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

Article Types

Countries / Regions

Search Results (6)

Search Parameters:
Keywords = mutual conditional entropy matrix

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
45 pages, 7545 KiB  
Review
Hamiltonian Computational Chemistry: Geometrical Structures in Chemical Dynamics and Kinetics
by Stavros C. Farantos
Entropy 2024, 26(5), 399; https://doi.org/10.3390/e26050399 - 30 Apr 2024
Viewed by 1783
Abstract
The common geometrical (symplectic) structures of classical mechanics, quantum mechanics, and classical thermodynamics are unveiled with three pictures. These cardinal theories, mainly at the non-relativistic approximation, are the cornerstones for studying chemical dynamics and chemical kinetics. Working in extended phase spaces, we show [...] Read more.
The common geometrical (symplectic) structures of classical mechanics, quantum mechanics, and classical thermodynamics are unveiled with three pictures. These cardinal theories, mainly at the non-relativistic approximation, are the cornerstones for studying chemical dynamics and chemical kinetics. Working in extended phase spaces, we show that the physical states of integrable dynamical systems are depicted by Lagrangian submanifolds embedded in phase space. Observable quantities are calculated by properly transforming the extended phase space onto a reduced space, and trajectories are integrated by solving Hamilton’s equations of motion. After defining a Riemannian metric, we can also estimate the length between two states. Local constants of motion are investigated by integrating Jacobi fields and solving the variational linear equations. Diagonalizing the symplectic fundamental matrix, eigenvalues equal to one reveal the number of constants of motion. For conservative systems, geometrical quantum mechanics has proved that solving the Schrödinger equation in extended Hilbert space, which incorporates the quantum phase, is equivalent to solving Hamilton’s equations in the projective Hilbert space. In classical thermodynamics, we take entropy and energy as canonical variables to construct the extended phase space and to represent the Lagrangian submanifold. Hamilton’s and variational equations are written and solved in the same fashion as in classical mechanics. Solvers based on high-order finite differences for numerically solving Hamilton’s, variational, and Schrödinger equations are described. Employing the Hénon–Heiles two-dimensional nonlinear model, representative results for time-dependent, quantum, and dissipative macroscopic systems are shown to illustrate concepts and methods. High-order finite-difference algorithms, despite their accuracy in low-dimensional systems, require substantial computer resources when they are applied to systems with many degrees of freedom, such as polyatomic molecules. We discuss recent research progress in employing Hamiltonian neural networks for solving Hamilton’s equations. It turns out that Hamiltonian geometry, shared with all physical theories, yields the necessary and sufficient conditions for the mutual assistance of humans and machines in deep-learning processes. Full article
(This article belongs to the Special Issue Kinetic Models of Chemical Reactions)
Show Figures

Figure 1

Figure 1
<p>Manifolds and functions which determine the geometrical structures of a classical system with <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> coordinates. <math display="inline"><semantics> <msup> <mi>q</mi> <mn>0</mn> </msup> </semantics></math> denotes a parameter and, specifically, the time for time-dependent systems. <math display="inline"><semantics> <mrow> <mi>q</mi> <mo>=</mo> <mo>(</mo> <msup> <mi>q</mi> <mn>1</mn> </msup> <mo>,</mo> <mo>…</mo> <mo>,</mo> <msup> <mi>q</mi> <mi>n</mi> </msup> <mo>)</mo> </mrow> </semantics></math> are the <span class="html-italic">n</span> coordinates that define the configurations of the system and <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>=</mo> <mo>(</mo> <msub> <mi>p</mi> <mn>1</mn> </msub> <mo>,</mo> <mo>…</mo> <mo>,</mo> <msub> <mi>p</mi> <mi>n</mi> </msub> <mo>)</mo> </mrow> </semantics></math> their canonical conjugate momenta. Details are given in the text.</p>
Full article ">Figure 2
<p>The description of the <span class="html-italic"><b>variation vector field</b></span>, <math display="inline"><semantics> <msub> <mi>Y</mi> <mi>x</mi> </msub> </semantics></math>, with respect to a reference trajectory with vector field <math display="inline"><semantics> <msub> <mi>X</mi> <mi>x</mi> </msub> </semantics></math> and initial condition <math display="inline"><semantics> <msub> <mi>x</mi> <mn>0</mn> </msub> </semantics></math>. <math display="inline"><semantics> <mrow> <msub> <mo>Φ</mo> <msub> <mi>x</mi> <mn>0</mn> </msub> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> denotes the Hamiltonian flow.</p>
Full article ">Figure 3
