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

Journals

Article Types

Countries / Regions

Search Results (105)

Search Parameters:
Keywords = pseudo-symmetry

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 293 KiB  
Article
The Quenched \({g_A}\) in Nuclei and Infrared Fixed Point in QCD
by Mannque Rho and Long-Qi Shao
Symmetry 2024, 16(12), 1704; https://doi.org/10.3390/sym16121704 (registering DOI) - 22 Dec 2024
Viewed by 48
Abstract
The possible consequence of an infrared (IR) fixed point in QCD for Nf=2,3 in nuclear matter is discussed. It is shown in terms of d(ilaton)-χ effective field theory (dχEFT) incorporated in a generalized effective field [...] Read more.
The possible consequence of an infrared (IR) fixed point in QCD for Nf=2,3 in nuclear matter is discussed. It is shown in terms of d(ilaton)-χ effective field theory (dχEFT) incorporated in a generalized effective field theory implemented with hidden local symmetry and hidden scale symmetry that the superallowed Gamow–Teller transition in the doubly magic-shell nucleus 100Sn recently measured at RIKEN indicates a large anomaly-induced quenching identified as a fundamental renormalization of gA from the free-space value of 1.276 to ≈0.8. Combined with the quenching expected from strong nuclear correlations “snc”, the effective coupling in nuclei gAeff would come to ∼1/2. If this result were reconfirmed, it would impact drastically not only nuclear structure and dense compact-star matter—where gA figures in π-N coupling via the Goldberger-Treiman relation—but also in search for physics Beyond the Standard Model (BSM), e.g., 0νββ decay, where the fourth power of gA figures. Full article
(This article belongs to the Special Issue Nuclear Symmetry Energy: From Finite Nuclei to Neutron Stars)
19 pages, 5695 KiB  
Article
Sparse Sensor Fusion for 3D Object Detection with Symmetry-Aware Colored Point Clouds
by Lele Wang, Peng Zhang, Ming Li and Faming Zhang
Symmetry 2024, 16(12), 1690; https://doi.org/10.3390/sym16121690 - 20 Dec 2024
Viewed by 360
Abstract
Multimodal fusion-based object detection is the foundational sensing task in scene understanding. It capitalizes on LiDAR and camera data to boost the robust results. However, there are still great challenges in establishing an effective fusion mechanism and performing accurate and diverse feature interaction [...] Read more.
Multimodal fusion-based object detection is the foundational sensing task in scene understanding. It capitalizes on LiDAR and camera data to boost the robust results. However, there are still great challenges in establishing an effective fusion mechanism and performing accurate and diverse feature interaction fusion. In particular, the relationship construction between the two modalities has not been comprehensively exploited, leading to sensor data utilization deficiencies and redundancies. In this paper, a novel 3D object-detection framework, namely a symmetry-aware sparse sensor fusion detection network (2SFNet), is proposed. This framework was designed to leverage point clouds and RGB images. The 2SFNet consists of three submodules, filtered colored point cloud generation, pseudo-image generation, and a dilated feature fusion network, to solve these problems. Firstly, filtered colored point cloud generation constructs non-ground colored point cloud (NCPC) data by employing an early fusion strategy and a ground-height-filtering module, selectively retaining only object-related information. Subsequently, 2D grid encoding is used on the reduced colored data. Finally, the processed colored data are fed into the improved PillarsNet architecture, which now has expanded receptive fields to enhance the fusion effect. This design optimizes the fusion process by ensuring a more balanced and effective data representation, aligning with the symmetry concept that underlies the model’s functionality. Experiments and evaluations were conducted on the KITTI dataset to present the effectuality, particularly for categories characterized by sparse point clouds. The results indicate that the symmetry-aware design of the 2SFNet leads to an improved performance when compared to other multimodal fusion networks, and alleviates the phenomenon caused by highly obscured and crowded scenes. Full article
(This article belongs to the Section Engineering and Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of early-level fusion, middle-level fusion, late-level fusion, and our fusion architectures.</p>
Full article ">Figure 2
<p>The proposed 2SFNet architecture.</p>
Full article ">Figure 3
<p>Schematic of filtered 7D colored point cloud generation. By means of a calibration matrix (calib.txt), the RGB pixels are cast onto corresponding points. R and T are the rotation matrix.</p>
Full article ">Figure 4
<p>Some visual examples of RGB image, raw point cloud, calibrated image, point cloud in image FOV, colored point cloud in image FOV, and corresponding filtered colored point cloud.</p>
Full article ">Figure 5
<p>1-Dilated Convolution, 2-Dilated Convolution and Framework of atrous convolution.</p>
Full article ">Figure 6
<p>Framework of the receptive field network.</p>
Full article ">Figure 7
<p>Visualization results for the 2SFNet model predictions and ground truths. The purple is the visualization result of the truth label and the blue is the visualization result of the network prediction.</p>
Full article ">Figure 8
<p>Visualization comparison for distant-object-detection results with PointPillars (<b>left</b>) and our 2SFNet (<b>right</b>). Notice that the predictions are entirely based on BEV maps derived from point clouds. Re-projecting to image space is for illustrative purposes only.</p>
Full article ">Figure 9
<p>Visualization comparison for partially occluded-object-detection results with PointPillars (<b>left</b>) and our 2SFNet (<b>right</b>). Notice that predictions are entirely based on BEV maps derived from point clouds. Re-projecting to image space is for illustrative purposes only.</p>
Full article ">
13 pages, 410 KiB  
Article
Parity Doubling in Dense Baryonic Matter as an Emergent Phenomenon and Pseudo-Conformal Phase
by Hyun Kyu Lee
Symmetry 2024, 16(12), 1598; https://doi.org/10.3390/sym16121598 - 30 Nov 2024
Viewed by 413
Abstract
The star matter composed of nucleons deep inside compact stars, such as neutron stars, is believed to be very dense, such that various types of new concepts and physical phenomena are naturally expected due to the nontrivial strong correlations between hadrons. The possibility [...] Read more.
The star matter composed of nucleons deep inside compact stars, such as neutron stars, is believed to be very dense, such that various types of new concepts and physical phenomena are naturally expected due to the nontrivial strong correlations between hadrons. The possibility of revealing the hidden scale symmetry in dense baryonic matter has been discussed recently, to uncover the pseudo-conformal phase in dense star matter. In the pseudo-conformal phase, the trace of the energy–momentum tensor becomes density-independent, and the speed of sound approaches the conformal velocity in scale symmetric matter. Interestingly, it is also observed that the effective nucleon mass becomes a density-independent finite quantity, which can be identified as the chiral invariant mass of the parity doublet model, indicating that the parity doubling is an emergent phenomenon. In this paper, we will discuss how parity-doubling symmetry emerges inside the core of a compact star as a consequence of the interplays between ω vector mesons and nucleons (or dilaton, χ, equivalently) and between the chiral symmetry and the scale symmetry. Full article
Show Figures

