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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (399)

Search Parameters:
Keywords = tensor space

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 1411 KiB  
Article
Covariant Formulation of the Brain’s Emerging Ohm’s Law
by Manuel Rivas and Manuel Reina
Symmetry 2024, 16(12), 1570; https://doi.org/10.3390/sym16121570 (registering DOI) - 23 Nov 2024
Viewed by 218
Abstract
It is essential to establish the validity of Ohm’s law in any reference frame if we aim to implement a relativistic approach to brain dynamics based on a Lorentz covariant microscopic response relation. Here, we obtain a covariant formulation of Ohm’s law for [...] Read more.
It is essential to establish the validity of Ohm’s law in any reference frame if we aim to implement a relativistic approach to brain dynamics based on a Lorentz covariant microscopic response relation. Here, we obtain a covariant formulation of Ohm’s law for an electromagnetic field tensor of any order derived from the emergent conductivity tensor in highly non-isotropic systems, employing the bidomain theory framework within brain tissue cells. With this, we offer a different perspective that we hope will lead to understanding the close relationship between brain dynamics and a seemingly ordinary yet profoundly crucial element: space. Full article
(This article belongs to the Section Mathematics)
Show Figures

Figure 1

Figure 1
<p>Trajectory of a charged particle under the action of a uniform magnetic field. If the particle’s velocity is not perpendicular to the magnetic field, the particle will move in a helical path.</p>
Full article ">Figure 2
<p>The definition of the parallel transport of vector <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>V</mi> </mrow> <mo>→</mo> </mover> </mrow> </semantics></math> along the tangent vector <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>U</mi> </mrow> <mo>→</mo> </mover> </mrow> </semantics></math> drawn on the curve <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> is a tensorial expression, valid in any reference frame.</p>
Full article ">Figure 3
<p>The geodesics of a curved space are the only lines capable of performing parallel transport on their own tangent vector.</p>
Full article ">Figure 4
<p>Local mapping of a differentiable manifold. For every point on the manifold, the mapping of the vector <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>V</mi> </mrow> <mo>→</mo> </mover> </mrow> </semantics></math> through the covariant derivative of the tensor <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>T</mi> </mrow> <mrow> <mi>i</mi> </mrow> <mrow> <mi>α</mi> <mi>β</mi> </mrow> </msubsup> </mrow> </semantics></math> corresponds to the vector <math display="inline"><semantics> <mrow> <mover accent="true"> <msup> <mi>V</mi> <mo>′</mo> </msup> <mo>→</mo> </mover> </mrow> </semantics></math> evaluated at the point by <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>T</mi> </mrow> <mrow> <mi>i</mi> </mrow> <mrow> <mi>α</mi> <mi>β</mi> </mrow> </msubsup> </mrow> </semantics></math>.</p>
Full article ">
42 pages, 6682 KiB  
Article
A Tensor Space for Multi-View and Multitask Learning Based on Einstein and Hadamard Products: A Case Study on Vehicle Traffic Surveillance Systems
by Fernando Hermosillo-Reynoso and Deni Torres-Roman
Sensors 2024, 24(23), 7463; https://doi.org/10.3390/s24237463 - 22 Nov 2024
Viewed by 154
Abstract
Since multi-view learning leverages complementary information from multiple feature sets to improve model performance, a tensor-based data fusion layer for neural networks, called Multi-View Data Tensor Fusion (MV-DTF), is used. It fuses M feature spaces X1,,XM, [...] Read more.
Since multi-view learning leverages complementary information from multiple feature sets to improve model performance, a tensor-based data fusion layer for neural networks, called Multi-View Data Tensor Fusion (MV-DTF), is used. It fuses M feature spaces X1,,XM, referred to as views, in a new latent tensor space, S, of order P and dimension J1××JP, defined in the space of affine mappings composed of a multilinear map T:X1××XMS—represented as the Einstein product between a (P+M)-order tensor A anda rank-one tensor, X=x(1)x(M), where x(m)Xm is the m-th view—and a translation. Unfortunately, as the number of views increases, the number of parameters that determine the MV-DTF layer grows exponentially, and consequently, so does its computational complexity. To address this issue, we enforce low-rank constraints on certain subtensors of tensor A using canonical polyadic decomposition, from which M other tensors U(1),,U(M), called here Hadamard factor tensors, are obtained. We found that the Einstein product AMX can be approximated using a sum of R Hadamard products of M Einstein products encoded as U(m)1x(m), where R is related to the decomposition rank of subtensors of A. For this relationship, the lower the rank values, the more computationally efficient the approximation. To the best of our knowledge, this relationship has not previously been reported in the literature. As a case study, we present a multitask model of vehicle traffic surveillance for occlusion detection and vehicle-size classification tasks, with a low-rank MV-DTF layer, achieving up to 92.81% and 95.10% in the normalized weighted Matthews correlation coefficient metric in individual tasks, representing a significant 6% and 7% improvement compared to the single-task single-view models. Full article
(This article belongs to the Section Vehicular Sensing)
10 pages, 241 KiB  
Article
Disaffinity Vectors on a Riemannian Manifold and Their Applications
by Sharief Deshmukh, Amira Ishan and Bang-Yen Chen
Mathematics 2024, 12(23), 3659; https://doi.org/10.3390/math12233659 - 22 Nov 2024
Viewed by 231
Abstract
A disaffinity vector on a Riemannian manifold (M,g) is a vector field whose affinity tensor vanishes. In this paper, we observe that nontrivial disaffinity functions offer obstruction to the topology of M and show that the existence of a [...] Read more.
A disaffinity vector on a Riemannian manifold (M,g) is a vector field whose affinity tensor vanishes. In this paper, we observe that nontrivial disaffinity functions offer obstruction to the topology of M and show that the existence of a nontrivial disaffinity function on M does not allow M to be compact. A characterization of the Euclidean space is also obtained by using nontrivial disaffinity functions. Further, we study properties of disaffinity vectors on M and prove that every Killing vector field is a disaffinity vector. Then, we prove that if the potential field ζ of a Ricci soliton M,g,ζ,λ is a disaffinity vector, then the scalar curvature is constant. As an application, we obtain conditions under which a Ricci soliton M,g,ζ,λ is trivial. Finally, we prove that a Yamabe soliton M,g,ξ,λ with a disaffinity potential field ξ is trivial. Full article
25 pages, 421 KiB  
Article
Propagation Speeds of Relativistic Conformal Particles from a Generalized Relaxation Time Approximation
by Alejandra Kandus and Esteban Calzetta
Entropy 2024, 26(11), 927; https://doi.org/10.3390/e26110927 - 30 Oct 2024
Viewed by 431
Abstract
The propagation speeds of excitations are a crucial input in the modeling of interacting systems of particles. In this paper, we assume the microscopic physics is described by a kinetic theory for massless particles, which is approximated by a generalized relaxation time approximation [...] Read more.
The propagation speeds of excitations are a crucial input in the modeling of interacting systems of particles. In this paper, we assume the microscopic physics is described by a kinetic theory for massless particles, which is approximated by a generalized relaxation time approximation (RTA) where the relaxation time depends on the energy of the particles involved. We seek a solution of the kinetic equation by assuming a parameterized one-particle distribution function (1-pdf) which generalizes the Chapman–Enskog (Ch-En) solution to the RTA. If developed to all orders, this would yield an asymptotic solution to the kinetic equation; we restrict ourselves to an approximate solution by truncating the Ch-En series to the second order. Our generalized Ch-En solution contains undetermined space-time-dependent parameters, and we derive a set of dynamical equations for them by applying the moments method. We check that these dynamical equations lead to energy–momentum conservation and positive entropy production. Finally, we compute the propagation speeds for fluctuations away from equilibrium from the linearized form of the dynamical equations. Considering relaxation times of the form τ=τ0(βμpμ)a, with <a<2, where βμ=uμ/T is the temperature vector in the Landau frame, we show that the Anderson–Witting prescription a=1 yields the fastest speed in all scalar, vector and tensor sectors. This fact ought to be taken into consideration when choosing the best macroscopic description for a given physical system. Full article
(This article belongs to the Section Non-equilibrium Phenomena)
Show Figures

