrt><mml:mrow><mml:msup><mml:mo>ℏ</mml:mo><mml:mn>2</mml:mn></mml:msup><mml:mo>/</mml:mo><mml:mrow><mml:mo>(</mml:mo><mml:mi>m</mml:mi><mml:msub><mml:mi>μ</mml:mi><mml:mn>1</mml:mn></mml:msub><mml:mo>)</mml:mo></mml:mrow></mml:mrow></mml:msqrt></mml:mrow></mml:semantics></mml:math></inline-formula> the healing length in terms of the atom mass <italic>m</italic> and the chemical potential <inline-formula><mml:math id="mm482"><mml:semantics><mml:msub><mml:mi>μ</mml:mi><mml:mn>1</mml:mn></mml:msub></mml:semantics></mml:math></inline-formula>. The temperature is <inline-formula><mml:math id="mm483"><mml:semantics><mml:mrow><mml:msub><mml:mi>k</mml:mi><mml:mi>B</mml:mi></mml:msub><mml:msub><mml:mi>T</mml:mi><mml:mn>0</mml:mn></mml:msub><mml:mo>/</mml:mo><mml:msub><mml:mi>μ</mml:mi><mml:mn>1</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>0.1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>. (<bold>b</bold>) the same quantity as in (<bold>a</bold>), in the absence of a black−hole horizon, where the flow remains sub-sonic everywhere. Image from [<xref ref-type="bibr" rid="B959-technologies-12-00064">959</xref>] (Permission to use this content granted by Creative Commons Licence <uri>https://creativecommons.org/licenses/by/4.0/</uri>). (<bold>c</bold>) Experimental observation of the predicted two−body correlation function in (<bold>a</bold>) panel, performed in [<xref ref-type="bibr" rid="B39-technologies-12-00064">39</xref>]. The origin represents the location of the horizon, and the dark bands emerging from the horizon indicate correlations. Image from [<xref ref-type="bibr" rid="B39-technologies-12-00064">39</xref>].</p>
Full article ">Figure 25
<p>Testing the foundations of quantum mechanics with atom technologies. Test of the phenomenological Continuous Spontaneous Localization (CSL) model for the collapse of the wavefunction, in the governing parameters space: the collapse rate <inline-formula><mml:math id="mm484"><mml:semantics><mml:mi>λ</mml:mi></mml:semantics></mml:math></inline-formula> and the correlation length <inline-formula><mml:math id="mm485"><mml:semantics><mml:msub><mml:mi>r</mml:mi><mml:mi>c</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula> of the collapse noise (see text). Diversely colored areas indicate the excluded regions based on results from different experiments: ground-based interferometric experiments (pink) [<xref ref-type="bibr" rid="B988-technologies-12-00064">988</xref>,<xref ref-type="bibr" rid="B1020-technologies-12-00064">1020</xref>,<xref ref-type="bibr" rid="B1021-technologies-12-00064">1021</xref>], non-interferometric experiments (blue) [<xref ref-type="bibr" rid="B1022-technologies-12-00064">1022</xref>,<xref ref-type="bibr" rid="B1023-technologies-12-00064">1023</xref>,<xref ref-type="bibr" rid="B1024-technologies-12-00064">1024</xref>,<xref ref-type="bibr" rid="B1025-technologies-12-00064">1025</xref>,<xref ref-type="bibr" rid="B1026-technologies-12-00064">1026</xref>], cold-atoms on the ground (orange) [<xref ref-type="bibr" rid="B1027-technologies-12-00064">1027</xref>], and non-interferometric experiments in space (green) [<xref ref-type="bibr" rid="B1022-technologies-12-00064">1022</xref>,<xref ref-type="bibr" rid="B1028-technologies-12-00064">1028</xref>]. Red line: the STE-QUEST prediction. Grey area: theoretically excluded region assuming a collapse at the macroscopic scale, which is a fundamental requirement of the model [<xref ref-type="bibr" rid="B1021-technologies-12-00064">1021</xref>]. Black dot: prediction of the Ghirardi-Rimini-Weber (GRW) model [<xref ref-type="bibr" rid="B1018-technologies-12-00064">1018</xref>]. Black interval: Adler model [<xref ref-type="bibr" rid="B1019-technologies-12-00064">1019</xref>]. Image from [<xref ref-type="bibr" rid="B52-technologies-12-00064">52</xref>] (Permission to use this content granted by Creative Commons Licence <uri>https://creativecommons.org/licenses/by/4.0/</uri>).</p>
Full article ">Figure 26
<p>Quantum-like paradigm and the brain. Quantum spin model for visual neurosciences: a network of neurons or groups of neurons is mapped into an open quantum chain of spins with given connectivity. (<bold>a</bold>) Different connectivities: nearest-neighbour (left) and all−to−all (right); (<bold>b</bold>) Excitations propagate via exchange <italic>J</italic>; (<bold>c</bold>) nearby excitations experience an energy offset <inline-formula><mml:math id="mm486"><mml:semantics><mml:msub><mml:mi mathvariant="sans-serif">Δ</mml:mi><mml:mn>0</mml:mn></mml:msub></mml:semantics></mml:math></inline-formula>; and (<bold>d</bold>) can be subject to different dissipative channels at rate <inline-formula><mml:math id="mm487"><mml:semantics><mml:msub><mml:mi>γ</mml:mi><mml:mi>i</mml:mi></mml:msub></mml:semantics></mml:math></inline-formula>; (<bold>e</bold>) Excitations of variable amplitude are injected via spin flip operations. (<bold>f</bold>) Magnetisation profile of one single excitation in nearest−neighbor (n.n., top) and all−to−all (a.a., bottom) connectivity. (<bold>g</bold>) Magnetisation