Figure 1

Figure 1
<p><math display="inline"><semantics> <mrow> <mo>−</mo> <mi>ϵ</mi> <mo>(</mo> <mi>n</mi> <mo>)</mo> <mo>+</mo> <mn>3</mn> <mi>P</mi> <mo>(</mo> <mi>n</mi> <mo>)</mo> <mo>=</mo> <mo>−</mo> <mi>TEMT</mi> </mrow> </semantics></math> vs. density for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (symmetric nuclear matter) and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (pure neutron matter).</p>
Full article ">Figure 2
<p>The ratio <math display="inline"><semantics> <mrow> <msubsup> <mi>m</mi> <mi>N</mi> <mo>*</mo> </msubsup> <mo>/</mo> <msub> <mi>m</mi> <mi>N</mi> </msub> <mo>≈</mo> <msup> <mrow> <mo>〈</mo> <mi>χ</mi> <mo>〉</mo> </mrow> <mo>*</mo> </msup> <mo>/</mo> <msub> <mrow> <mo>〈</mo> <mi>χ</mi> <mo>〉</mo> </mrow> <mn>0</mn> </msub> </mrow> </semantics></math> as a function of density for varying density dependence of <math display="inline"><semantics> <msubsup> <mi>g</mi> <mrow> <mi>V</mi> <mi>ω</mi> </mrow> <mo>*</mo> </msubsup> </semantics></math>. It stays more-or-less constant above that density. The Figure is borrowed from [<a href="#B18-symmetry-16-01598" class="html-bibr">18</a>].</p>
Full article ">
17 pages, 5612 KiB  
Article
Pseudo-Jahn–Teller Effect in Natural Compounds and Its Possible Role in Straintronics I: Hypericin and Its Analogs
by Dagmar Štellerová, Vladimír Lukeš and Martin Breza
Molecules 2024, 29(23), 5624; https://doi.org/10.3390/molecules29235624 - 28 Nov 2024
Viewed by 500
Abstract
The distortions and instability of high-symmetry configurations of polyatomic systems in nondegenerate states are usually ascribed to the pseudo-Jahn–Teller effect (PJTE). The geometries of hypericin, isohypericin, and fringelite D were optimized within various symmetry groups. Group-theoretical treatment and (TD-)DFT calculations were used to [...] Read more.
The distortions and instability of high-symmetry configurations of polyatomic systems in nondegenerate states are usually ascribed to the pseudo-Jahn–Teller effect (PJTE). The geometries of hypericin, isohypericin, and fringelite D were optimized within various symmetry groups. Group-theoretical treatment and (TD-)DFT calculations were used to identify the corresponding electronic states during the symmetry descent. The symmetry descent paths (up to the stable structures without imaginary vibrations) were determined using the corresponding imaginary vibrations as their kernel subgroups starting from the highest possible symmetry group. The vibronic interaction between the ground and excited electronic states relates to an increasing energy difference of both states during the symmetry decrease. This criterion was used to identify possible PJTE. We have shown that the PJTE in these naturally occurring compounds could explain only the symmetry descent paths C2v → C2 and C2v → Cs in hypericin, and the D2h → C2v, D2h → C2v → C2, and D2h → C2h ones in fringelite D. The electric dipole moments of hypericin and its analogs were determined prevailingly by the mutual orientations of the hydroxyl groups. The same held for the energies of frontier orbitals in these systems, but their changes during the symmetry descent were less significant. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Atom notation of hypericin (R<sub>1</sub> = R<sub>2</sub> = OH, R<sub>3</sub> = R<sub>4</sub> = CH<sub>3</sub>), isohypericin (R<sub>1</sub> = R<sub>3</sub> = OH, R<sub>2</sub> = R<sub>4</sub> = CH<sub>3</sub>) and fringelite D (R<sub>1</sub> = R<sub>2</sub> = R<sub>3</sub> = R<sub>4</sub> = OH).</p>
Full article ">Figure 2
<p>Optimized structures of hypericin Ia model systems (C—black, O—red, H—white).</p>