Figure 1

Figure 1
<p>(Color online) Speeds of the two vector modes from Equation (<a href="#FD89-entropy-26-00927" class="html-disp-formula">89</a>). The fastest mode (top dot-dashed blue curve) attains the same maximum speed for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>→</mo> <mo>−</mo> <mo>∞</mo> </mrow> </semantics></math> (top dotted light-blue horizontal line), indicating that the AW value is not exceeded at any value of <span class="html-italic">a</span>. Observe that the slowest mode speed (bottom, dashed light-blue curve) is zero for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, so we recover the AW case [<a href="#B51-entropy-26-00927" class="html-bibr">51</a>] where only one non-null mode exists.</p>
Full article ">Figure 2
<p>(Color online) Speeds of the three scalar modes from Equation (<a href="#FD91-entropy-26-00927" class="html-disp-formula">91</a>). We see that the maximum speed of the fastest mode (top red short-dashed curve) corresponds to the AW solution <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (top horizontal orange dotted line). The intermediate speed mode (long-dashed orange curve in the middle of the figure) also attains its minimum value at the AW value <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (bottom horizontal yellow dotted line), and the speeds of this mode never exceed the ones of the fastest mode. These two modes are the generalization of the AW modes found elsewhere. The bottom, single-line purple curve corresponds to the speeds of a new, slowest mode, whose velocity for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> is zero. Thus, we see that the AW case, for which there are only two propagating modes, is consistently included in our formalism.</p>
Full article ">Figure 3
<p>(Color online) Comparison of the maximum propagation speeds of each sector. The top dotted horizontal line corresponds to the AW (<math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>) scalar mode speed, short-dashed line immediately below corresponds to the velocities of the scalar fastest mode of our model. The middle dotted horizontal line and dot-dashed middle curve correspond to the vector mode speed for AW (<math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>) and to the speeds of our model, respectively. The bottom long-dashed horizontal line is the speed of the tensor mode, which agrees with the AW speed over the entire interval of <span class="html-italic">a</span> values considered. They verify <math display="inline"><semantics> <mrow> <msub> <mi>v</mi> <mi>T</mi> </msub> <mo>&lt;</mo> <msub> <mi>v</mi> <mi>V</mi> </msub> <mo>&lt;</mo> <msub> <mi>v</mi> <mi>S</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">
62 pages, 9349 KiB  
Article
Fokker-Planck Central Moment Lattice Boltzmann Method for Effective Simulations of Fluid Dynamics
by William Schupbach and Kannan Premnath
Fluids 2024, 9(11), 255; https://doi.org/10.3390/fluids9110255 - 29 Oct 2024
Viewed by 626
Abstract
We present a new formulation of the central moment lattice Boltzmann (LB) method based on a minimal continuous Fokker-Planck (FP) kinetic model, originally proposed for stochastic diffusive-drift processes (e.g., Brownian dynamics), by adapting it as a collision model for the continuous Boltzmann equation [...] Read more.
We present a new formulation of the central moment lattice Boltzmann (LB) method based on a minimal continuous Fokker-Planck (FP) kinetic model, originally proposed for stochastic diffusive-drift processes (e.g., Brownian dynamics), by adapting it as a collision model for the continuous Boltzmann equation (CBE) for fluid dynamics. The FP collision model has several desirable properties, including its ability to preserve the quadratic nonlinearity of the CBE, unlike that based on the common Bhatnagar-Gross-Krook model. Rather than using an equivalent Langevin equation as a proxy, we construct our approach by directly matching the changes in different discrete central moments independently supported by the lattice under collision to those given by the CBE under the FP-guided collision model. This can be interpreted as a new path for the collision process in terms of the relaxation of the various central moments to “equilibria”, which we term as the Markovian central moment attractors that depend on the products of the adjacent lower order moments and a diffusion coefficient tensor, thereby involving of a chain of attractors; effectively, the latter are nonlinear functions of not only the hydrodynamic variables, but also the non-conserved moments; the relaxation rates are based on scaling the drift coefficient by the order of the moment involved. The construction of the method in terms of the relevant central moments rather than via the drift and diffusion of the distribution functions directly in the velocity space facilitates its numerical implementation and analysis. We show its consistency to the Navier-Stokes equations via a Chapman-Enskog analysis and elucidate the choice of the diffusion coefficient based on the second order moments in accurately representing flows at relatively low viscosities or high Reynolds numbers. We will demonstrate the accuracy and robustness of our new central moment FP-LB formulation, termed as the FPC-LBM, using the D3Q27 lattice for simulations of a variety of flows, including wall-bounded turbulent flows. We show that the FPC-LBM is more stable than other existing LB schemes based on central moments, while avoiding numerical hyperviscosity effects in flow simulations at relatively very low physical fluid viscosities through a refinement to a model founded on kinetic theory. Full article
(This article belongs to the Special Issue Lattice Boltzmann Methods: Fundamentals and Applications)
Show Figures