<p>Manifolds and functions which determine the geometrical structures of a quantum system. The states of the quantum system are the elements of a <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>–</mo> </mrow> </semantics></math>dimensional complex vector space (Hilbert space <math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="script">H</mi> </mrow> <mi>n</mi> </msup> <mo>≅</mo> <msup> <mi mathvariant="double-struck">C</mi> <mi mathvariant="normal">n</mi> </msup> <mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> that includes the vectors <math display="inline"><semantics> <mrow> <mo>|</mo> <mi>ψ</mi> <mo>&gt;</mo> </mrow> </semantics></math>, their complex conjugate <math display="inline"><semantics> <mrow> <mo>&lt;</mo> <mi>ψ</mi> <mo>|</mo> </mrow> </semantics></math>, and linear transformations <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> </mrow> <mover accent="true"> <mi>ψ</mi> <mo>˙</mo> </mover> <mo>&gt;</mo> </mrow> </semantics></math>. Hermitian inner products, <math display="inline"><semantics> <mrow> <mo>&lt;</mo> <mi>ϕ</mi> <mo>|</mo> <mi>ψ</mi> <mo>&gt;</mo> </mrow> </semantics></math>, are employed for Hilbert spaces. The tangent bundle <math display="inline"><semantics> <mrow> <mo>(</mo> <msup> <mrow> <mi mathvariant="script">H</mi> </mrow> <mi>n</mi> </msup> <mo>×</mo> <msup> <mrow> <mi mathvariant="script">H</mi> </mrow> <mi>n</mi> </msup> <mo>)</mo> </mrow> </semantics></math> is mapped to the Extended Hilbert Space <math display="inline"><semantics> <mrow> <mo>(</mo> <msup> <mrow> <mi mathvariant="script">H</mi> </mrow> <mrow> <mi>n</mi> <mo>+</mo> <mn>1</mn> </mrow> </msup> <mo>×</mo> <msup> <mrow> <mi mathvariant="script">H</mi> </mrow> <mrow> <mi>n</mi> <mo>+</mo> <mn>1</mn> </mrow> </msup> <mo>)</mo> </mrow> </semantics></math> by the inclusion of a complex phase <math display="inline"><semantics> <mi>λ</mi> </semantics></math>, the elements of the unitary group <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math>, that produces the rays <math display="inline"><semantics> <mrow> <mo>{</mo> <mo>|</mo> <mi>ψ</mi> <mo>&gt;</mo> <mo>}</mo> <mo>:</mo> <mo>=</mo> <mo>{</mo> <mi>λ</mi> <mo>|</mo> <mi>ψ</mi> <mo>&gt;</mo> <mo>}</mo> </mrow> </semantics></math>. The canonical projection, <math display="inline"><semantics> <msub> <mi>π</mi> <mrow> <mrow> <mo>|</mo> </mrow> <msub> <mi>ψ</mi> <mi mathvariant="double-struck">P</mi> </msub> <mo>&gt;</mo> </mrow> </msub> </semantics></math>, projects the rays in <math display="inline"><semantics> <msup> <mrow> <mi mathvariant="script">H</mi> </mrow> <mrow> <mi>n</mi> <mo>+</mo> <mn>1</mn> </mrow> </msup> </semantics></math> onto the Projective Hilbert Space <math display="inline"><semantics> <mrow> <msup> <mi mathvariant="double-struck">P</mi> <mi>n</mi> </msup> <mrow> <mo>(</mo> <msup> <mrow> <mi mathvariant="script">H</mi> </mrow> <mrow> <mi>n</mi> <mo>+</mo> <mn>1</mn> </mrow> </msup> <mo>)</mo> </mrow> </mrow> </semantics></math>, the space where the physical states of the system live. Details are given in the text.</p>
Full article ">Figure 4
<p>Manifolds and functions which determine the geometrical structures of a thermodynamical system with <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> coordinates. <span class="html-italic">S</span> denotes the entropy and <math display="inline"><semantics> <mrow> <mi>q</mi> <mo>=</mo> <msup> <mrow> <mo>(</mo> <msup> <mi>q</mi> <mn>1</mn> </msup> <mo>,</mo> <mo>…</mo> <mo>,</mo> <msup> <mi>q</mi> <mi>n</mi> </msup> <mo>)</mo> </mrow> <mi>T</mi> </msup> </mrow> </semantics></math> are the coordinates of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>–</mo> </mrow> </semantics></math>extensive properties. <math display="inline"><semantics> <mi>γ</mi> </semantics></math> are the partial derivatives of entropy that correspond to the intensive properties of the system. With the inclusion of gauge <math display="inline"><semantics> <msub> <mi>P</mi> <mi>S</mi> </msub> </semantics></math>, the conjugate momenta <span class="html-italic">p</span> are defined with respect to which homogeneous Hamiltonians, <math display="inline"><semantics> <msub> <mi mathvariant="script">H</mi> <mi>e</mi> </msub> </semantics></math>, of first-degree are determined. Details are given in the text.</p>
Full article ">Figure 5
<p>Potential energy surface of the Hénon–Heiles model and isocontours in the configuration plane.</p>
Full article ">Figure 6
<p>(<b>a</b>) A dissociating trajectory of a time-dependent Hénon–Heiles potential. (<b>b</b>) The trajectory is initialized from a periodic orbit (red thick line) and with energy <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">H</mi> <mi>d</mi> </msub> <mo>=</mo> <mn>0.1575</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>(<b>a</b>) The initial wavepacket centered at the minimum of the potential well. (<b>b</b>) The evolved wavepacket after 2000 time units.</p>
Full article ">Figure 8
<p>(<b>a</b>) The autocorrelation function for the initial Gaussian wavepacket. (<b>b</b>) The power spectrum was obtained by taking the Fourier transformation of the autocorrelation function.</p>
Full article ">Figure 9
<p>Projections of two representative trajectories in the Hénon–Heiles phase space are shown with friction parameters <math display="inline"><semantics> <mrow> <msub> <mi>b</mi> <mn>11</mn> </msub> <mo>=</mo> <mn>0.3</mn> <mspace width="0.277778em"/> <mi>and</mi> <mspace width="0.277778em"/> <msub> <mi>b</mi> <mn>22</mn> </msub> <mo>=</mo> <mn>0.001</mn> </mrow> </semantics></math>; (<b>a</b>) in the <math display="inline"><semantics> <mrow> <mo>(</mo> <msup> <mi>σ</mi> <mn>1</mn> </msup> <mo>,</mo> <msup> <mi>σ</mi> <mn>2</mn> </msup> <mo>,</mo> <msub> <mi>π</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> </semantics></math> space and (<b>b</b>) in the <math display="inline"><semantics> <mrow> <mo>(</mo> <msup> <mi>σ</mi> <mn>1</mn> </msup> <mo>,</mo> <msup> <mi>σ</mi> <mn>2</mn> </msup> <mo>,</mo> <msub> <mi>π</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </semantics></math> space.</p>