propagations in the a.a. case for <inline-formula><mml:math id="mm488"><mml:semantics><mml:mrow><mml:mi>N</mml:mi><mml:mo>=</mml:mo><mml:mn>2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (top) and <inline-formula><mml:math id="mm489"><mml:semantics><mml:mrow><mml:mi>N</mml:mi><mml:mo>=</mml:mo><mml:mn>3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (bottom) excitations, the regular interference generates clear patterns with additional frequencies as more spin flips are introduced. This can be understood in the power spectrum of the time signals (<bold>h</bold>). (<bold>i</bold>) Weber’s law for numerosity is recovered using an ideal−observer decoding procedure independently of whether the excitation is injected with random amplitude (RR) or random amplitudes constrained to constant total energy (CE). Adapted from [<xref ref-type="bibr" rid="B1052-technologies-12-00064">1052</xref>] (Permission to use this content granted by Creative Commons Licence <uri>https://creativecommons.org/licenses/by/4.0/</uri>).</p>
Full article ">Figure 27
<p>Culturo-scientific storytelling (CSS) approach to Physics Outreach Research. (<bold>a</bold>) CSS consists in an experiential journey that mirrors the encounters of scientists and citizens, with each subject being approached through the process of scientific thinking, as depicted in (<bold>b</bold>). The journey begins by acquainting oneself with peripheral knowledge (arrow 1) and gradually delves deeper into the consolidation of fundamental concepts at the core (arrow 2). Practical applications are then explored and comprehended in the main body (arrow 3), followed by a return to the periphery to emphasize the ever-evolving and unfinished nature of the discipline-culture (arrow 4). Image inspired from [<xref ref-type="bibr" rid="B61-technologies-12-00064">61</xref>].</p>
Full article ">Figure 28
<p>Interactive tools for CSS. Screenshots from the art and science 6 sqm installation Quantum Jungle by Robin Baumgarten, visualizing the time evolution of one quantum particle. After touching the springs, the quantum particle is created in a superposition state (<bold>top left</bold>). The probability from computer simulations of the Schrödinger equation evolves according to a quantum walk (<bold>top right</bold>, <bold>bottom left</bold>), and is visualized via switching on of LEDs with proportional intensity. Later spring touch is interpreted as a measurement action, visualizing the collapse (<bold>bottom right</bold>).</p>
Full article ">
20 pages, 2332 KiB  
Article
Dynamic Behavior and Bifurcation Analysis of a Modified Reduced Lorenz Model
by Mohammed O. Al-Kaff, Ghada AlNemer, Hamdy A. El-Metwally, Abd-Elalim A. Elsadany and Elmetwally M. Elabbasy
Mathematics 2024, 12(9), 1354; https://doi.org/10.3390/math12091354 - 29 Apr 2024
Viewed by 856
Abstract
This study introduces a newly modified Lorenz model capable of demonstrating bifurcation within a specified range of parameters. The model demonstrates various bifurcation behaviors, which are depicted as distinct structures in the diagram. The study aims to discover and analyze the existence and [...] Read more.
This study introduces a newly modified Lorenz model capable of demonstrating bifurcation within a specified range of parameters. The model demonstrates various bifurcation behaviors, which are depicted as distinct structures in the diagram. The study aims to discover and analyze the existence and stability of fixed points in the model. To achieve this, the center manifold theorem and bifurcation theory are employed to identify the requirements for pitchfork bifurcation, period-doubling bifurcation, and Neimark–Sacker bifurcation. In addition to theoretical findings, numerical simulations, including bifurcation diagrams, phase pictures, and maximum Lyapunov exponents, showcase the nuanced, complex, and diverse dynamics. Finally, the study applies the Ott–Grebogi–Yorke (OGY) method to control the chaos observed in the reduced modified Lorenz model. Full article
(This article belongs to the Special Issue Applied Mathematics in Nonlinear Dynamics and Chaos)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>4.04</mn> <mo>,</mo> <mi>h</mi> <mo>∈</mo> <mo>[</mo> <mn>0.87</mn> <mo>,</mo> <mn>0.929</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>9</mn> <mo>,</mo> <mi>h</mi> <mo>∈</mo> <mo>[</mo> <mn>0.25</mn> <mo>,</mo> <mn>0.35</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>=</mo> <mn>0.8874794</mn> <mo>,</mo> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>4</mn> <mo>,</mo> <mn>4.36</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>=</mo> <mn>0.2519820562</mn> <mo>,</mo> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>8.5</mn> <mo>,</mo> <mn>12.34</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.6</mn> <mo>,</mo> <mi>h</mi> <mo>∈</mo> <mo>[</mo> <mn>1.49</mn> <mo>,</mo> <mn>2</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>The phase pictures associated with <a href="#mathematics-12-01354-f005" class="html-fig">Figure 5</a>a,b.</p>
Full article ">Figure 6 Cont.