Figure 1

Figure 1
<p>Routes for the derivation of the Boltzmann equation and the modeling of its collision term via BGK or FP approach under appropriate approximations, and their applications to representing the dynamics in fluids and plasmas (inspired from [<xref ref-type="bibr" rid="B41-fluids-09-00255">41</xref>]).</p>
Full article ">Figure 2
<p>Representation of collision processes at different levels of modeling description and the associated mathematical nature of the collision operator.</p>
Full article ">Figure 3
<p>Streamlines for two-dimensional lid-driven square cavity flow computed using the FPC-LBM at Reynolds numbers of <inline-formula><mml:math id="mm866"><mml:semantics><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> = 1000, 3200, 5000, and 7500. The formation of secondary and tertiary vortices is consistent with those in the benchmark results of Ghia et al. (1982) [<xref ref-type="bibr" rid="B69-fluids-09-00255">69</xref>] for each Reynolds number shown here.</p>
Full article ">Figure 4
<p>Comparisons of the horizontal velocity component along the vertical centerline in a two-dimensional lid-driven square cavity flow at different Reynolds numbers computed using the FPC-LBM with the reference results of Ghia et al. (1982) [<xref ref-type="bibr" rid="B69-fluids-09-00255">69</xref>]. (<bold>a</bold>) <inline-formula><mml:math id="mm867"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>1000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>b</bold>) <inline-formula><mml:math id="mm868"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>3200</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>c</bold>) <inline-formula><mml:math id="mm869"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>5000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>d</bold>) <inline-formula><mml:math id="mm870"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>7500</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 5
<p>Comparisons of the vertical velocity component along the horizontal centerline in a two-dimensional lid-driven square cavity flow at different Reynolds numbers computed using the FPC-LBM with the reference results of Ghia et al. (1982) [<xref ref-type="bibr" rid="B69-fluids-09-00255">69</xref>]. (<bold>a</bold>) <inline-formula><mml:math id="mm871"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>1000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>b</bold>) <inline-formula><mml:math id="mm872"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>3200</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>c</bold>) <inline-formula><mml:math id="mm873"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>5000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>d</bold>) <inline-formula><mml:math id="mm874"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>7500</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 6
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm875"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of lower Mach numbers of <inline-formula><mml:math id="mm876"><mml:semantics><mml:mrow><mml:mn>0.05</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm877"><mml:semantics><mml:mrow><mml:mn>0.2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm878"><mml:semantics><mml:mrow><mml:mn>0.3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm879"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm880"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm881"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) computed using the Maxwellian equilibria based MCM-LBM. The MCM-LBM is seen to sufficiently capture the physics of this case for all of the grid resolutions and Mach numbers considered here as it has not caused the formation of any spurious secondary vortices.</p>
Full article ">Figure 7
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm882"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of lower Mach numbers of <inline-formula><mml:math id="mm883"><mml:semantics><mml:mrow><mml:mn>0.05</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm884"><mml:semantics><mml:mrow><mml:mn>0.2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm885"><mml:semantics><mml:mrow><mml:mn>0.3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm886"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm887"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm888"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) computed using the Fokker-Planck equilibria based FPC-LBM. The FPC-LBM is seen to sufficiently capture the physics of this case for all of the grid resolutions and Mach numbers considered here as it has not caused the formation of any spurious secondary vortices. The FPC-LBM and MCM-LBM results are almost indistinguishable from one another for these cases indicating the robustness of central moment collision models in general and that we must consider more extreme cases to see significant differences between them.</p>
Full article ">Figure 8
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm889"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of higher Mach numbers of <inline-formula><mml:math id="mm890"><mml:semantics><mml:mrow><mml:mn>0.4</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm891"><mml:semantics><mml:mrow><mml:mn>0.5</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm892"><mml:semantics><mml:mrow><mml:mn>0.6</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm893"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm894"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm895"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) computed using the Maxwellian equilibria based MCM-LBM. It is seen that at larger Mach numbers and coarse grid resolutions, the MCM-LBM becomes unstable with the formation of spurious secondary vortices which form on the layers for the cases of <inline-formula><mml:math id="mm896"><mml:semantics><mml:mrow><mml:mrow><mml:mi>M</mml:mi><mml:mi>a</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>0.4</mml:mn><mml:mo>,</mml:mo><mml:mn>0.5</mml:mn><mml:mo>,</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm897"><mml:semantics><mml:mrow><mml:mn>0.6</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> at <inline-formula><mml:math id="mm898"><mml:semantics><mml:mrow><mml:msup><mml:mi>L</mml:mi><mml:mn>2</mml:mn></mml:msup><mml:mo>=</mml:mo><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 9
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm899"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of higher Mach numbers of <inline-formula><mml:math id="mm900"><mml:semantics><mml:mrow><mml:mn>0.4</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm901"><mml:semantics><mml:mrow><mml:mn>0.5</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm902"><mml:semantics><mml:mrow><mml:mn>0.6</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm903"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm904"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm905"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) computed using the Maxwellian equilibria based MCM-LBM.Vorticity contours at <inline-formula><mml:math id="mm906"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>. It is seen that the FPC-LBM remains stable and no spurious vortices are formed for the same cases of higher Mach numbers and grid resolutions as those shown previously for the MCM-LBM which did not remain stable (see <xref ref-type="fig" rid="fluids-09-00255-f008">Figure 8</xref>). This indicates that the FPC-LBM is a more robust collision model when compared to the MCM-LBM for simulations at coarse grid resolutions and at relatively large Mach numbers.</p>
Full article ">Figure 10
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm907"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of lower Mach numbers of <inline-formula><mml:math id="mm908"><mml:semantics><mml:mrow><mml:mn>0.05</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm909"><mml:semantics><mml:mrow><mml:mn>0.2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm910"><mml:semantics><mml:mrow><mml:mn>0.3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm911"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm912"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm913"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) at an extremely large bulk viscosity by setting the relaxation parameter associated with the bulk viscosity to <inline-formula><mml:math id="mm914"><mml:semantics><mml:mrow><mml:msub><mml:mi>ω</mml:mi><mml:mn>3</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>0.35</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and computed using the Maxwellian equilibria based MCM-LBM. It is seen that MCM-LBM becomes unstable under an extreme increase in bulk viscosity, especially for coarse grid resolutions which become progressively worse as the Mach number is increased. By increasing the relaxation time <inline-formula><mml:math id="mm915"><mml:semantics><mml:mrow><mml:msub><mml:mi>τ</mml:mi><mml:mn>3</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>1</mml:mn><mml:mo>/</mml:mo><mml:msub><mml:mi>ω</mml:mi><mml:mn>3</mml:mn></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula> associated with the bulk viscosity to <inline-formula><mml:math id="mm916"><mml:semantics><mml:mrow><mml:mn>2.857</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> from being <inline-formula><mml:math id="mm917"><mml:semantics><mml:mrow><mml:mo>∼</mml:mo><mml:mn>1.0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> or smaller, the range of beneficial limits is exceeded, and instead the simulations begin to numerically destabilize with MCM-LBM.</p>
Full article ">Figure 11
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm918"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of lower Mach numbers of <inline-formula><mml:math id="mm919"><mml:semantics><mml:mrow><mml:mn>0.05</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm920"><mml:semantics><mml:mrow><mml:mn>0.2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm921"><mml:semantics><mml:mrow><mml:mn>0.3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm92211"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm923"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm924"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) at an extremely large bulk viscosity by setting the relaxation parameter associated with the bulk viscosity to <inline-formula><mml:math id="mm925"><mml:semantics><mml:mrow><mml:msub><mml:mi>ω</mml:mi><mml:mn>3</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>0.35</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and computed using the Fokker-Planck equilibria based FPC-LBM. It is seen that the FPC-LBM remains stable even for the cases with an extreme increase in bulk viscosity shown previously, where the MCM-LBM did not remain stable (see <xref ref-type="fig" rid="fluids-09-00255-f010">Figure 10</xref>). This indicates that the FPC-LBM is numerically more stable when compared to the MCM-LBM in such cases as well.</p>
Full article ">Figure 12
<p>Streamlines for three-dimensional lid-driven cubic cavity flow at Reynolds numbers of <inline-formula><mml:math id="mm926"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>100</mml:mn><mml:mo>,</mml:mo><mml:mn>400</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and 1000 computed using the FPC-LBM along the <inline-formula><mml:math id="mm927"><mml:semantics><mml:mrow><mml:mi>z</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> centerplane shown in (<bold>a</bold>,<bold>d</bold>,<bold>g</bold>), the <inline-formula><mml:math id="mm928"><mml:semantics><mml:mrow><mml:mi>y</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> centerplane shown in (<bold>b</bold>, <bold>e</bold>,<bold>h</bold>), and the <inline-formula><mml:math id="mm929"><mml:semantics><mml:mrow><mml:mi>x</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> centerplane shown in (<bold>c</bold>,<bold>f</bold>,<bold>i</bold>).</p>
Full article ">Figure 13
<p>Comparisons of the horizontal velocity <inline-formula><mml:math id="mm930"><mml:semantics><mml:mrow><mml:mi>u</mml:mi><mml:mo>/</mml:mo><mml:mi>U</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> along the vertical coordinate <inline-formula><mml:math id="mm931"><mml:semantics><mml:mrow><mml:mi>y</mml:mi><mml:mo>/</mml:mo><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> at <inline-formula><mml:math id="mm932"><mml:semantics><mml:mrow><mml:mi>x</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm933"><mml:semantics><mml:mrow><mml:mi>z</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> (left) and vertical velocity <inline-formula><mml:math id="mm934"><mml:semantics><mml:mrow><mml:mi>v</mml:mi><mml:mo>/</mml:mo><mml:mi>U</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> along the horizontal coordinate <inline-formula><mml:math id="mm935"><mml:semantics><mml:mrow><mml:mi>x</mml:mi><mml:mo>/</mml:mo><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> at <inline-formula><mml:math id="mm936"><mml:semantics><mml:mrow><mml:mi>y</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm937"><mml:semantics><mml:mrow><mml:mi>z</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> (right) in a three-dimensional cubic cavity flow at different Reynolds numbers computed using the FPC-LBM with the reference results of Ku et al. (1987) [<xref ref-type="bibr" rid="B71-fluids-09-00255">71</xref>] and Shu et al. (2003) [<xref ref-type="bibr" rid="B72-fluids-09-00255">72</xref>]. (<bold>a</bold>,<bold>b</bold>) <inline-formula><mml:math id="mm938"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>100</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>c</bold>,<bold>d</bold>) <inline-formula><mml:math id="mm939"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>400</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>e</bold>,<bold>f</bold>) <inline-formula><mml:math id="mm940"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>1000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 14
<p>Parallel performance of the MPI implementation of our 3D FPC-LBM for lid-driven cubic cavity flow simulations using a grid resolution of <inline-formula><mml:math id="mm941"><mml:semantics><mml:mrow><mml:mn>150</mml:mn><mml:mo>×</mml:mo><mml:mn>150</mml:mn><mml:mo>×</mml:mo><mml:mn>150</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> in our in-house computer cluster.</p>
Full article ">Figure 15
<p>The maximum possible Reynolds number for which the three-dimensional lid-driven cubic cavity flow simulations remain stable at different grid resolutions of <inline-formula><mml:math id="mm942"><mml:semantics><mml:msup><mml:mn>48</mml:mn><mml:mn>3</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm943"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>3</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm944"><mml:semantics><mml:msup><mml:mn>80</mml:mn><mml:mn>3</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm945"><mml:semantics><mml:msup><mml:mn>96</mml:mn><mml:mn>3</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> for different types of LB collision models—raw moments-based MRT-LBM, Maxwellian central moments-based MCM-LBM, factorized central moments-based Factorized LBM, cumulant LBM, and Fokker-Planck central moments-based FPC-LBM.</p>
Full article ">Figure 16
<p>A comparison of the decay rates produced by different LB collision models—SRT-LBM, Maxwellian central moments-based MCM-LBM, cumulant LBM, and Fokker-Planck central moments-based FPC-LBM as compared to the analytically predicted decay rate for the simulation of orthogonal crossing shear waves. Figure (<bold>a</bold>) indicates that the MCM-LBM fails to produce a decay rate similar to that of the analytical solution, and thus is not able to deal with the numerical hyperviscosity effects associated with this problem. Figure (<bold>b</bold>), which is a highly zoomed version of the left figure, indicates that the SRT-LBM can deal with the hyperviscosity effects but also contains unwanted noise. Furthermore, the cumulant LBM and the FPC-LBM are seen to have nearly identical decay rates that are consistent with the analytical solution.</p>
Full article ">Figure 17
<p>Comparisons of turbulence statistics for fully developed turbulent channel flow at a shear Reynolds number of <inline-formula><mml:math id="mm946"><mml:semantics><mml:mrow><mml:msub><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>*</mml:mo></mml:msub><mml:mo>=</mml:mo><mml:mn>180</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> computed using the FPC-LBM and compared to the direct numerical simulations (DNS) data of Lee and Moser (2015) [<xref ref-type="bibr" rid="B74-fluids-09-00255">74</xref>] and experimental data of Kreplin and Eckelmann (1979) [<xref ref-type="bibr" rid="B76-fluids-09-00255">76</xref>]. (<bold>a</bold>) Mean streamwise velocity, (<bold>b</bold>) Root-mean-square (rms) velocity fluctuations, and (<bold>c</bold>) Reynolds stress along the streamwise-wall normal direction.</p>
Full article ">
20 pages, 25073 KiB  
Article
Development of 6DOF Hardware-in-the-Loop Ground Testbed for Autonomous Robotic Space Debris Removal
by Ahmad Al Ali, Bahador Beigomi and Zheng H. Zhu
Aerospace 2024, 11(11), 877; https://doi.org/10.3390/aerospace11110877 - 25 Oct 2024
Viewed by 607
Abstract
This paper presents the development of a hardware-in-the-loop ground testbed featuring active gravity compensation via software-in-the-loop integration, specially designed to support research in autonomous robotic removal of space debris. The testbed is designed to replicate six degrees of freedom (6DOF) motion maneuvering to [...] Read more.
This paper presents the development of a hardware-in-the-loop ground testbed featuring active gravity compensation via software-in-the-loop integration, specially designed to support research in autonomous robotic removal of space debris. The testbed is designed to replicate six degrees of freedom (6DOF) motion maneuvering to accurately simulate the dynamic behaviors of free-floating robotic manipulators and free-tumbling space debris under microgravity conditions. The testbed incorporates two industrial 6DOF robotic manipulators, a three-finger robotic gripper, and a suite of sensors, including cameras, force/torque sensors, and tactile tensors. Such a setup provides a robust platform for testing and validating technologies related to autonomous tracking, capture, and post-capture stabilization within the context of active space debris removal missions. Preliminary experimental results have demonstrated advancements in motion control, computer vision, and sensor fusion. This facility is positioned to become an essential resource for the development and validation of robotic manipulators in space, offering substantial improvements to the effectiveness and reliability of autonomous capture operations in space missions. Full article
(This article belongs to the Special Issue Space Mechanisms and Robots)
Show Figures