Full article ">Figure 10
<p>Panel (<b>a</b>) is the time evolution of the entropy production and (<b>b</b>) the trajectory length calculated with the Ruppeiner metric for the two trajectories shown in <a href="#entropy-26-00399-f009" class="html-fig">Figure 9</a>.</p>
Full article ">Figure A1
<p>(<b>a</b>) Concentrations of the constituent chemical species as functions of time, (<b>b</b>) reaction metric, (<b>c</b>) reaction coordinates <math display="inline"><semantics> <mrow> <mo>(</mo> <msup> <mi>ξ</mi> <mn>1</mn> </msup> <mo>,</mo> <msup> <mi>ξ</mi> <mn>2</mn> </msup> <mo>)</mo> </mrow> </semantics></math>, (<b>d</b>) conjugate momenta <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>p</mi> <msup> <mi>ξ</mi> <mn>1</mn> </msup> </msub> <mo>,</mo> <msub> <mi>p</mi> <msup> <mi>ξ</mi> <mn>2</mn> </msup> </msub> <mo>)</mo> </mrow> </semantics></math> to reaction coordinates. The quantities have been calculated from a trajectory run with initial concentrations, <math display="inline"><semantics> <mrow> <mo>[</mo> <msub> <mi mathvariant="normal">Q</mi> <mn>1</mn> </msub> <mo>]</mo> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mo>[</mo> <msub> <mi mathvariant="normal">Q</mi> <mn>2</mn> </msub> <mo>]</mo> <mo>=</mo> <mn>0.03</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mo>[</mo> <mi mathvariant="normal">P</mi> <mo>]</mo> <mo>=</mo> <mn>0.001</mn> </mrow> </semantics></math>. The rate constants for the forward reactions are taken equal to <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mrow> <mi>f</mi> <mn>1</mn> </mrow> </msub> <mo>=</mo> <mn>1.5</mn> <mo>,</mo> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mrow> <mi>f</mi> <mn>2</mn> </mrow> </msub> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics></math>, and the backward reactions <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mrow> <mi>b</mi> <mn>1</mn> </mrow> </msub> <mo>=</mo> <mn>0.5</mn> <mo>,</mo> <mo> </mo> <msub> <mi>k</mi> <mrow> <mi>b</mi> <mn>2</mn> </mrow> </msub> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math>.</p>
Full article ">
14 pages, 448 KiB  
Article
Estimating Quantum Mutual Information of Continuous-Variable Quantum States by Measuring Purity and Covariance Matrix
by Jiyong Park
Entropy 2022, 24(7), 940; https://doi.org/10.3390/e24070940 - 6 Jul 2022
Cited by 2 | Viewed by 3516
Abstract
We derive accessible upper and lower bounds for continuous-variable (CV) quantum states on quantum mutual information. The derivations are based on the observation that some functions of purities bound the difference between quantum mutual information of a quantum state and its Gaussian reference. [...] Read more.
We derive accessible upper and lower bounds for continuous-variable (CV) quantum states on quantum mutual information. The derivations are based on the observation that some functions of purities bound the difference between quantum mutual information of a quantum state and its Gaussian reference. The bounds are efficiently obtainable by measuring purities and the covariance matrix without multimode quantum state reconstruction. We extend our approach to the upper and lower bounds for the quantum total correlation of CV multimode quantum states. Furthermore, we investigate the relations of the bounds for the quantum mutual information with the bounds for the quantum conditional entropy. Full article
(This article belongs to the Section Quantum Information)
Show Figures

Figure 1

Figure 1
<p>Quantum mutual information <math display="inline"><semantics> <mrow> <mi mathvariant="script">I</mi> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </semantics></math> (black solid line), its lower and upper bounds <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">I</mi> <mo>−</mo> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (red dashed line), and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">I</mi> <mo>+</mo> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (blue dot-dashed line), respectively, and the maximum of the classical mutual information extractable by joint homodyne detection <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">C</mi> <mi>HD</mi> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (purple dotted line) for the two-mode squeezed thermal state with <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>n</mi> <mo>¯</mo> </mover> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> against the squeezing parameter <span class="html-italic">r</span>. Note that one nat is equal to <math display="inline"><semantics> <mrow> <mfrac> <mn>1</mn> <mrow> <mo form="prefix">ln</mo> <mn>2</mn> </mrow> </mfrac> <mo>≃</mo> <mn>1.44</mn> </mrow> </semantics></math> bits.</p>
Full article ">Figure 2
<p>Quantum mutual information <math display="inline"><semantics> <mrow> <mi mathvariant="script">I</mi> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </semantics></math> (black solid line), its lower and upper bounds <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">I</mi> <mo>−</mo> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (red dashed line), and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">I</mi> <mo>+</mo> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (blue dot-dashed line), respectively, and the maximum of the classical mutual information extractable by joint homodyne detection <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">C</mi> <mi>HD</mi> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (purple dotted line) for the pair coherent state <math display="inline"><semantics> <mrow> <mrow> <mi>ρ</mi> <mo>=</mo> <mo>|</mo> </mrow> <msub> <mi>φ</mi> <mi>x</mi> </msub> <mrow> <mo>〉</mo> <mo>〈</mo> </mrow> <msub> <mi>φ</mi> <mi>x</mi> </msub> <mrow> <mo>|</mo> </mrow> </mrow> </semantics></math> against the parameter <span class="html-italic">x</span>.</p>