Full article ">Figure 3
<p>Optimized structures of hypericin Ib model systems (C—black, O—red, H—white).</p>
Full article ">Figure 4
<p>Optimized structures of hypericin Ic model systems (C—black, O—red, H—white).</p>
Full article ">Figure 5
<p>Optimized structures of isohypericin IIa model systems (C—black, O—red, H—white).</p>
Full article ">Figure 6
<p>Optimized structures of isohypericin IIb model systems (C—black, O—red, H—white).</p>
Full article ">Figure 7
<p>Optimized structures of hypericin IIc model systems (C—black, O—red, H—white).</p>
Full article ">Figure 8
<p>Optimized structures of fringelite D IIIa model systems (C—black, O—red, H—white).</p>
Full article ">Figure 9
<p>Optimized structures of fringelite D IIIb model systems (C—black, O—red, H—white).</p>
Full article ">Figure 10
<p>Optimized structures of fringelite D IIIc model systems (C—black, O—red, H—white).</p>
Full article ">Figure 11
<p>Optimized structures of fringelite D IIId model systems (C—black, O—red, H—white).</p>
Full article ">
25 pages, 504 KiB  
Article
Discrete Pseudo-Quasi Overlap Functions and Their Applications in Fuzzy Multi-Attribute Group Decision-Making
by Mei Jing, Jingqian Wang, Mei Wang and Xiaohong Zhang
Mathematics 2024, 12(22), 3569; https://doi.org/10.3390/math12223569 - 15 Nov 2024
Viewed by 746
Abstract
The overlap function, a continuous aggregation function, is widely used in classification, decision-making, image processing, etc. Compared to applications, overlap functions have also achieved fruitful results in theory, such as studies on the fundamental properties of overlap functions, various generalizations of the concept [...] Read more.
The overlap function, a continuous aggregation function, is widely used in classification, decision-making, image processing, etc. Compared to applications, overlap functions have also achieved fruitful results in theory, such as studies on the fundamental properties of overlap functions, various generalizations of the concept of overlap functions, and the construction of additive and multiplicative generators based on overlap functions. However, most of the research studies on the overlap functions mentioned above contain commutativity and continuity, which can limit their practical applications. In this paper, we remove the symmetry and continuity from overlap functions and define discrete pseudo-quasi overlap functions on finite chains. Meanwhile, we also discuss their related properties. Then, we introduce pseudo-quasi overlap functions on sub-chains and construct discrete pseudo-quasi overlap functions on finite chains using pseudo-quasi overlap functions on these sub-chain functions. Unlike quasi-overlap functions on finite chains generated by the ordinal sum, discrete pseudo-quasi overlap functions on finite chains constructed through pseudo-quasi overlap functions on different sub-chains are dissimilar. Eventually, we remove the continuity from pseudo-automorphisms and propose the concept of pseudo-quasi-automorphisms. Based on this, we utilize pseudo-overlap functions, pseudo-quasi-automorphisms, and integral functions to obtain discrete pseudo-quasi overlap functions on finite chains, moreover, we apply them to fuzzy multi-attribute group decision-making. The results indicate that compared to overlap functions and pseudo-overlap functions, discrete pseudo-quasi overlap functions on finite chains have stronger flexibility and a wider range of practical applications. Full article
(This article belongs to the Special Issue Fuzzy Sets and Fuzzy Systems)
Show Figures