Figure 1

Figure 1
<p>Space robotic manipulator—Canadarm [<a href="#B3-aerospace-11-00877" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>(<b>A</b>) Air-bearing testbed at York University [<a href="#B6-aerospace-11-00877" class="html-bibr">6</a>], (<b>B</b>) air-bearing testbed at the Polish Academy of Sciences [<a href="#B7-aerospace-11-00877" class="html-bibr">7</a>], (<b>C</b>) HIL testbed at Shenzhen Space Technology Center [<a href="#B12-aerospace-11-00877" class="html-bibr">12</a>], (<b>D</b>) European proximity operations simulator at the German Aerospace Center [<a href="#B13-aerospace-11-00877" class="html-bibr">13</a>], (<b>E</b>) CSA SPDM task verification facility [<a href="#B14-aerospace-11-00877" class="html-bibr">14</a>], (<b>F</b>) dual robotic testbed at Tsinghua University [<a href="#B15-aerospace-11-00877" class="html-bibr">15</a>,<a href="#B16-aerospace-11-00877" class="html-bibr">16</a>], (<b>G</b>) MTVF at the China Academy of Space Technology [<a href="#B16-aerospace-11-00877" class="html-bibr">16</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Shenzhen Space Technology Center [<a href="#B12-aerospace-11-00877" class="html-bibr">12</a>], (<b>b</b>) German Aerospace Center [<a href="#B13-aerospace-11-00877" class="html-bibr">13</a>], (<b>c</b>) China Academy of Space Technology [<a href="#B15-aerospace-11-00877" class="html-bibr">15</a>], and (<b>d</b>) Tsinghua University [<a href="#B16-aerospace-11-00877" class="html-bibr">16</a>].</p>
Full article ">Figure 4
<p>Schematic of dual-robot HIL testbed.</p>
Full article ">Figure 5
<p>(<b>A</b>) FANUC manipulator, (<b>B</b>) robotic gripper, (<b>C</b>) ATI force/load sensor.</p>
Full article ">Figure 6
<p>Flowchart illustrating how the simulated motion can be applied to real-world execution.</p>
Full article ">Figure 7
<p>ATI Sensor Frame.</p>
Full article ">Figure 8
<p>The target’s free-floating motion disturbed by an external force.</p>
Full article ">Figure 9
<p>Robotiq three-finger gripper movement.</p>
Full article ">Figure 10
<p>(<b>a</b>) Tactile sensor, (<b>b</b>) Intel camera.</p>
Full article ">Figure 11
<p>Full-scale monitoring of all angles.</p>
Full article ">Figure 12
<p>Mock-up satellite and components.</p>
Full article ">Figure 13
<p>The 6DOF hardware-in-the-loop ground testbed.</p>
Full article ">Figure 14
<p>(<b>A</b>) Yolo feature recognition, (<b>B</b>) AI computer vision for target tracking result.</p>
Full article ">Figure 15
<p>Skeletal representation of the 6DOF simulation environment.</p>
Full article ">Figure 16
<p>Motion equivalence.</p>
Full article ">Figure 17
<p>Joint positions of robot B to deliver the target motion.</p>
Full article ">Figure 18
<p>Mock-up satellite tumbling in space.</p>
Full article ">Figure 19
<p>ATI force/load sensor: force values.</p>
Full article ">Figure 20
<p>ATI force/load sensor: torque values.</p>
Full article ">Figure 21
<p>Robot A/gripper joint positions.</p>
Full article ">Figure 22
<p>Robot A/gripper cartesian positions.</p>
Full article ">Figure 23
<p>Gripper capture of target.</p>
Full article ">Figure 24
<p>Full debris capture mission.</p>
Full article ">Figure 25
<p>Camera’s field of view and bounding box during the pre-capture phase.</p>
Full article ">
15 pages, 312 KiB  
Article
Spinor–Vector Duality and Mirror Symmetry
by Alon E. Faraggi
Universe 2024, 10(10), 402; https://doi.org/10.3390/universe10100402 - 19 Oct 2024
Viewed by 517
Abstract
Mirror symmetry was first observed in worldsheet string constructions, and was shown to have profound implications in the Effective Field Theory (EFT) limit of string compactifications, and for the properties of Calabi–Yau manifolds. It opened up a new field in pure mathematics, and [...] Read more.
Mirror symmetry was first observed in worldsheet string constructions, and was shown to have profound implications in the Effective Field Theory (EFT) limit of string compactifications, and for the properties of Calabi–Yau manifolds. It opened up a new field in pure mathematics, and was utilised in the area of enumerative geometry. Spinor–Vector Duality (SVD) is an extension of mirror symmetry. This can be readily understood in terms of the moduli of toroidal compactification of the Heterotic String, which includes the metric the antisymmetric tensor field and the Wilson line moduli. In terms of the toroidal moduli, mirror symmetry corresponds to mappings of the internal space moduli, whereas Spinor–Vector Duality corresponds to maps of the Wilson line moduli. In the past few of years, we demonstrated the existence of Spinor–Vector Duality in the effective field theory compactifications of string theories. This was achieved by starting with a worldsheet orbifold construction that exhibited Spinor–Vector Duality and resolving the orbifold singularities, hence generating a smooth, effective field theory limit with an imprint of the Spinor–Vector Duality. Just like mirror symmetry, the Spinor–Vector Duality can be used to study the properties of complex manifolds with vector bundles. Spinor–Vector Duality offers a top-down approach to the “Swampland” program, by exploring the imprint of the symmetries of the ultra-violet complete worldsheet string constructions in the effective field theory limit. The SVD suggests a demarcation line between (2,0) EFTs that possess an ultra-violet complete embedding versus those that do not. Full article
Show Figures