Full article ">Figure 3
<p>Quantum mutual information <math display="inline"><semantics> <mrow> <mi mathvariant="script">I</mi> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </semantics></math> (black solid line), its lower and upper bounds <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">I</mi> <mo>−</mo> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (red dashed line), and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">I</mi> <mo>+</mo> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (blue dot-dashed line), respectively, and the maximum of the classical mutual information extractable by joint homodyne detection <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">C</mi> <mi>HD</mi> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (purple dotted line) for the CV Werner state <math display="inline"><semantics> <mrow> <mrow> <mi>ρ</mi> <mo>=</mo> <mi>f</mi> <mo>|</mo> </mrow> <msub> <mi>ψ</mi> <mi>r</mi> </msub> <mrow> <mo>〉</mo> <mo>〈</mo> </mrow> <msub> <mi>ψ</mi> <mi>r</mi> </msub> <mrow> <mo>|</mo> <mo>+</mo> <mrow> <mo>(</mo> <mn>1</mn> <mo>−</mo> <mi>f</mi> <mo>)</mo> </mrow> <mo>|</mo> <mn>0</mn> <mo>〉</mo> <mo>〈</mo> <mn>0</mn> <mo>|</mo> <mo>⊗</mo> <mo>|</mo> <mn>0</mn> <mo>〉</mo> <mo>〈</mo> <mn>0</mn> <mo>|</mo> </mrow> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> against the fraction <span class="html-italic">f</span>.</p>
Full article ">Figure 4
<p>Quantum total correlation <math display="inline"><semantics> <mrow> <mi mathvariant="script">T</mi> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </semantics></math> (black solid line), its lower and upper bounds <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">T</mi> <mo>−</mo> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (red dashed line), and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="script">T</mi> <mo>+</mo> </msub> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (blue dot-dashed line), respectively, and the quantum total correlation of the reference Gaussian state <math display="inline"><semantics> <mrow> <mi mathvariant="script">T</mi> <mo>(</mo> <msub> <mi>ρ</mi> <mi mathvariant="normal">G</mi> </msub> <mo>)</mo> </mrow> </semantics></math> (gray dotted line) for <span class="html-italic">M</span>-mode entangled coherent states <math display="inline"><semantics> <mrow> <mrow> <mi>ρ</mi> <mo>=</mo> <mo>|</mo> </mrow> <msub> <mo>Ψ</mo> <mi>M</mi> </msub> <mrow> <mo>〉</mo> <mo>〈</mo> </mrow> <msub> <mo>Ψ</mo> <mi>M</mi> </msub> <mrow> <mo>|</mo> </mrow> </mrow> </semantics></math> with (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, and (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> against the coherent amplitude <math display="inline"><semantics> <mi>γ</mi> </semantics></math>.</p>
Full article ">Figure 5
<p>Quantum conditional entropy <math display="inline"><semantics> <mrow> <mi>S</mi> <msub> <mrow> <mo>(</mo> <mn>1</mn> <mo>|</mo> <mn>2</mn> <mo>)</mo> </mrow> <mi>ρ</mi> </msub> </mrow> </semantics></math> (black solid line), its upper and lower bounds <math display="inline"><semantics> <mrow> <msub> <mi>S</mi> <mo>+</mo> </msub> <msub> <mrow> <mo>(</mo> <mn>1</mn> <mo>|</mo> <mn>2</mn> <mo>)</mo> </mrow> <mi>ρ</mi> </msub> </mrow> </semantics></math> (red dashed line), and <math display="inline"><semantics> <mrow> <msub> <mi>S</mi> <mo>−</mo> </msub> <msub> <mrow> <mo>(</mo> <mn>1</mn> <mo>|</mo> <mn>2</mn> <mo>)</mo> </mrow> <mi>ρ</mi> </msub> </mrow> </semantics></math> (blue dot-dashed line), respectively, and <math display="inline"><semantics> <mrow> <mo>−</mo> <mi>S</mi> <mo>(</mo> <msub> <mi>ρ</mi> <mrow> <mn>2</mn> <mo>,</mo> <mi mathvariant="normal">G</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> (purple dotted line) for the CV Werner state <math display="inline"><semantics> <mrow> <mrow> <mi>ρ</mi> <mo>=</mo> <mi>f</mi> <mo>|</mo> </mrow> <msub> <mi>ψ</mi> <mi>r</mi> </msub> <mrow> <mo>〉</mo> <mo>〈</mo> </mrow> <msub> <mi>ψ</mi> <mi>r</mi> </msub> <mrow> <mo>|</mo> <mo>+</mo> <mrow> <mo>(</mo> <mn>1</mn> <mo>−</mo> <mi>f</mi> <mo>)</mo> </mrow> <mo>|</mo> <mn>0</mn> <mo>〉</mo> <mo>〈</mo> <mn>0</mn> <mo>|</mo> <mo>⊗</mo> <mo>|</mo> <mn>0</mn> <mo>〉</mo> <mo>〈</mo> <mn>0</mn> <mo>|</mo> </mrow> </mrow> </semantics></math> against the fraction <span class="html-italic">f</span>.</p>
Full article ">
34 pages, 11427 KiB  
Article
Categorical Exploratory Data Analysis: From Multiclass Classification and Response Manifold Analytics Perspectives of Baseball Pitching Dynamics
by Fushing Hsieh and Elizabeth P. Chou
Entropy 2021, 23(7), 792; https://doi.org/10.3390/e23070792 - 22 Jun 2021
Cited by 7 | Viewed by 3530
Abstract
All features of any data type are universally equipped with categorical nature revealed through histograms. A contingency table framed by two histograms affords directional and mutual associations based on rescaled conditional Shannon entropies for any feature-pair. The heatmap of the mutual association matrix [...] Read more.