Figure 1

Figure 1
<p>Framework diagram of the paper.</p>
Full article ">Figure 2
<p>A discrete pseudo-quasi overlap function <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>Q</mi> <msub> <mi>O</mi> <mi mathvariant="script">L</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>A discrete pseudo-quasi overlap function <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>Q</mi> <msub> <mi>O</mi> <mi>L</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>A discrete pseudo-quasi overlap function <math display="inline"><semantics> <msub> <mrow> <mi>P</mi> <mi>Q</mi> <mi>O</mi> </mrow> <msub> <mi>L</mi> <mi>N</mi> </msub> </msub> </semantics></math>.</p>
Full article ">Figure 5
<p>A discrete pseudo-quasi overlap function <math display="inline"><semantics> <msub> <mrow> <mi>P</mi> <mi>Q</mi> <mi>O</mi> </mrow> <msub> <mi>L</mi> <msup> <mi>N</mi> <mo>+</mo> </msup> </msub> </msub> </semantics></math>.</p>
Full article ">
13 pages, 3806 KiB  
Article
Stereodivergent Synthesis of Aldol Products Using Pseudo-C2 Symmetric N-benzyl-4-(trifluoromethyl)piperidine-2,6-dione
by Rina Yada, Tomoko Kawasaki-Takasuka and Takashi Yamazaki
Molecules 2024, 29(21), 5129; https://doi.org/10.3390/molecules29215129 - 30 Oct 2024
Viewed by 441
Abstract
The present article describes the successful performance of crossed aldol reactions of the CF3-containing pseudo-C2 symmetric cyclic imide with various aldehydes. The utilization of HMPA as an additive attained the preferential formation of the anti-products in good to excellent [...] Read more.
The present article describes the successful performance of crossed aldol reactions of the CF3-containing pseudo-C2 symmetric cyclic imide with various aldehydes. The utilization of HMPA as an additive attained the preferential formation of the anti-products in good to excellent yields, which contrasts with our previous method without this additive, proceeding to furnish the corresponding syn-isomers. The effective participation of ketones and α,β-unsaturated carbonyl compounds in reactions with this imide was also demonstrated to expand the application of this imide. Full article
(This article belongs to the Section Organic Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Possible transition state models of the present reactions (<b>M</b>: metal, <b>L</b>: ligand).</p>
Full article ">Figure 2
<p>Byproducts of the present reaction.</p>
Full article ">Scheme 1
<p>Crossed aldol reactions of <b>1</b> with aldehydes.</p>
Full article ">Scheme 2
<p>Confirmation of the stereochemistry.</p>
Full article ">Scheme 3
<p>Reaction of <b><span class="html-italic">anti</span>-2e</b> with LDA.</p>
Full article ">Scheme 4
<p>Reaction of <b>1</b> with cyclohex-2-en-1-one.</p>
Full article ">
15 pages, 8549 KiB  
Article
Advances in the Parameter Space Concept towards Picometer Precise Crystal Structure Refinement—A Resolution Study
by Matthias Zschornak, Christian Wagner, Melanie Nentwich, Muthu Vallinayagam and Karl F. Fischer
Crystals 2024, 14(8), 684; https://doi.org/10.3390/cryst14080684 - 26 Jul 2024
Viewed by 927
Abstract
The Parameter Space Concept (PSC) is an alternative approach to solving and refining (partial) crystal structures from very few pre-chosen X-ray or neutron diffraction amplitudes without the use of Fourier inversion. PSC interprets those amplitudes as piecewise analytic hyper-surfaces, so-called isosurfaces, in the [...] Read more.
The Parameter Space Concept (PSC) is an alternative approach to solving and refining (partial) crystal structures from very few pre-chosen X-ray or neutron diffraction amplitudes without the use of Fourier inversion. PSC interprets those amplitudes as piecewise analytic hyper-surfaces, so-called isosurfaces, in the Parameter Space, which is spanned by the spatial coordinates of all atoms of interest. The intersections of all isosurfaces constitute the (possibly degenerate) structure solution. The present feasibility study investigates the La and Sr split position of the potential high-temperature super-conductor (La0.5Sr1.5)MnO4, I4/mmm, with a postulated total displacement between La and Sr of a few pm by theoretical amplitudes of pre-selected 00l reflections (l=2,4,,20). The revision of 15-year-old results with state-of-the-art computing equipment enhances the former simplified model by varying the scattering power ratio fLa/fSr, as exploitable by means of resonant scattering contrast at synchrotron facilities, and irrevocably reveals one of the two originally proposed solutions as being a “blurred” pseudo-solution. Finally, studying the resolution limits of PSC as a function of intensity errors by means of Monte-Carlo simulations shows both that the split can only be resolved for sufficiently low errors and, particularly for the resonant scattering contrast, a theoretical precision down to ±0.19 pm can be achieved for this specific structural problem. Full article
(This article belongs to the Section Crystal Engineering)
Show Figures