Figure 1

Figure 1
<p>Density plot depicting the Spinor–Vector Duality in the space of fermionic <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mn>2</mn> </msub> <mo>×</mo> <msub> <mi>Z</mi> <mn>2</mn> </msub> </mrow> </semantics></math> Heterotic String orbifolds. The figure shows the number of models with a given number of <math display="inline"><semantics> <mrow> <mo>(</mo> <mn>16</mn> <mo>+</mo> <mover> <mn>16</mn> <mo>¯</mo> </mover> <mo>)</mo> </mrow> </semantics></math> and 10 representations of <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>O</mi> <mo>(</mo> <mn>10</mn> <mo>)</mo> </mrow> </semantics></math>. It is symmetric under the exchange of rows and columns, reflecting the SVD that underlies the entire space of vacua.</p>
Full article ">Figure 2
<p>The <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mn>2</mn> </msub> <mo>×</mo> <msubsup> <mi>Z</mi> <mn>2</mn> <mo>′</mo> </msubsup> </mrow> </semantics></math> partition function of the <math display="inline"><semantics> <msup> <mi>g</mi> <mo>′</mo> </msup> </semantics></math>-twist and <span class="html-italic">g</span> Wilson line with discrete torsion <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>±</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">
25 pages, 323 KiB  
Article
Classification of Petrov Homogeneous Spaces
by V. V. Obukhov
Symmetry 2024, 16(10), 1385; https://doi.org/10.3390/sym16101385 - 17 Oct 2024
Viewed by 377
Abstract
In this paper, the final stage of the Petrov classification is carried out. As it is known, the Killing vector fields specify infinitesimal transformations of the group of motions of space V4. In the case where the group of motions [...] Read more.
In this paper, the final stage of the Petrov classification is carried out. As it is known, the Killing vector fields specify infinitesimal transformations of the group of motions of space V4. In the case where the group of motions G3 acts in a simply transitive way in the homogeneous space V4, the geometry of the non-isotropic hypersurface is determined by the geometry of the transitivity space V3 of the group G3. In this case, the metric tensor of the space V3 can be given by a nonholonomic reper consisting of three independent vectors (a)α, which define the generators of the group G3 of finite transformations in the space V3. The representation of the metric tensor of V4 spaces by means of vector fields (a)α has a great physical meaning and makes it possible to substantially simplify the equations of mathematical physics in such spaces. Therefore, the Petrov classification should be complemented by the classification of vector fields (a)α connected to Killing vector fields. For homogeneous spaces, this problem has been largely solved. A complete solution of this problem is presented in the present paper, where I refine the Petrov classification for homogeneous spaces in which the group G3, which belongs to type VIII according to the Petrov classification, acts simply transitively. In addition, this paper provides the complete classification of vector fields (a)α for space V4 in which the group G3 acts simply transitivity on isotropic hypersurfaces. Full article
(This article belongs to the Special Issue Symmetry: Feature Papers 2024)
17 pages, 39089 KiB  
Article
Electronic and Optical Properties of 2D Heterostructure Bilayers of Graphene, Borophene and 2D Boron Carbides from First Principles
by Lu Niu, Oliver J. Conquest, Carla Verdi and Catherine Stampfl
Nanomaterials 2024, 14(20), 1659; https://doi.org/10.3390/nano14201659 - 16 Oct 2024
Viewed by 791
Abstract
In the present work the atomic, electronic and optical properties of two-dimensional graphene, borophene, and boron carbide heterojunction bilayer systems (Graphene–BC3, Graphene–Borophene and Graphene–B4C3) as well as their constituent monolayers are investigated on the basis of first-principles [...] Read more.
In the present work the atomic, electronic and optical properties of two-dimensional graphene, borophene, and boron carbide heterojunction bilayer systems (Graphene–BC3, Graphene–Borophene and Graphene–B4C3) as well as their constituent monolayers are investigated on the basis of first-principles calculations using the HSE06 hybrid functional. Our calculations show that while borophene is metallic, both monolayer BC3 and B4C3 are indirect semiconductors, with band-gaps of 1.822 eV and 2.381 eV as obtained using HSE06. The Graphene–BC3 and Graphene–B4C3 bilayer heterojunction systems maintain the Dirac point-like character of graphene at the K-point with the opening of a very small gap (20–50 meV) and are essentially semi-metals, while Graphene–Borophene is metallic. All bilayer heterostructure systems possess absorbance in the visible region where the resonance frequency and resonance absorption peak intensity vary between structures. Remarkably, all heterojunctions support plasmons within the range 16.5–18.5 eV, while Graphene–B4C3 and Graphene–Borophene exhibit a π-type plasmon within the region 4–6 eV, with the latter possessing an additional plasmon at the lower energy of 1.5–3 eV. The dielectric tensor for Graphene–B4C3 exhibits complex off-diagonal elements due to the lower P3 space group symmetry indicating it has anisotropic dielectric properties and could exhibit optically active (chiral) effects. Our study shows that the two-dimensional heterostructures have desirable optical properties broadening the potential applications of the constituent monolayers. Full article
Show Figures

Figure 1

Figure 1
<p>Optimized atomic structures of BC<sub>3</sub>, borophene, B<sub>4</sub>C<sub>3</sub> and graphene. Boron and carbon atoms are denoted by the green and brown spheres, respectively. Borophene has three unique bonds indicated by <math display="inline"><semantics> <msub> <mi>l</mi> <mn>1</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>l</mi> <mn>2</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>l</mi> <mn>3</mn> </msub> </semantics></math>, while B<sub>4</sub>C<sub>3</sub> has four unique bonds indicated by <math display="inline"><semantics> <msub> <mi>l</mi> <mn>4</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>l</mi> <mn>5</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>l</mi> <mn>6</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>l</mi> <mn>7</mn> </msub> </semantics></math>. The unit cells are highlighted in orange.</p>
Full article ">Figure 2
<p>Band structure of the four monolayer systems, BC<sub>3</sub>, B<sub>4</sub>C<sub>3</sub>, borophene and graphene as calculated using the PBE (blue) and HSE06 (orange) functionals. The Fermi level is indicated by the purple dashed line.</p>
Full article ">Figure 3
<p>Top view of the optimized atomic structures of Graphene–BC<sub>3</sub>, Graphene–Borophene, and Graphene–B<sub>4</sub>C<sub>3</sub>. Boron and carbon atoms are denoted by the green and brown spheres, respectively. The unit cell is indicated by the orange parallelogram.</p>
Full article ">Figure 4
<p>Charge density difference (<math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>ρ</mi> <mo>(</mo> <mi mathvariant="bold">r</mi> <mo>)</mo> </mrow> </semantics></math>, calculated using Equation (<a href="#FD10-nanomaterials-14-01659" class="html-disp-formula">10</a>)) between the monolayers and the heterostructures. Regions of charge accumulation are shown in yellow and regions of charge depletion are shown in blue. The isosurface level is <math display="inline"><semantics> <mrow> <mn>1.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math> a<sub>0</sub><sup>−3</sup> and the top layer is always graphene. The unit cell is indicated by the orange parallelogram.</p>
Full article ">Figure 5
<p>Band structure and total DOS for Graphene–BC<sub>3</sub> as calculated using the PBE (blue) and HSE06 (orange) functionals. The Fermi level is indicated by the purple dashed line.</p>
Full article ">Figure 6
<p>Band structure and total DOS for Graphene–Borophene as calculated using the PBE (blue) and HSE06 (orange) functionals. The Fermi level is indicated by the purple dashed line.</p>
Full article ">Figure 7
<p>Band structure and total DOS for Graphene–B<sub>4</sub>C<sub>3</sub> as calculated using the PBE (blue) and HSE06 (orange) functionals. The Fermi level is indicated by the purple dashed line.</p>
Full article ">Figure 8
<p><b>Left</b>: schematic of the Schottky-Mott model showing the valence and conduction band energies, the Fermi energy, and <span class="html-italic">n</span>-type and <span class="html-italic">p</span>-type Schottky barriers labelled <math display="inline"><semantics> <msub> <mi>E</mi> <mi>V</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>E</mi> <mi>C</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>E</mi> <mi>F</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mo>Φ</mo> <mi>n</mi> </msub> </semantics></math>, and <math display="inline"><semantics> <msub> <mo>Φ</mo> <mi>p</mi> </msub> </semantics></math>, respectively. <b>Right</b>: Example for the HSE06 calculated Graphene–B<sub>4</sub>C<sub>3</sub> heterojunction showing the determined Schottky barrier height from the band structure.</p>
Full article ">Figure 9
<p>The real and imaginary parts of the in-plane and out-of-plane dielectric function as a function of photon energy as calculated using the PBE and HSE06 functionals for the Graphene–Borophene heterostructure.</p>
Full article ">Figure 10
<p>The real and imaginary parts of the in-plane and out-of-plane dielectric function as a function of photon energy as calculated using the PBE and HSE06 functionals for the Graphene–BC<sub>3</sub> heterostructure.</p>
Full article ">Figure 11
<p>The real and imaginary parts of the in-plane and out-of-plane dielectric function as a function of photon energy as calculated using the PBE and HSE06 functionals for the Graphene–B<sub>4</sub>C<sub>3</sub> heterostructure.</p>
Full article ">Figure 12
<p>The in-plane and out-of-plane adsorption coefficient (upper) and energy loss spectrum (lower) as a function of photon energy as calculated using the HSE06 functional.</p>
Full article ">Figure 13
<p>The in-plane and out-of-plane reflectivity (upper) and refractive index (lower) as a function of photon energy as calculated using the HSE06 functional.</p>
Full article ">
8 pages, 2764 KiB  
Proceeding Paper
Deep Learning System for E-Waste Management
by Godfrey Perfectson Oise and Susan Konyeha
Eng. Proc. 2024, 67(1), 66; https://doi.org/10.3390/engproc2024067066 - 16 Oct 2024
Viewed by 632
Abstract
The deep learning system for e-waste management presented in this proposal is a transformative solution designed to address the escalating challenges of garbage collection and management in urban environments. Rapid urbanization has resulted in increased waste generation, necessitating a more intelligent and efficient [...] Read more.
The deep learning system for e-waste management presented in this proposal is a transformative solution designed to address the escalating challenges of garbage collection and management in urban environments. Rapid urbanization has resulted in increased waste generation, necessitating a more intelligent and efficient approach to e-waste collection and disposal. This system integrates cutting-edge technologies, primarily Artificial Intelligence (AI), to improve e-waste management processes, enhance resource utilization, and contribute to the creation of cleaner and more sustainable urban spaces. Urban areas are experiencing unprecedented growth, leading to a surge in the volume of waste generated daily; as such, traditional waste management systems struggle to cope with this influx, resulting in environmental pollution, compromised public health, and inefficient resource utilization. The proposed deep learning model with accuracy of 83% seeks to revolutionize existing practices by leveraging the capabilities of AI. The aim of this research is to develop a sequential neural network using a Keras and TensorFlow image analysis: a deep learning convolutional neural network (CNN) for e-waste management. The Python programming tool will be used to develop the deep learning model as well as a GUI that will facilitate human–computer interactions. The system will be tested and the result evaluated to assess the functionality and adequacy of the system. Full article
(This article belongs to the Proceedings of The 3rd International Electronic Conference on Processes)
Show Figures