<p>The phase pictures associated with <a href="#mathematics-12-01354-f005" class="html-fig">Figure 5</a>a,b.</p>
Full article ">Figure 7
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>2.1</mn> <mo>,</mo> <mi>h</mi> <mo>∈</mo> <mo>[</mo> <mn>0.96</mn> <mo>,</mo> <mn>1.12</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>The phase pictures associated with <a href="#mathematics-12-01354-f007" class="html-fig">Figure 7</a>a,b.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>=</mo> <mn>0.9336269198</mn> <mo>,</mo> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>2.6</mn> <mo>,</mo> <mn>2.91</mn> <mo>]</mo> </mrow> </semantics></math> and (<b>d</b>) LA for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>2.8</mn> <mo>,</mo> <mn>2.91</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>The phase pictures associated with <a href="#mathematics-12-01354-f009" class="html-fig">Figure 9</a>a,b.</p>
Full article ">Figure 11
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>=</mo> <mn>1.495455526</mn> <mo>,</mo> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>0.5</mn> <mo>,</mo> <mn>1.21</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>(<b>a</b>,<b>b</b>) Bifurcation diagram and (<b>c</b>) ML of model (<a href="#FD4-mathematics-12-01354" class="html-disp-formula">4</a>) for value of <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>=</mo> <mn>0.9645336217</mn> <mo>,</mo> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>1.98</mn> <mo>,</mo> <mn>2.81</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>(<b>a</b>) The region of stability for the controlled model (<a href="#FD36-mathematics-12-01354" class="html-disp-formula">36</a>). (<b>b</b>) Diagrams of bifurcation for the controlled model (<a href="#FD36-mathematics-12-01354" class="html-disp-formula">36</a>) with <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>p</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mover accent="true"> <mi>ϰ</mi> <mo stretchy="false">^</mo> </mover> <mo>,</mo> <mover accent="true"> <mi>γ</mi> <mo stretchy="false">^</mo> </mover> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.482548585</mn> <mo>,</mo> <mn>0.2328531368</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mn>2</mn> </msub> <mo>∈</mo> <mrow> <mo>[</mo> <mo>−</mo> <mn>3.5</mn> <mo>,</mo> <mn>1</mn> <mo>]</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 14
<p>The Phase of Chaos of the controlled model (<a href="#FD36-mathematics-12-01354" class="html-disp-formula">36</a>).</p>
Full article ">
16 pages, 7351 KiB  
Article
Study of the Spatiotemporal Distribution Characteristics of Rainfall Using Hybrid Dimensionality Reduction-Clustering Model: A Case Study of Kunming City, China
by Weijie Lin, Yuanyuan Liu, Na Li, Jing Wang, Nianqiang Zhang, Yanyan Wang, Mingyang Wang, Hancheng Ren and Min Li
Atmosphere 2024, 15(5), 534; https://doi.org/10.3390/atmos15050534 - 26 Apr 2024
Viewed by 767
Abstract
In recent years, the frequency and intensity of global extreme weather events have gradually increased, leading to significant changes in urban rainfall patterns. The uneven distribution of rainfall has caused varying degrees of water security issues in different regions. Accurately grasping the spatiotemporal [...] Read more.