Figure 1

Figure 1
<p>The 2-dimensional Parameter Space <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mo>⊗</mo> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </msub> </mrow> </semantics></math> for the La/Sr Partial Geometric Structure Factors <math display="inline"><semantics> <mrow> <msub> <mi>G</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mi>l</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> (Equation (<a href="#FD1-crystals-14-00684" class="html-disp-formula">1</a>)) of (La<sub>0.5</sub>Sr<sub>1.5</sub>)MnO<sub>4</sub>, <math display="inline"><semantics> <mrow> <mi>I</mi> <mn>4</mn> <mo>/</mo> </mrow> </semantics></math>mmm. The data are visualized in the full functional region (<b>left</b>) and the magnified region of interest (<b>right</b>) as isosurfaces, including positive (solid lines) and negative signs (dotted lines). The permutation symmetry of <math display="inline"><semantics> <msup> <mi mathvariant="script">P</mi> <mn>2</mn> </msup> </semantics></math> adds an exact pseudo-solution <math display="inline"><semantics> <mrow> <msubsup> <mi>z</mi> <mrow> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </mrow> <mo>*</mo> </msubsup> <mo>=</mo> <mn>0.35316</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mi>z</mi> <mrow> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </mrow> <mo>*</mo> </msubsup> <mo>=</mo> <mn>0.36316</mn> </mrow> </semantics></math> to the presumed split position <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mo>=</mo> <mn>0.36316</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mo>=</mo> <mn>0.35316</mn> </mrow> </semantics></math> (green dots). The gray shaded areas depict two different but equivalent asymmetric units, the conventional one (dark gray) and the one used in the literature (light gray), for better comparison. Gray solid lines depict mirrors and gray dashed lines anti-mirrors.</p>
Full article ">Figure 2
<p>The 2-dimensional Parameter Space <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mo>⊗</mo> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> </mrow> </semantics></math> for the complete structure of (La<sub>0.5</sub>Sr<sub>1.5</sub>)MnO<sub>4</sub>, <math display="inline"><semantics> <mrow> <mi>I</mi> <mn>4</mn> <mo>/</mo> </mrow> </semantics></math>mmm, with computed diffraction data <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>(</mo> <mi>l</mi> <mo>)</mo> </mrow> </semantics></math> (Equation (<a href="#FD2-crystals-14-00684" class="html-disp-formula">2</a>)) in the first quadrant (<b>left</b>) and the magnified region of interest (<b>right</b>). The functions are visualized as isosurfaces including positive (solid lines) and negative signs (dotted lines), with asymmetric units and mirrors in analogy to <a href="#crystals-14-00684-f001" class="html-fig">Figure 1</a>. The mirror along <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mo>=</mo> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> </mrow> </semantics></math> lifts via loss of permutation symmetry from <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mo>≠</mo> <msub> <mi>f</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> </mrow> </semantics></math> and the pseudo-solution <math display="inline"><semantics> <mrow> <mo>(</mo> <msubsup> <mi>z</mi> <mrow> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </mrow> <mo>*</mo> </msubsup> <mo>,</mo> <msubsup> <mi>z</mi> <mrow> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </mrow> <mo>*</mo> </msubsup> <mo>)</mo> </mrow> </semantics></math> with approximately interchanged coordinates shifts.</p>
Full article ">Figure 3
<p>(<b>a</b>) Model study of both the true and the pseudo-symmetric solutions as a function of scattering strength ratio by fitting intersecting isosurfaces <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>(</mo> <mi>l</mi> <mo>)</mo> </mrow> </semantics></math> (Equation (<a href="#FD2-crystals-14-00684" class="html-disp-formula">2</a>), least-squares). The ratio was varied by a scaling factor of 10, keeping the mean product <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mspace width="0.166667em"/> <mo>·</mo> <mspace width="0.166667em"/> <msub> <mi>f</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </msub> </mrow> </semantics></math> fixed. The isosurfaces <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>(</mo> <mi>l</mi> <mo>)</mo> </mrow> </semantics></math> were calculated for reflections <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mn>4</mn> <mo>,</mo> <mo>…</mo> <mo>,</mo> <mn>20</mn> </mrow> </semantics></math>. Confidence regions are given as error bars of <math display="inline"><semantics> <mrow> <mn>2.6</mn> <mi>σ</mi> </mrow> </semantics></math>, magnified by a factor of 10 for better visibility. The true solution <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mo>,</mo> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </msub> <mo>)</mo> </mrow> </semantics></math> is independent of the varying ratio (green dot), whereas the pseudo-solutions result in a linear series between the limits at a distance of <math display="inline"><semantics> <mrow> <mn>2</mn> <mspace width="0.166667em"/> <mo>·</mo> <mspace width="0.166667em"/> <mo>Δ</mo> <mi>z</mi> </mrow> </semantics></math> (red dots). As expected, the positional errors scale inversely to the scattering power. The change in scattering power directly reflects the distortion of the respective isosurface features, i.e., light weights act as elongations that increase the respective positional errors, shown for the limits of the series in (<b>b</b>) for “light La” (violet dot) and in (<b>c</b>) for “light Sr” (yellow dot).</p>
Full article ">Figure 4
<p>Monte-Carlo study of the split position with <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>z</mi> <mo>=</mo> <mn>0.0034</mn> </mrow> </semantics></math> for different Gaussian-distributed random errors of the reflection intensities for <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mn>4</mn> <mo>,</mo> <mo>…</mo> <mo>,</mo> <mn>20</mn> </mrow> </semantics></math> with 100 test samples each (Equation (<a href="#FD2-crystals-14-00684" class="html-disp-formula">2</a>), least-squares fits). The solution is in black and the pseudo-solution in red, each with two confidence regions, <math display="inline"><semantics> <mrow> <mn>1</mn> <mi>σ</mi> <mo>≈</mo> <mn>68</mn> <mo>%</mo> </mrow> </semantics></math> (heavy color shade) and <math display="inline"><semantics> <mrow> <mn>2.6</mn> <mi>σ</mi> <mo>≈</mo> <mn>99</mn> <mo>%</mo> </mrow> </semantics></math> (light color shade). Whereas an intensity error of <math display="inline"><semantics> <mrow> <mn>5</mn> <mo>%</mo> </mrow> </semantics></math> resolves the true solution within <math display="inline"><semantics> <mi>σ</mi> </semantics></math> confidence, highly precise intensity data with an intensity error better than <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math> are needed for <math display="inline"><semantics> <mrow> <mn>2.6</mn> <mi>σ</mi> </mrow> </semantics></math> confidence.</p>
Full article ">Figure 5
<p>Fitted positions <math display="inline"><semantics> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> </semantics></math> (blue +) and <math display="inline"><semantics> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> </mrow> </mtd> </mtr> </mtable> </msub> </semantics></math> (green ×) for a series of presumed splits <math display="inline"><semantics> <mrow> <mn>0</mn> <mo>≤</mo> <mo>Δ</mo> <mi>z</mi> <mo>≤</mo> <mn>0.02</mn> </mrow> </semantics></math> (between green and blue circles, accordingly) and three data qualities, given with a confidence level of 99% (error envelopes, 100 samples each, Equation (<a href="#FD2-crystals-14-00684" class="html-disp-formula">2</a>), least-squares fit). For the three different intensity errors <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>I</mi> <mo>/</mo> <mi>I</mi> </mrow> </semantics></math> = 20%, 5% and 1%, the distinct split positions can be resolved for splits <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>z</mi> <mo>&gt;</mo> <mn>0.02</mn> </mrow> </semantics></math> (≈25 pm), <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>z</mi> <mo>&gt;</mo> <mn>0.007</mn> </mrow> </semantics></math> (≈8.5 pm), and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>z</mi> <mo>&gt;</mo> <mn>0.003</mn> </mrow> </semantics></math> (≈3.7 pm), respectively.</p>
Full article ">Figure 6