Figure 1

Figure 1
<p>Dataset images of e-waste categories.</p>
Full article ">Figure 2
<p>Sequential neural network architecture.</p>
Full article ">Figure 3
<p>Model training process.</p>
Full article ">Figure 4
<p>Performance graphs. Accuracy performance graph (<b>a</b>); Loss performance graph (<b>b</b>).</p>
Full article ">Figure 5
<p>Confusion matrix.</p>
Full article ">Figure 6
<p>Prediction and accuracy of the battery class.</p>
Full article ">Figure 7
<p>Prediction and accuracy of the keyboard class.</p>
Full article ">
16 pages, 278 KiB  
Article
Characterizations of Spheres and Euclidean Spaces by Conformal Vector Fields
by Sharief Deshmukh, Nasser Bin Turki and Ramesh Sharma
Mathematics 2024, 12(20), 3163; https://doi.org/10.3390/math12203163 - 10 Oct 2024
Viewed by 473
Abstract
A nontrivial conformal vector field ω on an m-dimensional connected Riemannian manifold Mm,g has naturally associated with it the conformal potential θ, a smooth function on Mm, and a skew-symmetric tensor T of type [...] Read more.
A nontrivial conformal vector field ω on an m-dimensional connected Riemannian manifold Mm,g has naturally associated with it the conformal potential θ, a smooth function on Mm, and a skew-symmetric tensor T of type (1,1) called the associated tensor. There is a third entity, namely the vector field Tω, called the orthogonal reflection field, and in this article, we study the impact of the condition that commutator ω,Tω=0; this condition that we refer to as the orthogonal reflection field is commutative. As a natural impact of this condition, we see the existence of a smooth function ρ on Mm such that θ=ρω; this function ρ is called the proportionality function. First, we show that an m-dimensional compact and connected Riemannian manifold Mm,g admits a nontrivial conformal vector field ω with a commuting orthogonal reflection Tω and constant proportionality function ρ if and only if Mm,g is isometric to the sphere Sm(c) of constant curvature c. Secondly, we find three more characterizations of the sphere and two characterizations of a Euclidean space using these ideas. Finally, we provide a condition for a conformal vector field on a compact Riemannian manifold to be closed. Full article
(This article belongs to the Special Issue Advances in Differential Geometry and Its Applications)
18 pages, 4198 KiB  
Article
Two-Stage Optimization Model Based on Neo4j-Dueling Deep Q Network
by Tie Chen, Pingping Yang, Hongxin Li, Jiaqi Gao and Yimin Yuan
Energies 2024, 17(19), 4998; https://doi.org/10.3390/en17194998 - 8 Oct 2024
Viewed by 498
Abstract
To alleviate the power flow congestion in active distribution networks (ADNs), this paper proposes a two-stage load transfer optimization model based on Neo4j-Dueling DQN. First, the Neo4j graph model was established as the training environment for Dueling DQN. Meanwhile, the power supply paths [...] Read more.
To alleviate the power flow congestion in active distribution networks (ADNs), this paper proposes a two-stage load transfer optimization model based on Neo4j-Dueling DQN. First, the Neo4j graph model was established as the training environment for Dueling DQN. Meanwhile, the power supply paths from the congestion point to the power source point were obtained using the Cypher language built into Neo4j, forming a load transfer space that served as the action space. Secondly, based on various constraints in the load transfer process, a reward and penalty function was formulated to establish the Dueling DQN training model. Finally, according to the εgreedy action selection strategy, actions were selected from the action space and interacted with the Neo4j environment, resulting in the optimal load transfer operation sequence. In this paper, Python was used as the programming language, TensorFlow open-source software library was used to form a deep reinforcement network, and Py2neo toolkit was used to complete the linkage between the python platform and Neo4j. We conducted experiments on a real 79-node system, using three power flow congestion scenarios for validation. Under the three power flow congestion scenarios, the time required to obtain the results was 2.87 s, 4.37 s and 3.45 s, respectively. For scenario 1 before and after load transfer, the line loss, voltage deviation and line load rate were reduced by about 56.0%, 76.0% and 55.7%, respectively. For scenario 2 before and after load transfer, the line loss, voltage deviation and line load rate were reduced by 41.7%, 72.9% and 56.7%, respectively. For scenario 3 before and after load transfer, the line loss, voltage deviation and line load rate were reduced by 13.6%, 47.1% and 37.7%, respectively. The experimental results show that the trained model can quickly and accurately derive the optimal load transfer operation sequence under different power flow congestion conditions, thereby validating the effectiveness of the proposed model. Full article
(This article belongs to the Section F1: Electrical Power System)
Show Figures