In recent years, the frequency and intensity of global extreme weather events have gradually increased, leading to significant changes in urban rainfall patterns. The uneven distribution of rainfall has caused varying degrees of water security issues in different regions. Accurately grasping the spatiotemporal distribution patterns of rainfall is crucial for understanding the hydrological cycle and predicting the availability of water resources. This study collected rainfall data every five minutes from 62 rain gauge stations in the main urban area of Kunming City from 2019 to 2021, constructing an unsupervised hybrid dimensionality reduction-clustering (HDRC) model. The model employs the Locally Linear Embedding (LLE) algorithm from manifold learning for dimensionality reduction of the data samples and uses the dynamic clustering K-Means algorithm for cluster analysis. The results show that the model categorizes the rainfall in the Kunming area into three types: The first type has its rainfall center distributed on the north shore of Dian Lake and the southern part of Kunming’s main urban area, with spatial dynamics showing the rainfall distribution gradually developing from the Dian Lake water body towards the land. The second type’s rainfall center is located in the northern mountainous area of Kunming, with a smaller spatial dynamic change trend. The water vapor has a relatively fixed and concentrated rainfall center due to the orographic uplift effect of the mountains. The third type’s rainfall center is located in the main urban area of Kunming, with this type of rainfall showing smaller variations in all indicators, mainly occurring in May and September when the temperature is lower, related to the urban heat island effect. This research provides a general workflow for spatial rainfall classification, capable of mining the spatiotemporal distribution patterns of regional rainfall based on extensive data and generating typical samples of rainfall types. Full article
(This article belongs to the Special Issue Characteristics of Extreme Climate Events over China)
Show Figures

Figure 1

Figure 1
<p>Flow chart of the analysis of spatial and temporal distribution characteristics of rainfall.</p>
Full article ">Figure 2
<p>Study scope of the rainfall area.</p>
Full article ">Figure 3
<p>Low-dimensional sample set clustering visualization.</p>
Full article ">Figure 4
<p>Cumulative rainfall distribution for Type I rainfall.</p>
Full article ">Figure 5
<p>Cumulative rainfall distribution for Type Ⅱ rainfall.</p>
Full article ">Figure 6
<p>Cumulative rainfall distribution for Type Ⅲ rainfall.</p>
Full article ">Figure 7
<p>Cumulative rainfall ratios for different regions for three different types of rainfall: (<b>a</b>) Type I rainfall; (<b>b</b>) Type II rainfall; and (<b>c</b>) Type III rainfall.</p>
Full article ">
17 pages, 4906 KiB  
Article
Sensing at the Nanoscale Using Nitrogen-Vacancy Centers in Diamond: A Model for a Quantum Pressure Sensor
by Hari P. Paudel, Gary R. Lander, Scott E. Crawford and Yuhua Duan
Nanomaterials 2024, 14(8), 675; https://doi.org/10.3390/nano14080675 - 12 Apr 2024
Viewed by 1496
Abstract
The sensing of stress under harsh environmental conditions with high resolution has critical importance for a range of applications including earth’s subsurface scanning, geological CO2 storage monitoring, and mineral and resource recovery. Using a first-principles density functional theory (DFT) approach combined with [...] Read more.
The sensing of stress under harsh environmental conditions with high resolution has critical importance for a range of applications including earth’s subsurface scanning, geological CO2 storage monitoring, and mineral and resource recovery. Using a first-principles density functional theory (DFT) approach combined with the theoretical modelling of the low-energy Hamiltonian, here, we investigate a novel approach to detect unprecedented levels of pressure by taking advantage of the solid-state electronic spin of nitrogen-vacancy (NV) centers in diamond. We computationally explore the effect of strain on the defect band edges and band gaps by varying the lattice parameters of a diamond supercell hosting a single NV center. A low-energy Hamiltonian is developed that includes the effect of stress on the energy level of a ±1 spin manifold at the ground state. By quantifying the energy level shift and split, we predict pressure sensing of up to 0.3 MPa/Hz using the experimentally measured spin dephasing time. We show the superiority of the quantum sensing approach over traditional optical sensing techniques by discussing our results from DFT and theoretical modelling for the frequency shift per unit pressure. Importantly, we propose a quantum manometer that could be useful to measure earth’s subsurface vibrations as well as for pressure detection and monitoring in high-temperature superconductivity studies and in material sciences. Our results open avenues for the development of a sensing technology with high sensitivity and resolution under extreme pressure limits that potentially has a wider applicability than the existing pressure sensing technologies. Full article
(This article belongs to the Special Issue First-Principle Calculation Study of Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>A diamond structure with a NV center. Shown are (<b>a</b>) the unit cell and (<b>b</b>) a 3 × 3 × 3 supercell structure with dimensions of 1.07 nm. The DOS for NV<sup>−</sup> defective diamond. (<b>c</b>) The impurity states can be seen around the Fermi level with some contributions from C. (<b>d</b>). The charge distribution around the N impurity and C vacancy centered around the supercell with charge <math display="inline"><semantics> <mrow> <mi>q</mi> <mo>=</mo> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math>. The color map in (<b>d</b>) indicates positive (yellow) and negative (blue) charge surfaces.</p>
Full article ">Figure 2
<p>Diamond bandstructures calculated using the PAW-PBE potential (<b>a</b>) with a N impurity, (<b>b</b>) NV<sup>0</sup>, and (<b>c</b>) with NV<sup>−</sup> centers in a 3 × 3 × 3 supercell for a NV center oriented along the [111] direction. As a NV center was introduced, additional defect bands arose with a band gap of nearly 1.5 eV in this calculation, allowing for strong electron-hole polarizations under green laser illumination in both neutral and negative NV defects. Dark blue denotes the C states, whereas red and light blue indicate impurity states (zero energy is placed at the Fermi level).</p>