<p>Two-dimensional Parameter Space <math display="inline"><semantics> <mrow> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>Sr</mi> <mi/> </mrow> </mtd> </mtr> </mtable> </msub> <mo>⊗</mo> <msub> <mi>z</mi> <mtable> <mtr> <mtd columnalign="right"> <mrow> <mi>La</mi> </mrow> </mtd> </mtr> </mtable> </msub> </mrow> </semantics></math> for the structure (La<sub>0.5</sub>Sr<sub>1.5</sub>)MnO<sub>4</sub>, <math display="inline"><semantics> <mrow> <mi>I</mi> <mn>4</mn> <mo>/</mo> </mrow> </semantics></math>mmm, in the vicinity of the split position (green dot). The diffraction data are given as intensity isosurfaces <math display="inline"><semantics> <mrow> <mi>I</mi> <mo>(</mo> <mi>l</mi> <mo>)</mo> </mrow> </semantics></math> (Equation (<a href="#FD2-crystals-14-00684" class="html-disp-formula">2</a>)) for two photon energies <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>15</mn> <mo> </mo> <mrow> <mi mathvariant="normal">k</mi> <mi mathvariant="normal">e</mi> <mspace width="-0.21251pt"/> <mi mathvariant="normal">V</mi> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>16.1</mn> <mo> </mo> <mrow> <mi mathvariant="normal">k</mi> <mi mathvariant="normal">e</mi> <mspace width="-0.21251pt"/> <mi mathvariant="normal">V</mi> </mrow> </mrow> </semantics></math> (just below the Sr absorption edge) and for error envelopes of <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>1</mn> </mrow> </semantics></math>% (blue and red lines). The contrast enhancement and respective superior resolution (small black region) originate from multiple large-angle intersections (upper right <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mn>4</mn> <mo>,</mo> <mo>…</mo> <mo>,</mo> <mn>10</mn> </mrow> </semantics></math>; lower right <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mn>4</mn> <mo>,</mo> <mo>…</mo> <mo>,</mo> <mn>20</mn> </mrow> </semantics></math>), and especially the small envelope of the low-intensity high-indexed reflection <math display="inline"><semantics> <mrow> <mn>0</mn> <mspace width="0.166667em"/> <mn>0</mn> <mspace width="0.166667em"/> <mn>16</mn> </mrow> </semantics></math> (increased opacity in lower inset). In addition, the second wavelength severely lifts the degeneracy of the pseudo-solution.</p>
Full article ">
16 pages, 329 KiB  
Article
The Effective Potential of Scalar Pseudo-Quantum Electrodynamics in (2 + 1)D
by Leandro O. Nascimento, Carlos A. P. C. Junior and José R. Santos
Condens. Matter 2024, 9(2), 25; https://doi.org/10.3390/condmat9020025 - 30 May 2024
Viewed by 1231
Abstract
The description of the electron–electron interactions in two-dimensional materials has a dimensional mismatch, where electrons live in (2 + 1)D while photons propagate in (3 + 1)D. In order to define an action in (2 + 1)D, one may perform a dimensional reduction [...] Read more.
The description of the electron–electron interactions in two-dimensional materials has a dimensional mismatch, where electrons live in (2 + 1)D while photons propagate in (3 + 1)D. In order to define an action in (2 + 1)D, one may perform a dimensional reduction of quantum electrodynamics in (3 + 1)D (QED4) into pseudo-quantum electrodynamics (PQED). The main difference between this model and QED4 is the presence of a pseudo-differential operator in the Maxwell term. However, besides the Coulomb repulsion, electrons in a material are subjected to several microscopic interactions, which are inherent in a many-body system. These are expected to reduce the range of the Coulomb potential, leading to a short-range interaction. Here, we consider the coupling to a scalar field in PQED for explaining such a mechanism, which resembles the spontaneous symmetry breaking (SSB) in Abelian gauge theories. In order to do so, we consider two cases: (i) by coupling the quantum electrodynamics to a Higgs field in (3 + 1)D and, thereafter, performing the dimensional reduction; and (ii) by coupling a Higgs field to the gauge field in PQED and, subsequently, calculating its effective potential. In case (i), we obtain a model describing electrons interacting through the Yukawa potential and, in case (ii), we show that SSB does not occur at one-loop approximation. The relevance of the model for describing electronic interactions in two-dimensional materials is also addressed. Full article
(This article belongs to the Special Issue PQED: 30 Years of Reduced Quantum Electrodynamics)
Show Figures