Figure 1

Figure 1
<p>Model structure and framework of Neo4j−Dueling DQN.</p>
Full article ">Figure 2
<p>Mapping of Neo4j node attributes.</p>
Full article ">Figure 3
<p>Results of potential power supply path search.</p>
Full article ">Figure 4
<p>Structure of Dueling DQN neural network.</p>
Full article ">Figure 5
<p><b>The</b> interaction model between Dueling DQN and Neo4j.</p>
Full article ">Figure 6
<p>The algorithm process of Neo4j-Dueling DQN.</p>
Full article ">Figure 7
<p>ADN topology diagram.</p>
Full article ">Figure 8
<p>Cumulative reward value results of model training.</p>
Full article ">Figure 9
<p>The output curve of DG.</p>
Full article ">Figure 10
<p>Three types of load demand curves.</p>
Full article ">Figure 11
<p>L1-L9 power flow congestion condition.</p>
Full article ">Figure 12
<p>Changes in evaluation indexes in the process of load transfer under three scenarios. (<b>a</b>) Changes in evaluation indicators in scenario 1, (<b>b</b>) changes in evaluation indicators in scenario 2, and (<b>c</b>) changes in evaluation indicators in scenario 3.</p>
Full article ">Figure 13
<p>Changes in closing current during load transfer in three scenarios. (<b>a</b>) Change in closing current in scenario 1, (<b>b</b>) change in closing current in scenario 2; (<b>c</b>) change in closing current in scenario 3.</p>
Full article ">Figure 14
<p>Constraint changes in penalty item during load transfer operation. (<b>a</b>) The change in penalty constraints in scenario 1, (<b>b</b>) the change in penalty constraints in scenario 2, and (<b>c</b>) the change in penalty constraints in scenario 3.</p>
Full article ">Figure 15
<p>Comparison of the average reward value of the three algorithms.</p>
Full article ">
16 pages, 2454 KiB  
Article
Numerical Modeling of Plasticity in Metal Matrix Fiber Composites
by Gennadiy Lvov and Maria Tănase
Appl. Sci. 2024, 14(19), 8679; https://doi.org/10.3390/app14198679 - 26 Sep 2024
Viewed by 552
Abstract
This paper presents micromechanical analyses of an orthogonally reinforced composite with new constitutive equations of kinematic plastic hardening. The homogenization of plastic properties was performed through a numerical analysis of a representative volume using the finite element method. A modification of Prager’s theory [...] Read more.
This paper presents micromechanical analyses of an orthogonally reinforced composite with new constitutive equations of kinematic plastic hardening. The homogenization of plastic properties was performed through a numerical analysis of a representative volume using the finite element method. A modification of Prager’s theory was used to construct physical relations for an equivalent orthotropic material. In the proposed version of the theory, a special tensor for back stresses is introduced, which takes into account the difference in the rate of hardening for different types of plastic deformation. For boron–aluminum orthogonally reinforced composite with known mechanical properties of fibers and matrix, all material parameters of the theory were determined, deformation diagrams were constructed, and the equation for a plasticity surface in a six-dimensional stress space was obtained. The advantage of the developed method of numerical homogenization is that it only requires a minimal amount of experimental data. The efficiency of micromechanical analysis makes it possible to optimally design metal matrix composites with the required plastic properties. Full article
Show Figures

Figure 1

Figure 1
<p>Representative volume element. (<b>a</b>) Geometrical model; (<b>b</b>) finite element model.</p>
Full article ">Figure 2
<p>Distributions of equivalent stresses: (<b>a</b>) uniaxial stretching along the 0X axis; (<b>b</b>) uniaxial stretching along the 0Z axis; (<b>c</b>) shear in the XY pane; (<b>d</b>) shear in the XZ plane.</p>
Full article ">Figure 3
<p>Distributions of equivalent plastic strains: (<b>a</b>) uniaxial stretching along the 0X axis; (<b>b</b>) uniaxial stretching along the 0Z axis; (<b>c</b>) shear in the XY pane; (<b>d</b>) shear in the XZ plane.</p>
Full article ">Figure 4
<p>Diagrams for (<b>a</b>) uniaxial stretching along the 0X and 0Z axes; (<b>b</b>) shear in the XY and XZ planes.</p>
Full article ">Figure 5
<p>Sections of plasticity surface in coordinates: (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>x</mi> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>z</mi> </mrow> </msub> </mrow> </semantics></math> under tension along x-axis; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>x</mi> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>z</mi> </mrow> </msub> </mrow> </semantics></math> under tension along z-axis; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>x</mi> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics></math> under tension along x-axis; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>x</mi> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics></math> under shear in plane X0Y.</p>
Full article ">
18 pages, 764 KiB  
Tutorial
A Tutorial on the Use of Physics-Informed Neural Networks to Compute the Spectrum of Quantum Systems
by Lorenzo Brevi, Antonio Mandarino and Enrico Prati
Technologies 2024, 12(10), 174; https://doi.org/10.3390/technologies12100174 - 26 Sep 2024
Cited by 1 | Viewed by 1631
Abstract
Quantum many-body systems are of great interest for many research areas, including physics, biology, and chemistry. However, their simulation is extremely challenging, due to the exponential growth of the Hilbert space with system size, making it exceedingly difficult to parameterize the wave functions [...] Read more.
Quantum many-body systems are of great interest for many research areas, including physics, biology, and chemistry. However, their simulation is extremely challenging, due to the exponential growth of the Hilbert space with system size, making it exceedingly difficult to parameterize the wave functions of large systems by using exact methods. Neural networks and machine learning, in general, are a way to face this challenge. For instance, methods like tensor networks and neural quantum states are being investigated as promising tools to obtain the wave function of a quantum mechanical system. In this tutorial, we focus on a particularly promising class of deep learning algorithms. We explain how to construct a Physics-Informed Neural Network (PINN) able to solve the Schrödinger equation for a given potential, by finding its eigenvalues and eigenfunctions. This technique is unsupervised, and utilizes a novel computational method in a manner that is barely explored. PINNs are a deep learning method that exploit automatic differentiation to solve integro-differential equations in a mesh-free way. We show how to find both the ground and the excited states. The method discovers the states progressively by starting from the ground state. We explain how to introduce inductive biases in the loss to exploit further knowledge of the physical system. Such additional constraints allow for a faster and more accurate convergence. This technique can then be enhanced by a smart choice of collocation points in order to take advantage of the mesh-free nature of the PINN. The methods are made explicit by applying them to the infinite potential well and the particle in a ring, a challenging problem to be learned by an artificial intelligence agent due to the presence of complex-valued eigenfunctions and degenerate states Full article
(This article belongs to the Section Quantum Technologies)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of the approach followed in this paper. The circles represent variables; the rectangles represent operations. The variables are also represented by primary colors, while the methods are represented by the secondary colors obtained by combining the colors of their input and their output. In green, there are the neural network inputs. In yellow, there are the two neural networks: the main one that computes the eigenstates and the one that computes the energy. In blue, there are the outputs of the neural networks: the auxiliary outputs, the main outputs, which are the eigenfunctions, and the eigenvalue <span class="html-italic">E</span>. In purple, there are the operations needed to calculate the losses. The losses are the integral loss, the boundary condition loss, the normalization loss, the inductive biases, and the differential equation loss. SAE is the sum of the absolute errors; SSE is the sum of the squared errors. In red, there is the final loss. It is calculated by summing the partial losses weighted by the empirically adjusted hyperparameters <math display="inline"><semantics> <msub> <mi>w</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>w</mi> <mrow> <mi>b</mi> <mi>c</mi> </mrow> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>w</mi> <mrow> <mi>n</mi> <mi>o</mi> <mi>r</mi> <mi>m</mi> </mrow> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>w</mi> <mrow> <mi>b</mi> <mi>i</mi> <mi>a</mi> <mi>s</mi> </mrow> </msub> </semantics></math>, and <math display="inline"><semantics> <msub> <mi>w</mi> <mrow> <mi>e</mi> <mi>q</mi> </mrow> </msub> </semantics></math>. In cyan, there are the criteria and the optimization method.</p>
Full article ">Figure 2
<p>(<b>a</b>) Ground state for the well. The red dots are the PINN’s predictions, while the blue line is the ground truth. The yellow vertical lines represent the walls of the well. (<b>b</b>) Fidelity throughout training for the ground state. (<b>c</b>) Loss behavior throughout training. The y-axis is in logarithmic scale; therefore, the oscillations on the y-axis for low losses are overemphasized.</p>
Full article ">Figure 3
<p>(<b>a</b>) Fifth excited state for the well Equation (<a href="#FD22-technologies-12-00174" class="html-disp-formula">22</a>). The red dots are the PINN’s predictions, while the blue line is the ground truth. The yellow vertical lines represent the walls of the well. (<b>b</b>) Fidelity throughout training for the ground state. (<b>c</b>) Loss behavior throughout training. The y-axis is in logarithmic scale; therefore, the oscillations on the y-axis for low losses are overemphasized.</p>
Full article ">Figure 4
<p>(<b>a</b>) One of the two first excited states for the ring Equation (<a href="#FD26-technologies-12-00174" class="html-disp-formula">26</a>). The red dots are the PINN’s predictions, while the blue line is the ground truth. Note that in this case, the ground truth and the neural prediction are actually different states of the pair of degenerate states for <math display="inline"><semantics> <mrow> <mo>|</mo> <mi>n</mi> <mo>|</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>. This discrepancy is due to the fact that the two minima given by the states for <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> are equally valid, and thus the neural network can converge freely to either one. However, in order to have fidelity be an informative metric, the real and imaginary parts of the neural network predictions are first multiplied by <math display="inline"><semantics> <mrow> <mi>s</mi> <mi>i</mi> <mi>g</mi> <mi>n</mi> <mfenced separators="" open="(" close=")"> <mrow> <mo>〈</mo> <mi>R</mi> <mi>e</mi> <mfenced separators="" open="(" close=")"> <msub> <mi>ψ</mi> <mrow> <mi>E</mi> <mi>x</mi> </mrow> </msub> </mfenced> </mrow> <mo stretchy="false">|</mo> <mrow> <mi>R</mi> <mi>e</mi> <mfenced separators="" open="(" close=")"> <msub> <mi>ψ</mi> <mrow> <mi>P</mi> <mi>I</mi> <mi>N</mi> <mi>N</mi> </mrow> </msub> </mfenced> <mo>〉</mo> </mrow> </mfenced> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>s</mi> <mi>i</mi> <mi>g</mi> <mi>n</mi> <mfenced separators="" open="(" close=")"> <mrow> <mo>〈</mo> <mi>I</mi> <mi>m</mi> <mfenced separators="" open="(" close=")"> <msub> <mi>ψ</mi> <mrow> <mi>E</mi> <mi>x</mi> </mrow> </msub> </mfenced> </mrow> <mo stretchy="false">|</mo> <mrow> <mi>I</mi> <mi>m</mi> <mfenced separators="" open="(" close=")"> <msub> <mi>ψ</mi> <mrow> <mi>P</mi> <mi>I</mi> <mi>N</mi> <mi>N</mi> </mrow> </msub> </mfenced> </mrow> <mo>〉</mo> </mfenced> </mrow> </semantics></math>, respectively, resulting in the two plots superimposing in the figure. (<b>b</b>) Fidelity throughout training for the ground state. (<b>c</b>) Loss behavior throughout training. The y-axis is in logarithmic scale; therefore, the oscillations on the y-axis for low losses are overemphasized.</p>
Full article ">
20 pages, 1029 KiB  
Article
Tensor Network Space-Time Spectral Collocation Method for Time-Dependent Convection-Diffusion-Reaction Equations
by Dibyendu Adak, Duc P. Truong, Gianmarco Manzini, Kim Ø. Rasmussen and Boian S. Alexandrov
Mathematics 2024, 12(19), 2988; https://doi.org/10.3390/math12192988 - 25 Sep 2024
Viewed by 670
Abstract
Emerging tensor network techniques for solutions of partial differential equations (PDEs), known for their ability to break the curse of dimensionality, deliver new mathematical methods for ultra-fast numerical solutions of high-dimensional problems. Here, we introduce a Tensor Train (TT) Chebyshev spectral collocation method, [...] Read more.
Emerging tensor network techniques for solutions of partial differential equations (PDEs), known for their ability to break the curse of dimensionality, deliver new mathematical methods for ultra-fast numerical solutions of high-dimensional problems. Here, we introduce a Tensor Train (TT) Chebyshev spectral collocation method, in both space and time, for the solution of the time-dependent convection-diffusion-reaction (CDR) equation with inhomogeneous boundary conditions, in Cartesian geometry. Previous methods for numerical solution of time-dependent PDEs often used finite difference for time, and a spectral scheme for the spatial dimensions, which led to a slow linear convergence. Spectral collocation space-time methods show exponential convergence; however, for realistic problems they need to solve large four-dimensional systems. We overcome this difficulty by using a TT approach, as its complexity only grows linearly with the number of dimensions. We show that our TT space-time Chebyshev spectral collocation method converges exponentially, when the solution of the CDR is smooth, and demonstrate that it leads to a very high compression of linear operators from terabytes to kilobytes in TT-format, and a speedup of tens of thousands of times when compared to a full-grid space-time spectral method. These advantages allow us to obtain the solutions at much higher resolutions. Full article
(This article belongs to the Section Computational and Applied Mathematics)
Show Figures