Full article ">Figure 3
<p>NV-center diamond band structures for changes in the transverse lattice parameter ratio <math display="inline"><semantics> <mrow> <mfrac> <mrow> <msup> <mrow> <mi>a</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msup> </mrow> <mrow> <mi>a</mi> </mrow> </mfrac> <mo>=</mo> <mfrac> <mrow> <msup> <mrow> <mi>b</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msup> </mrow> <mrow> <mi>b</mi> </mrow> </mfrac> </mrow> </semantics></math> by (<b>a</b>) 1.0074, (<b>b</b>) 1.014, (<b>c</b>) 0.988 and (<b>d</b>) 0.9813. The blue curves represent the C bands. whereas the red curves represent the impurity band.</p>
Full article ">Figure 4
<p>Energy splitting of the excited <sup>3</sup>E level under a variation in the ratios of the longitudinal (<b>a</b>) and transverse (<b>b</b>) lattice parameters. The transverse lattice parameters a and b were changed simultaneously in (<b>b</b>) up to 2%. Only the change in <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>a</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msup> <mo>/</mo> <mi>a</mi> </mrow> </semantics></math> is shown, due to the cubic symmetry of the supercell.</p>
Full article ">Figure 5
<p>NV center in a diamond unit cell showing coordinate systems with respect to NV axes (red) and crystallographic axes (green) [<a href="#B63-nanomaterials-14-00675" class="html-bibr">63</a>].</p>
Full article ">Figure 6
<p>Energy level shift (blue) and splitting (red and green) of the <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>1</mn> </mrow> </semantics></math> spin manifold as a function of the applied stress along the <math display="inline"><semantics> <mrow> <mi mathvariant="bold-italic">p</mi> <mo>∥</mo> <mfenced open="[" close="]" separators="|"> <mrow> <mn>100</mn> </mrow> </mfenced> </mrow> </semantics></math> direction for the NV centers with orientation <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">e</mi> </mrow> <mrow> <mi>z</mi> </mrow> </msub> <mo>∈</mo> <mo>{</mo> <mn>111</mn> <mo>}</mo> </mrow> </semantics></math> in a zero magnetic field. Zero-field splitting is aligned with the zero energy.</p>
Full article ">
22 pages, 699 KiB  
Article
Mathematical Modeling of Two Interacting Populations’ Dynamics of Onchocerciasis Disease Spread with Nonlinear Incidence Functions
by Kabiru Michael Adeyemo, Umar Muhammad Adam, Adejimi Adeniji and Kayode Oshinubi
Mathematics 2024, 12(2), 222; https://doi.org/10.3390/math12020222 - 9 Jan 2024
Viewed by 796
Abstract
The transmission dynamics of onchocerciasis in two interacting populations are examined using a deterministic compartmental model with nonlinear incidence functions. The model undergoes qualitative analysis to examine how it behaves near disease-free equilibrium (DFE) and endemic equilibrium. Using the Lyapunov function, it is [...] Read more.
The transmission dynamics of onchocerciasis in two interacting populations are examined using a deterministic compartmental model with nonlinear incidence functions. The model undergoes qualitative analysis to examine how it behaves near disease-free equilibrium (DFE) and endemic equilibrium. Using the Lyapunov function, it is demonstrated that the DFE is globally stable when the threshold parameter R01 is taken into account. When R0>1, it suffices to show globally how asymptotically stable the endemic equilibrium is and its existence. We conduct the bifurcation analysis by looking at the possibility of the model’s equilibria coexisting at R0<1 but near R0=1 using the Center Manifold Theory. We use the sensitivity analysis method to understand how some parameters influence the R0, hence the transmission and mitigation of the disease dynamics. Furthermore, we simulate the model developed numerically to understand the population dynamics. The outcome presented in this article offers valuable understanding of the transmission dynamics of onchocerciasis, specifically in the context of two populations that interact with each other, considering the presence of nonlinear incidence. Full article
Show Figures

Figure 1

Figure 1
<p>Forward Bifurcation plot of the Onchocerciasis model.</p>
Full article ">Figure 2
<p>Simulation of the population state with varying parameters of onchocerciasis model dynamics, (<b>left</b>): <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mi>h</mi> </msub> <mo>=</mo> <mn>0.09</mn> <mo>,</mo> <msub> <mi>γ</mi> <mi>h</mi> </msub> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mi>r</mi> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, (<b>right</b>): <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mi>h</mi> </msub> <mo>=</mo> <mn>0.05</mn> <mo>,</mo> <msub> <mi>γ</mi> <mi>h</mi> </msub> <mo>=</mo> <mn>0.09</mn> <mo>,</mo> <mi>r</mi> <mo>=</mo> <mn>0.0667</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Susceptible and exposed human population. (<b>a</b>) Dynamic behavior of susceptible human population with varying parameters and (<b>b</b>) Dynamic behavior of exposed human population with varying parameters.</p>
Full article ">Figure 4
<p>Infected and recovered human population. (<b>a</b>) Dynamic behavior of infected human population with varying parameters and (<b>b</b>) Dynamic behavior of recovered human population with varying parameters.</p>
Full article ">Figure 5
<p>Graph of parameters and their sensitivity indices.</p>
Full article ">Figure 6
<p>Graph of parameters and their sensitivity indices.</p>
Full article ">
23 pages, 1490 KiB  
Article
Spatiotemporal Dynamics of a Diffusive Immunosuppressive Infection Model with Nonlocal Competition and Crowley–Martin Functional Response
by Yuan Xue, Jinli Xu and Yuting Ding
Axioms 2023, 12(12), 1085; https://doi.org/10.3390/axioms12121085 - 27 Nov 2023
Viewed by 1052
Abstract
In this paper, we introduce the Crowley–Martin functional response and nonlocal competition into a reaction–diffusion immunosuppressive infection model. First, we analyze the existence and stability of the positive constant steady states of the systems with nonlocal competition and local competition, respectively. Second, we [...] Read more.