All features of any data type are universally equipped with categorical nature revealed through histograms. A contingency table framed by two histograms affords directional and mutual associations based on rescaled conditional Shannon entropies for any feature-pair. The heatmap of the mutual association matrix of all features becomes a roadmap showing which features are highly associative with which features. We develop our data analysis paradigm called categorical exploratory data analysis (CEDA) with this heatmap as a foundation. CEDA is demonstrated to provide new resolutions for two topics: multiclass classification (MCC) with one single categorical response variable and response manifold analytics (RMA) with multiple response variables. We compute visible and explainable information contents with multiscale and heterogeneous deterministic and stochastic structures in both topics. MCC involves all feature-group specific mixing geometries of labeled high-dimensional point-clouds. Upon each identified feature-group, we devise an indirect distance measure, a robust label embedding tree (LET), and a series of tree-based binary competitions to discover and present asymmetric mixing geometries. Then, a chain of complementary feature-groups offers a collection of mixing geometric pattern-categories with multiple perspective views. RMA studies a system’s regulating principles via multiple dimensional manifolds jointly constituted by targeted multiple response features and selected major covariate features. This manifold is marked with categorical localities reflecting major effects. Diverse minor effects are checked and identified across all localities for heterogeneity. Both MCC and RMA information contents are computed for data’s information content with predictive inferences as by-products. We illustrate CEDA developments via Iris data and demonstrate its applications on data taken from the PITCHf/x database. Full article
(This article belongs to the Special Issue Information Complexity in Structured Data)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Threespecies of Iris: Setosa (red), Versicolor (green), and Virginica (blue); (<b>B</b>) four features’ possibly gapped histograms; and (<b>C</b>) corresponding piecewise linear approximations on empirical distribution functions.</p>
Full article ">Figure 2
<p>Three contingency tables (row-vs.-column) for mixing geometries: (<b>A</b>) petal width-vs.-petal length; (<b>B</b>) sepal length-vs.-(petal length, petal width); and (<b>C</b>) (sepal length, sepal width)-vs.-(petal length, petal width). Within each cell, the triplet of numbers is: #Setosa/#Versicolor/#Virginica.</p>
Full article ">Figure 3
<p>(<b>A</b>) Mutualconditional entropy matrices and (<b>B</b>) marked synergistic feature-groups on its heatmap.</p>
Full article ">Figure 4
<p>Three mutual conditional entropy matrices with marked synergistic feature-sets: (<b>A</b>) slider; (<b>B</b>) curveball; and (<b>C</b>) fastball.</p>
Full article ">Figure 5
<p>Slider’s four 3D geometries of five triplets of features (url-address for rotatable 3D plot): (<b>A</b>) {“z0”, “x0”, “vx0”}; (<b>B</b>) {“x0”, “start_speed”, “end_speed”}; (<b>C</b>) {“spin_rate”, “start_speed”, “end_speed”}; and (<b>D</b>) {“pfx_x”, “spin_dir”, “pfx_z”}. See corresponding rotatable 3D plots at <a href="https://rpubs.com/CEDA/baseball" target="_blank">https://rpubs.com/CEDA/baseball</a> (accessed on 17 June 2021).</p>
Full article ">Figure 6
<p>Relative distance among three Iris species’ (petal-length, petal-width) point-clouds based on 1000 triplet competitions: (<b>A</b>) label-pair dominance matrix; (<b>B</b>) relative distance matrix of three labels (species); and (<b>C</b>) label-embedding tree.</p>
Full article ">Figure 7
<p>Three pie charts of the Iris example.</p>
Full article ">Figure 8
<p>Global geometry of point-cloud based on Feature-Group C: (<b>A</b>) distance matrix of three sets of randomly selected data points from three pitchers: 453286 (red), 543243 (green) and 596112 (blue); and (<b>B</b>) LET (label embedding tree) of five pitchers.</p>
Full article ">Figure 9
<p>Predictive map of five slider pitchers based on Feature-Set C with threshold range [0.65, 100/65] for pseudo-likelihood values via K (=20) nearest neighbors.</p>
Full article ">Figure 10
<p>Slider’s two-label embedding tree and two predictive maps (with threshold value 1): (<b>A</b>,<b>C</b>) Feature-Set A; and (<b>B</b>,<b>D</b>) 19 features.</p>
Full article ">Figure 11
<p>Curveball’s MCC setting with threshold range <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>]</mo> </mrow> </semantics></math> with respect to feature-group: DEF: (<b>A</b>) label embedding tree of feature set DEF; and (<b>B</b>) predictive graphs and predictive matrix.</p>
Full article ">Figure 12
<p>Fastball’s two manifolds: (<b>A</b>) {“pfx_x”, “pfx_z”, “ spin_dir”}; and (<b>B</b>) {“x0”, “start_speed”, “end_speed”} (see corresponding rotatable 3D plots at <a href="https://rpubs.com/CEDA/baseball" target="_blank">https://rpubs.com/CEDA/baseball</a> (accessed on 17 June 2021)).</p>
Full article ">Figure 13
<p>Manifolds of {“pfx_x”, “pfx_z”, “spin_dir”} with marked strips: (<b>A</b>) three strips of “spin_rate”; and (<b>B</b>) three strips of “spin_rate” intersecting with three strips “spin_dir” (see corresponding rotatable 3D plots at <a href="https://rpubs.com/CEDA/baseball" target="_blank">https://rpubs.com/CEDA/baseball</a> (accessed on 17 June 2021)).</p>
Full article ">Figure 14
<p>Manifolds of {“pfx_x”, “pfx_z”, “spin_dir”} with marked strips: (<b>A</b>) three strips of “start_speed”; and (<b>B</b>) three strips of “start_speed” intersecting with three strips “spin_dir” (see corresponding rotatable 3D plots at (<a href="https://rpubs.com/CEDA/baseball" target="_blank">https://rpubs.com/CEDA/baseball</a> (accessed on 17 June 2021)).</p>
Full article ">Figure 15
<p>Minor features selection based Shannon entropy on nine patches on manifolds of {“pfx_x”, “pfx_z”, “spin_dir”} framed by nine intersecting patches via three strips of “spin_dir” with three strips “spin_dir”, as seen in <a href="#entropy-23-00792-f013" class="html-fig">Figure 13</a>.</p>
Full article ">
28 pages, 7583 KiB  
Article
Mimicking Complexity of Structured Data Matrix’s Information Content: Categorical Exploratory Data Analysis
by Fushing Hsieh, Elizabeth P. Chou and Ting-Li Chen
Entropy 2021, 23(5), 594; https://doi.org/10.3390/e23050594 - 11 May 2021
Cited by 6 | Viewed by 2517
Abstract
We develop Categorical Exploratory Data Analysis (CEDA) with mimicking to explore and exhibit the complexity of information content that is contained within any data matrix: categorical, discrete, or continuous. Such complexity is shown through visible and explainable serial multiscale structural dependency with heterogeneity. [...] Read more.