Figure 1

Figure 1
<p>One-dimensional space-time grid with <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> collocation nodes.</p>
Full article ">Figure 2
<p>TT decomposition of a 4D tensor <math display="inline"><semantics> <mi mathvariant="script">X</mi> </semantics></math>, with TT rank <math display="inline"><semantics> <mrow> <mi mathvariant="bold">r</mi> <mo>=</mo> <mfenced separators="" open="[" close="]"> <msub> <mi>r</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>r</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>r</mi> <mn>3</mn> </msub> </mfenced> </mrow> </semantics></math> and approximation error <math display="inline"><semantics> <mi>ε</mi> </semantics></math>, in accordance with Equation (<a href="#FD21-mathematics-12-02988" class="html-disp-formula">21</a>).</p>
Full article ">Figure 3
<p>Representation of a linear matrix <math display="inline"><semantics> <mi mathvariant="bold">A</mi> </semantics></math> in the TT-matrix format. First, we reshape the operation matrix <b>A</b> and permute its indices to create the tensor <math display="inline"><semantics> <mi mathvariant="script">A</mi> </semantics></math>. Then, we factorize the tensor in the tensor-train matrix format according to Equation (<a href="#FD23-mathematics-12-02988" class="html-disp-formula">23</a>) to obtain <math display="inline"><semantics> <msup> <mi mathvariant="script">A</mi> <mrow> <mi>T</mi> <mi>T</mi> </mrow> </msup> </semantics></math>.</p>
Full article ">Figure 4
<p>CUR decomposition.</p>
Full article ">Figure 5
<p>Test 1: <b>Left Panel</b>: Relative error curve in <math display="inline"><semantics> <msup> <mi>L</mi> <mn>2</mn> </msup> </semantics></math> norm showing the exponential convergence of SP-SP schemes. <b>Middle Panel</b>: Elapsed time in seconds. <b>Right Panel</b>: Compression ratio of the solution. All plots are versus the number of points per dimension.</p>
Full article ">Figure 6
<p>Test 1: Comparison between TT SP-SP and TT FD-FD. <b>Left Panel</b>: Plot of relative error versus number of points per dimension. <b>Middle Panel</b>: Plot of elapsed time versus number of points per dimension. <b>Right Panel</b>: Plot of relative error versus elapsed times, which shows that the TT SP-SP is more efficient compared to TT FD-FD.</p>
Full article ">Figure 7
<p>Test 2. <b>Left Panel</b>: Relative error curve in <math display="inline"><semantics> <msup> <mi>L</mi> <mn>2</mn> </msup> </semantics></math> norm showing the exponential convergence of the SP-SP schemes. <b>Middle Panel</b>: Elapsed time in seconds. TT format allows solutions in much higher resolution compared to the full-grid format. <b>Right Panel</b>: Compression ratio of the solution. All plots are versus the number of points per dimension.</p>
Full article ">Figure 8
<p>Test 2: Comparison between TT SP-SP and TT FD-FD. <b>Left Panel</b>: Relative error versus number of points per dimension. <b>Middle Panel</b>: Elapsed time versus number of points per dimension. <b>Right Panel</b>: Relative error versus elapsed times, which shows that the TT SP-SP is more efficient compared to TT FD-FD.</p>
Full article ">Figure 9
<p>Test 3. <b>Left Panel</b>: Relative error curve in <math display="inline"><semantics> <msup> <mi>L</mi> <mn>2</mn> </msup> </semantics></math> norm showing the linear convergence of SP-SP schemes, while the theoretical expected quadratic convergence is shown by the triangle. <b>Middle Panel</b>: Elapsed time in seconds. TT format allows solutions in much higher resolution compared to the full-grid format. <b>Right Panel</b>: Compression ratio of the solution. All plots are versus the number of points per dimension.</p>
Full article ">Figure 10
<p>Test case 3: Comparison between TT SP-SP and TT FD-FD. <b>Left Panel</b>: Relative error versus number of points per dimension. <b>Middle Panel</b>: Elapsed time versus number of points per dimension. As the grid becomes larger, the TT SP-SP requires more computational time. <b>Right Panel</b>: Relative error versus elapsed times, which shows that the TT SP-SP is more efficient compared to TT FD-FD.</p>
Full article ">
Back to TopTop