In this paper, we introduce the Crowley–Martin functional response and nonlocal competition into a reaction–diffusion immunosuppressive infection model. First, we analyze the existence and stability of the positive constant steady states of the systems with nonlocal competition and local competition, respectively. Second, we deduce the conditions for the occurrence of Turing, Hopf, and Turing–Hopf bifurcations of the system with nonlocal competition, as well as the conditions for the occurrence of Hopf bifurcations of the system with local competition. Furthermore, we employ the multiple time scales method to derive the normal forms of the Hopf bifurcations reduced on the center manifold for both systems. Finally, we conduct numerical simulations for both systems under the same parameter settings, compare the impact of nonlocal competition, and find that the nonlocal term can induce spatially inhomogeneous stable periodic solutions. We also provide corresponding biological explanations for the simulation results. Full article
(This article belongs to the Special Issue Recent Advances in Applied Mathematics and Artificial Intelligence)
Show Figures

Figure 1

Figure 1
<p>If <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math> holds, then system (<a href="#FD2-axioms-12-01085" class="html-disp-formula">2</a>) has three positive constant steady states.</p>
Full article ">Figure 2
<p>System (<a href="#FD2-axioms-12-01085" class="html-disp-formula">2</a>) has two positive constant steady states under various conditions. (<b>a</b>,<b>b</b>) Correspond to when <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>3</mn> <mo>)</mo> </mrow> </semantics></math> holds. (<b>c</b>) Corresponds to when <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>4</mn> <mo>)</mo> </mrow> </semantics></math> holds. (<b>d</b>) Corresponds to when <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>2</mn> <mo>)</mo> </mrow> </semantics></math> holds.</p>
Full article ">Figure 3
<p>System (<a href="#FD2-axioms-12-01085" class="html-disp-formula">2</a>) has one positive constant steady state under various conditions. (<b>a</b>,<b>b</b>) Correspond to when <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>5</mn> <mo>)</mo> </mrow> </semantics></math> holds. (<b>c</b>,<b>d</b>) Correspond to when <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>6</mn> <mo>)</mo> </mrow> </semantics></math> holds. (<b>e</b>) Corresponds to when <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>7</mn> <mo>)</mo> </mrow> </semantics></math> holds. (<b>f</b>) Left corresponds to when <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>8</mn> <mo>)</mo> </mrow> </semantics></math> holds, and right corresponds to when <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="normal">H</mi> <mn>9</mn> <mo>)</mo> </mrow> </semantics></math> holds.</p>
Full article ">Figure 4
<p>Choose <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0.1</mn> <mo>&lt;</mo> <mn>0.1128</mn> </mrow> </semantics></math> and the initial function is (<math display="inline"><semantics> <mrow> <mn>1.337684786</mn> <mo>+</mo> <mn>0.01</mn> <mo form="prefix">cos</mo> <mi>x</mi> <mo>,</mo> <mspace width="3.33333pt"/> <mn>3.063620649</mn> <mo>+</mo> <mn>0.025</mn> <mo form="prefix">cos</mo> <mi>x</mi> </mrow> </semantics></math>). Then the positive constant steady state <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mo>∗</mo> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mn>1.337684786</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mn>3.063620649</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> of system (<a href="#FD2-axioms-12-01085" class="html-disp-formula">2</a>) is locally asymptotically stable.</p>
Full article ">Figure 5
<p>Choose <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0.1165</mn> <mo>&gt;</mo> <mn>0.1128</mn> </mrow> </semantics></math> and consider the initial function as (<math display="inline"><semantics> <mrow> <mn>1.337684786</mn> <mo>+</mo> <mn>0.01</mn> <mo form="prefix">cos</mo> <mi>x</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>3.063620649</mn> <mo>+</mo> <mn>0.025</mn> <mo form="prefix">cos</mo> <mi>x</mi> </mrow> </semantics></math>). Then system (<a href="#FD2-axioms-12-01085" class="html-disp-formula">2</a>) exhibits a spatially inhomogeneous stable periodic solution.</p>
Full article ">Figure 6
<p>Choose <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0.12</mn> <mo>&gt;</mo> <mn>0.1128</mn> </mrow> </semantics></math> and consider the initial function as (<math display="inline"><semantics> <mrow> <mn>1.337684786</mn> <mo>+</mo> <mn>0.01</mn> <mo form="prefix">cos</mo> <mi>x</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>3.063620649</mn> <mo>+</mo> <mn>0.025</mn> <mo form="prefix">cos</mo> <mi>x</mi> </mrow> </semantics></math>). The constant steady state solution <math display="inline"><semantics> <msub> <mi>E</mi> <mo>∗</mo> </msub> </semantics></math> of system (<a href="#FD2-axioms-12-01085" class="html-disp-formula">2</a>) is unstable.</p>
Full article ">Figure 7
<p>Choose <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0.165</mn> <mo>&lt;</mo> <mn>0.1774</mn> </mrow> </semantics></math> and the initial function is (<math display="inline"><semantics> <mrow> <mn>1.337684786</mn> <mo>+</mo> <mn>0.003</mn> <mo>+</mo> <mn>0.005</mn> <mo form="prefix">cos</mo> <mi>x</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>3.063620649</mn> <mo>+</mo> <mn>0.001</mn> <mo form="prefix">cos</mo> <mi>x</mi> </mrow> </semantics></math>). Then the constant steady state <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mo>∗</mo> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mn>1.337684786</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mn>3.063620649</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> of system (<a href="#FD1-axioms-12-01085" class="html-disp-formula">1</a>) is locally asymptotically stable.</p>
Full article ">Figure 8