We develop Categorical Exploratory Data Analysis (CEDA) with mimicking to explore and exhibit the complexity of information content that is contained within any data matrix: categorical, discrete, or continuous. Such complexity is shown through visible and explainable serial multiscale structural dependency with heterogeneity. CEDA is developed upon all features’ categorical nature via histogram and it is guided by all features’ associative patterns (order-2 dependence) in a mutual conditional entropy matrix. Higher-order structural dependency of k(3) features is exhibited through block patterns within heatmaps that are constructed by permuting contingency-kD-lattices of counts. By growing k, the resultant heatmap series contains global and large scales of structural dependency that constitute the data matrix’s information content. When involving continuous features, the principal component analysis (PCA) extracts fine-scale information content from each block in the final heatmap. Our mimicking protocol coherently simulates this heatmap series by preserving global-to-fine scales structural dependency. Upon every step of mimicking process, each accepted simulated heatmap is subject to constraints with respect to all of the reliable observed categorical patterns. For reliability and robustness in sciences, CEDA with mimicking enhances data visualization by revealing deterministic and stochastic structures within each scale-specific structural dependency. For inferences in Machine Learning (ML) and Statistics, it clarifies, upon which scales, which covariate feature-groups have major-vs.-minor predictive powers on response features. For the social justice of Artificial Intelligence (AI) products, it checks whether a data matrix incompletely prescribes the targeted system. Full article
(This article belongs to the Special Issue Information Complexity in Structured Data)
Show Figures

Figure 1

Figure 1
<p>Two pictorial Illustrations: (<b>A</b>) strike “zone” in MLB; (<b>B</b>) Magnus effect for fastball having the back-spin and up-ward force, see <a href="https://upload.wikimedia.org/wikipedia/commons/I/15/Sketch_of_Magnus_effect" target="_blank">https://upload.wikimedia.org/wikipedia/commons/I/15/Sketch_of_Magnus_effect</a> (accessed on 8 May 2021).</p>
Full article ">Figure 2
<p>(<b>A</b>) Heatmap of <math display="inline"><semantics> <mrow> <mn>10</mn> <mo>×</mo> <mn>10</mn> </mrow> </semantics></math> MCE matrix (with a HC tree using the single linkage module) and (<b>B</b>) a directed network of 10 nodes with thin and thick edges for directional conditional entropy falling in threshold regions <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>0.9</mn> <mo>,</mo> <mn>0.95</mn> <mo>]</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>0.9</mn> <mo>]</mo> </mrow> </semantics></math>, respectively.</p>
Full article ">Figure 3
<p>Contingency tables marked with constraints with <math display="inline"><semantics> <mrow> <mn>95</mn> <mo>%</mo> </mrow> </semantics></math> reliability: (<b>A</b>) Zone-vs.-Batting-results; (<b>B</b>) Strike-and-Ball-vs. Zone; and, (<b>C</b>) Batting-results-vs.-Strike-and-Ball. (the darker color, the higher order.)</p>
Full article ">Figure 4
<p>The <math display="inline"><semantics> <mrow> <mn>13</mn> <mo>×</mo> <mn>95</mn> </mrow> </semantics></math> contingency table of the zone feature against the bivariate feature Ball-vs.-Strike count and Batting result.</p>
Full article ">Figure 5
<p>Piecewise linear approximations via histograms of four features related to Magnus effects: (<b>A</b>) “pfx_x”; (<b>B</b>) “pfx_z”; (<b>C</b>) “spin_dir”; and, (<b>D</b>) “spin_rate”.</p>
Full article ">Figure 6
<p>The 4D manifold of 4 features related to Magnus effects. The color-coding scheme is the 9 bins of histogram of “spin_dir”: (<b>A</b>) front view; (<b>B</b>) back view. See corresponding rotatable 3D plots through the link: <a href="https://rpubs.com/CEDA/Mimicking" target="_blank">https://rpubs.com/CEDA/Mimicking</a> (accessed on 8 May 2021).</p>
Full article ">Figure 7
<p>Two <math display="inline"><semantics> <mrow> <mn>9</mn> <mo>×</mo> <mn>9</mn> </mrow> </semantics></math> contingency tables marked with <math display="inline"><semantics> <mrow> <mn>95</mn> <mo>%</mo> </mrow> </semantics></math> pattern-based rules: (<b>A</b>) {“pfx_x”, “spin_dir”}; (<b>B</b>) {“pfx_z”, “spin_rate”}. Pattern-based rules are marked along columns: darker colors for larger counts.</p>
Full article ">Figure 8
<p>Contingency-4D-lattice of 4 features governing Magnus effects: { “pfx_x”, “spin_dir” }-vs.- { “pfx_z”, “spin_rate”}.</p>
Full article ">Figure 9
<p>Eigenvalues and eigenvectors form six color-coded blocks.</p>
Full article ">Figure 10
<p>Individual histograms of <math display="inline"><semantics> <mrow> <msubsup> <mover accent="true"> <mi>ε</mi> <mo>˜</mo> </mover> <mi>i</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> <mo>,</mo> <msubsup> <mover accent="true"> <mi>ε</mi> <mo>˜</mo> </mover> <mi>i</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>,</mo> <msubsup> <mover accent="true"> <mi>ε</mi> <mo>˜</mo> </mover> <mi>i</mi> <mrow> <mo>(</mo> <mn>3</mn> <mo>)</mo> </mrow> </msubsup> <mo>,</mo> <msubsup> <mover accent="true"> <mi>ε</mi> <mo>˜</mo> </mover> <mi>i</mi> <mrow> <mo>(</mo> <mn>4</mn> <mo>)</mo> </mrow> </msubsup> </mrow> </semantics></math> pertaining to the 6 color-coded blocks: (<b>A</b>) Yellow; (<b>B</b>) Orange; (<b>C</b>) Green; (<b>D</b>) Blue; (<b>E</b>) Pink; and, (<b>F</b>) Purple.</p>
Full article ">Figure 11