<p>Choose <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0.178</mn> <mo>&gt;</mo> <mn>0.1774</mn> </mrow> </semantics></math>, and consider the initial function as (<math display="inline"><semantics> <mrow> <mn>1.337684786</mn> <mo>+</mo> <mn>0.003</mn> <mo>+</mo> <mn>0.005</mn> <mo form="prefix">cos</mo> <mi>x</mi> <mo>,</mo> <mspace width="3.33333pt"/> <mn>3.063620649</mn> <mo>+</mo> <mn>0.001</mn> <mo form="prefix">cos</mo> <mi>x</mi> </mrow> </semantics></math>), then system (<a href="#FD1-axioms-12-01085" class="html-disp-formula">1</a>) exhibits a spatially homogeneous stable periodic solution.</p>
Full article ">
34 pages, 3811 KiB  
Article
Dynamics Analysis of a Discrete-Time Commensalism Model with Additive Allee for the Host Species
by Yanbo Chong, Ankur Jyoti Kashyap, Shangming Chen and Fengde Chen
Axioms 2023, 12(11), 1031; https://doi.org/10.3390/axioms12111031 - 2 Nov 2023
Cited by 1 | Viewed by 1154
Abstract
We propose and study a class of discrete-time commensalism systems with additive Allee effects on the host species. First, the single species with additive Allee effects is analyzed for existence and stability, then the existence of fixed points of discrete systems is given, [...] Read more.
We propose and study a class of discrete-time commensalism systems with additive Allee effects on the host species. First, the single species with additive Allee effects is analyzed for existence and stability, then the existence of fixed points of discrete systems is given, and the local stability of fixed points is given by characteristic root analysis. Second, we used the center manifold theorem and bifurcation theory to study the bifurcation of a codimension of one of the system at non-hyperbolic fixed points, including flip, transcritical, pitchfork, and fold bifurcations. Furthermore, this paper used the hybrid chaos method to control the chaos that occurs in the flip bifurcation of the system. Finally, the analysis conclusions were verified by numerical simulations. Compared with the continuous system, the similarities are that both species’ densities decrease with increasing Allee values under the weak Allee effect and that the host species hastens extinction under the strong Allee effect. Further, when the birth rate of the benefited species is low and the time is large enough, the benefited species will be locally asymptotically stabilized. Thus, our new finding is that both strong and weak Allee effects contribute to the stability of the benefited species under certain conditions. Full article
(This article belongs to the Special Issue Mathematical Models and Simulations)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">V</span> and <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi mathvariant="normal">V</mi> <mo>(</mo> <mi>y</mi> <mo>)</mo> </mrow> </semantics></math> function with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>2.5</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Positive fixed point existence diagram for map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>). (<b>a</b>) If <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>0.54</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has two positive fixed points <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mspace width="3.33333pt"/> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>. (<b>b</b>–<b>d</b>) Local diagrams with <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>0.5</mn> <mo>]</mo> <mo>,</mo> <mspace width="3.33333pt"/> <mi>y</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>0.1</mn> <mo>]</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>∈</mo> <mo>[</mo> <mn>0.1</mn> <mo>,</mo> <mn>0.5</mn> <mo>]</mo> </mrow> </semantics></math> corresponding to (<b>a</b>), respectively. (<b>e</b>) If <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>0.5625</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has a unique positive fixed point <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math>. (<b>f</b>) Local diagrams with <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>0.25</mn> <mo>]</mo> </mrow> </semantics></math> corresponding to (<b>e</b>), respectively. (<b>g</b>) If <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has no positive fixed point. (<b>h</b>) If <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>3</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) also has no positive fixed point.</p>
Full article ">Figure 3
<p>Flip bifurcation diagrams of commensal species at initial value <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>0</mn> </msub> <mo>,</mo> <msub> <mi>y</mi> <mn>0</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.5</mn> <mo>,</mo> <mn>0.3</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>4</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Topological classification at <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mi>α</mi> <mo>,</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>4</mn> <mo>]</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>3</mn> <mo>]</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Topological classification of the <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>−</mo> <msubsup> <mi>β</mi> <mrow> <mi>i</mi> </mrow> <mo>*</mo> </msubsup> </mrow> </semantics></math> plane at <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mrow> <mn>3</mn> <mi>i</mi> </mrow> </msub> <mrow> <mo>(</mo> <msubsup> <mi>x</mi> <mrow> <mi>i</mi> </mrow> <mo>*</mo> </msubsup> <mo>,</mo> <msub> <mi>y</mi> <mi>i</mi> </msub> <mo>)</mo> </mrow> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>0.52</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>0.45</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>4</mn> <mo>]</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msubsup> <mi>β</mi> <mrow> <mi>i</mi> </mrow> <mo>*</mo> </msubsup> <mo>∈</mo> <mrow> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>10</mn> <mo>]</mo> </mrow> <mo>,</mo> <mrow> <mo>(</mo> <mi>i</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>(<b>a</b>) With <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, and the initial value as <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mn>0.5</mn> <mo>,</mo> <mn>1</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has positive fixed points <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.5616</mn> </mrow> </semantics></math>; (<b>b</b>) with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> and the initial value as <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mn>0.5</mn> <mo>,</mo> <mn>1</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has positive fixed points <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>; (<b>c</b>) with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>0.54</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, and the initial value as <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0.09</mn> <mo>,</mo> <mn>0.05</mn> <mo>,</mo> <mn>0.02</mn> <mo>,</mo> <mn>0.2</mn> <mo>,</mo> <mn>0.3</mn> <mo>.</mo> <mn>0.5</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has two positive fixed points <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mspace width="3.33333pt"/> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>; (<b>d</b>) with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>0.5625</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, and the initial value as <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0.09</mn> <mo>,</mo> <mn>0.05</mn> <mo>,</mo> <mn>0.02</mn> <mo>,</mo> <mn>0.2</mn> <mo>,</mo> <mn>0.3</mn> <mo>.</mo> <mn>0.5</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has positive fixed points <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>e</b>) with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, and the initial value as <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0.15</mn> <mo>,</mo> <mn>0.3</mn> <mo>,</mo> <mn>1</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has no positive fixed point; (<b>f</b>) with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mspace width="3.33333pt"/> <mi>A</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, and the initial value as <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>=</mo> <mn>0.15</mn> <mo>,</mo> <mn>0.3</mn> <mo>,</mo> <mn>1</mn> </mrow> </semantics></math>, the map (<a href="#FD6-axioms-12-01031" class="html-disp-formula">6</a>) has no positive fixed point.</p>
Full article ">Figure 7
<p>Stability of the fixed point <math display="inline"><semantics> <msubsup> <mi>E</mi> <mrow> <mn>32</mn> </mrow> <mo>*</mo> </msubsup> </semantics></math> of the map (<a href="#FD5-axioms-12-01031" class="html-disp-formula">5</a>) with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mi>β</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>M</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, and the initial values as <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>x</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>,</mo> <mi>y</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mo>(</mo> <mn>1.5</mn> <mo>,</mo> <mn>0.3</mn> <mo>)</mo> <mo>,</mo> <mo>(</mo> <mn>1.2</mn> <mo>,</mo> <mn>0.6</mn> <mo>)</mo> <mo>,</mo> <mo>(</mo> <mn>0.8</mn> <mo>,</mo> <mn>0.2</mn> <mo>)</mo> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Transcritical bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>y</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>3</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>0.15</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Pitchfork bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>y</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>3</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>(<b>a</b>) Flip bifurcation diagrams on the <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>−</mo> <mi>x</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>M</mi> <mo>=</mo> <mn>0.4</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>; (<b>b</b>) maximum Lyapunov exponents corresponding to (<b>a</b>).</p>
Full article ">Figure 11
<p>Local bifurcation diagrams corresponding to <a href="#axioms-12-01031-f010" class="html-fig">Figure 10</a>a.</p>
Full article ">Figure 12
<p>Phase diagrams corresponding to <a href="#axioms-12-01031-f010" class="html-fig">Figure 10</a>a.</p>
Full article ">Figure 13
<p>(<b>a</b>) Bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>x</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>; (<b>b</b>) bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>y</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>; (<b>c</b>) bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>x</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>; (<b>d</b>) bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>y</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>; (<b>e</b>) bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>x</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>2.5</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>; (<b>f</b>) bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>x</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>2.5</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 14
<p>Fold bifurcation diagrams in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>y</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 15
<p>Time series of the commensal species for (<a href="#FD28-axioms-12-01031" class="html-disp-formula">28</a>) with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.5</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>0.5</mn> <mo>,</mo> <mi>M</mi> <mo>=</mo> <mn>0.2</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>. (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>=</mo> <mn>0.57</mn> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>=</mo> <mn>0.58</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>Stability region diagram in the <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>−</mo> <mi>ρ</mi> </mrow> </semantics></math> plane with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>2.5</mn> <mo>,</mo> <mi>β</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>A</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>.</p>
Full article ">
Back to TopTop