<p>Snapshots of 3D plots of {“pfx_x”, “pfx_z”, “spin_rate”}: <math display="inline"><semantics> <msubsup> <mrow> <mi mathvariant="script">C</mi> </mrow> <mo>□</mo> <mo>*</mo> </msubsup> </semantics></math> (dark-colored-points) and <math display="inline"><semantics> <msub> <mi mathvariant="script">C</mi> <mo>□</mo> </msub> </semantics></math> (light-colored-points) the six color-coded mimicked blocks: (<b>A</b>) local view; (<b>B</b>) global view. Block-specific color-coding: Yellow; Orange; Green; Blue; Pink; Purple. See the corresponding rotatable 3D plots through the link: <a href="https://rpubs.com/CEDA/Mimicking" target="_blank">https://rpubs.com/CEDA/Mimicking</a> (accessed on 8 May 2021).</p>
Full article ">Figure 12
<p>(<b>A</b>) Histogram of spin-pair (“spin_dir”, “spin_rate”) to (“pfx_x”, “pfx_z”) correspondence among 293 pieces of 4D hypercubes. (<b>B</b>) The snapshot of 3D plot of {“VX0”, “X0”, “Z0”} for data points belonging to five 4D hypercubes sharing the spin-pair (5,7). See the corresponding rotatable 3D plots through the link: <a href="https://rpubs.com/CEDA/Mimicking" target="_blank">https://rpubs.com/CEDA/Mimicking</a> (accessed on 8 May 2021).</p>
Full article ">Figure 13
<p>Two contingency tables of two feature-pairs: {“zone”,“VX0”} and {“X0”, “Z0”}. Observed column-wise ordering rules with 95% of reliability are marked. The collective acceptance rate is about 56% with respect to simply random sampling.</p>
Full article ">Figure 14
<p>The <math display="inline"><semantics> <mrow> <mn>127</mn> <mo>×</mo> <mn>96</mn> </mrow> </semantics></math> contingency-4D-lattice of two feature-pairs: {“X0”, “Z0”} vs. {“zone”,“VX0”}. The row-axis is superimposed with a HC-tree derived from <math display="inline"><semantics> <mrow> <mn>127</mn> <mo>×</mo> <mn>127</mn> </mrow> </semantics></math> adjacency matrix under eight-neighbor-neighborhood system. These five blocks are marked across the four peripheral zones.</p>
Full article ">Figure 15
<p>Geometry of five mimicked blocks marked on the heatmap of <math display="inline"><semantics> <mrow> <mn>127</mn> <mo>×</mo> <mn>96</mn> </mrow> </semantics></math> contingency-4D-lattice of two feature-pairs: {“X0”, “Z0”} vs. {“zone”,“VX0”}, as shown in <a href="#entropy-23-00594-f014" class="html-fig">Figure 14</a>. These 5 blocks are marked across the 4 peripheral zones. See corresponding rotatable 3D plots through the link: <a href="https://rpubs.com/CEDA/Mimicking" target="_blank">https://rpubs.com/CEDA/Mimicking</a> (accessed on 8 May 2021).</p>
Full article ">
417 KiB  
Proceeding Paper
Spatiotemporal Graph Imaging Associated with Multilevel Atomic Excitations
by Abu Mohamed Alhasan
Proceedings 2020, 67(1), 16; https://doi.org/10.3390/ASEC2020-07886 - 11 Nov 2020
Viewed by 1138
Abstract
In this paper, we establish a graph imaging technique to manifest local stabilization within atomic systems of multiple levels. Specifically, we address the interrelation between local stabilization and image entropy. As an example, we consider the mutual interaction of two pair of pulses [...] Read more.
In this paper, we establish a graph imaging technique to manifest local stabilization within atomic systems of multiple levels. Specifically, we address the interrelation between local stabilization and image entropy. As an example, we consider the mutual interaction of two pair of pulses propagating in a double-Λ configuration. Thus, we have two different sets of two pulses that share the same shape and phase, initially. The first (second) set belongs to lower (upper) -Λ subsystems, respectively. The configuration of two pair of pulses is considered as a dynamical graph model with four nodes. The dynamic transition matrix describes the connectivity matrix in the static graph model. It is to be emphasized that the graph and its image have the same transition matrix. In particular, the graph model exposes the stabilization in terms of the singular-value decomposition of energies for the transition matrix, that is, irrespectively of the structure of the transition matrix. The image model of the graph displays the details of the matrix structure in terms of row and column probabilities. Therefore, it enables one to study conditional probabilities and mutual information inherent in the network of the graph. Furthermore, the graph imaging provides the main row/column contribution to the transition matrix in terms of image entropy. Our results show that image entropy exposes spatial dependence, which is irrelevant to graph entropy. Full article
(This article belongs to the Proceedings of The 1st International Electronic Conference on Applied Sciences)
Show Figures

Figure 1

Figure 1
<p>Energy-level diagram of the <math display="inline"> <semantics> <mrow> <msup> <mrow/> <mn>87</mn> </msup> <mi>R</mi> <mi>b</mi> </mrow> </semantics> </math><math display="inline"> <semantics> <msub> <mi>D</mi> <mn>1</mn> </msub> </semantics> </math> line including the hyperfine (hf) structure. The Rabi frequencies <math display="inline"> <semantics> <msub> <mo>Ω</mo> <mrow> <mi>i</mi> <mi>j</mi> </mrow> </msub> </semantics> </math> couple the dipole-allowed transition <math display="inline"> <semantics> <mrow> <mrow> <mrow> <mo>|</mo> </mrow> <msub> <mi>F</mi> <mi>i</mi> </msub> <mrow> <mo>〉</mo> </mrow> </mrow> <mo>⇔</mo> <mrow> <mrow> <mo>|</mo> </mrow> <msub> <mi>F</mi> <mi>j</mi> </msub> <mrow> <mo>〉</mo> </mrow> </mrow> </mrow> </semantics> </math>, where <math display="inline"> <semantics> <mi>α</mi> </semantics> </math> and <math display="inline"> <semantics> <mi>β</mi> </semantics> </math> denote the level label with <math display="inline"> <semantics> <mrow> <mi>i</mi> <mo>=</mo> <mo>{</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> <mo>}</mo> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>j</mi> <mo>=</mo> <mo>{</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>}</mo> </mrow> </semantics> </math>, respectively. The number <span class="html-italic">F</span> denotes the total angular momentum quantum number associated with the hf level.</p>
Full article ">Figure 2
<p>Complete graph with four nodes and six edges. The edges are formed by dipole-allowed and forbidden transitions. <math display="inline"> <semantics> <msubsup> <mi>S</mi> <mrow> <mi>i</mi> <mi>j</mi> </mrow> <mn>2</mn> </msubsup> </semantics> </math> represents the Bloch metric associated with the transition <math display="inline"> <semantics> <mrow> <mrow> <mo>|</mo> <mi>i</mi> <mo>〉</mo> </mrow> <mo>↔</mo> <mrow> <mo>|</mo> <mi>j</mi> <mo>〉</mo> </mrow> </mrow> </semantics> </math>.</p>
Full article ">Figure 3
<p>Spatial mutual information for different transition matrixes. The links are represented by Bloch metrics between dipole-allowed and forbidden transitions; <math display="inline"> <semantics> <mrow> <mi>k</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>3</mn> </msup> </mrow> </semantics> </math>.</p>
Full article ">Figure 4
<p>The spatial cross-section in relative units that is associated with the total light scattering. The cross-section is normalized to its maximum in the course of propagation; <math display="inline"> <semantics> <mrow> <mi>k</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>3</mn> </msup> </mrow> </semantics> </math>.</p>
Full article ">
10 pages, 281 KiB  
Article
Simple Stopping Criteria for Information Theoretic Feature Selection
by Shujian Yu and José C. Príncipe
Entropy 2019, 21(1), 99; https://doi.org/10.3390/e21010099 - 21 Jan 2019
Cited by 8 | Viewed by 4967
Abstract
Feature selection aims to select the smallest feature subset that yields the minimum generalization error. In the rich literature in feature selection, information theory-based approaches seek a subset of features such that the mutual information between the selected features and the class labels [...] Read more.
Feature selection aims to select the smallest feature subset that yields the minimum generalization error. In the rich literature in feature selection, information theory-based approaches seek a subset of features such that the mutual information between the selected features and the class labels is maximized. Despite the simplicity of this objective, there still remain several open problems in optimization. These include, for example, the automatic determination of the optimal subset size (i.e., the number of features) or a stopping criterion if the greedy searching strategy is adopted. In this paper, we suggest two stopping criteria by just monitoring the conditional mutual information (CMI) among groups of variables. Using the recently developed multivariate matrix-based Rényi’s α-entropy functional, which can be directly estimated from data samples, we showed that the CMI among groups of variables can be easily computed without any decomposition or approximation, hence making our criteria easy to implement and seamlessly integrated into any existing information theoretic feature selection methods with a greedy search strategy. Full article
(This article belongs to the Special Issue Information Theoretic Learning and Kernel Methods)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) shows the the values of mutual information (MI) <math display="inline"><semantics> <mrow> <mi mathvariant="bold">I</mi> <mo>(</mo> <msup> <mi>S</mi> <mo>′</mo> </msup> <mo>;</mo> <mi>y</mi> <mo>)</mo> </mrow> </semantics></math> and conditional mutual information (CMI) <math display="inline"><semantics> <mrow> <mi mathvariant="bold">I</mi> <mo>(</mo> <mrow> <mo>{</mo> <mi>S</mi> <mo>−</mo> <msup> <mi>S</mi> <mo>′</mo> </msup> <mo>}</mo> </mrow> <mo>;</mo> <mi>y</mi> <mo>|</mo> <msup> <mi>S</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> </semantics></math> with respect to different number of selected features, i.e., the size of <math display="inline"><semantics> <msup> <mi>S</mi> <mo>′</mo> </msup> </semantics></math>. <math display="inline"><semantics> <mrow> <mi mathvariant="bold">I</mi> <mo>(</mo> <msup> <mi>S</mi> <mo>′</mo> </msup> <mo>;</mo> <mi>y</mi> <mo>)</mo> </mrow> </semantics></math> is monotonically increasing, whereas <math display="inline"><semantics> <mrow> <mi mathvariant="bold">I</mi> <mo>(</mo> <mrow> <mo>{</mo> <mi>S</mi> <mo>−</mo> <msup> <mi>S</mi> <mo>′</mo> </msup> <mo>}</mo> </mrow> <mo>;</mo> <mi>y</mi> <mo>|</mo> <msup> <mi>S</mi> <mo>′</mo> </msup> <mo>)</mo> </mrow> </semantics></math> is monotonically decreasing. (<b>b</b>) shows the terminated points produced by different stopping criteria, namely CMI-heuristic (black solid line), CMI-permutation (black dashed line), <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>MI</mi> </mrow> </semantics></math>-<math display="inline"><semantics> <msup> <mi>χ</mi> <mn>2</mn> </msup> </semantics></math> (green solid line), and MI-permutation (blue solid line). The red curve with the shaded area indicates the average bootstrap classification accuracy with <math display="inline"><semantics> <mrow> <mn>95</mn> <mo>%</mo> </mrow> </semantics></math> confidence interval. In this example, the bootstrap classification accuracy reaches its statistical maximum value with 11 features and CMI-heuristic performs the best.</p>
Full article ">
Back to TopTop