Figure 1

Figure 1
<p>The effective potential of the reduced-scalar model. We plot Equation (<a href="#FD21-condensedmatter-09-00025" class="html-disp-formula">21</a>) with <math display="inline"><semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mo>−</mo> <mn>0.5</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Λ</mi> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics></math>. Note that the local minimum in <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> is the only acceptable ground state, whether we assume that <math display="inline"><semantics> <mi>ρ</mi> </semantics></math> is always much less than <math display="inline"><semantics> <mi mathvariant="sans-serif">Λ</mi> </semantics></math>.</p>
Full article ">
29 pages, 9097 KiB  
Review
Pseudo-Spin Symmetry and the Hints for Unstable and Superheavy Nuclei
by Jing Geng, Zhiheng Wang, Jia Liu, Jiajie Li and Wenhui Long
Symmetry 2024, 16(5), 631; https://doi.org/10.3390/sym16050631 - 20 May 2024
Viewed by 1428
Abstract
The pseudo-spin symmetry (PSS) provides an important angle to understand nuclear microscopic structure and the novel phenomena found in unstable nuclei. The relativistic Hartree–Fock (RHF) theory, that takes the important degrees of freedom associated with the π-meson and ρ-tensor (ρ [...] Read more.
The pseudo-spin symmetry (PSS) provides an important angle to understand nuclear microscopic structure and the novel phenomena found in unstable nuclei. The relativistic Hartree–Fock (RHF) theory, that takes the important degrees of freedom associated with the π-meson and ρ-tensor (ρ-T) couplings into account, provides an appropriate description of the PSS restoration in realistic nuclei, particularly for the pseudo-spin (PS) doublets with high angular momenta (l˜). The investigations of the PSS within the RHF theory are recalled in this paper by focusing on the effects of the Fock terms. Aiming at common artificial shell closures appearing in previous relativistic mean-field calculations, the mechanism responsible for the PSS restoration of high-l˜ orbits is stressed, revealing the manifestation of nuclear in-medium effects on the PSS, and thus, providing qualitative guidance on modeling the in-medium balance between nuclear attractions and repulsions. Moreover, the essential role played by the ρ-T coupling, that contributes mainly via the Fock terms, is introduced as combined with the relations between the PSS and various nuclear phenomena, including the shell structure and the evolution, novel halo and bubble-like phenomena, and the superheavy magicity. As the consequences of the nuclear force in complicated nuclear many-body systems, the PSS itself and the mechanism therein can not only deepen our understanding of nuclear microscopic structure and relevant phenomena, but also provide special insight into the nature of the nuclear force, which can further enrich our knowledge of nuclear physics. Full article
(This article belongs to the Special Issue Restoration of Broken Symmetries in the Nuclear Many-Body Problem)
Show Figures

Figure 1

Figure 1
<p>Feynman diagrams of the Hartree and Fock terms under the meson-exchange picture of the nuclear force, where the solid lines with arrows represent the interacting nucleons, and the blue and red colors are used to identify the nucleons dressed by different quantum numbers.</p>
Full article ">Figure 2
<p>The PSO splitting <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>E</mi> <mi>PSO</mi> </msub> <mo>=</mo> <mrow> <mo stretchy="false">(</mo> <msub> <mi>E</mi> <mrow> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mi>j</mi> <mo>=</mo> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mo>−</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> <mo>−</mo> <msub> <mi>E</mi> <mrow> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mi>j</mi> <mo>=</mo> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mo>+</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> <mo stretchy="false">)</mo> </mrow> <mo>/</mo> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mo>+</mo> <mn>1</mn> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> versus the average s.p. energy <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>E</mi> <mo stretchy="false">¯</mo> </mover> <mi>PSO</mi> </msub> <mo>=</mo> <mrow> <mo stretchy="false">(</mo> <msub> <mi>E</mi> <mrow> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mi>j</mi> <mo>=</mo> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mo>−</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> <mo>+</mo> <msub> <mi>E</mi> <mrow> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mi>j</mi> <mo>=</mo> <mover accent="true"> <mi>l</mi> <mo stretchy="false">˜</mo> </mover> <mo>+</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> <mo stretchy="false">)</mo> </mrow> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> for neutron PS doublets <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>1</mn> <mover accent="true"> <mi>p</mi> <mo stretchy="false">˜</mo> </mover> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>1</mn> <mover accent="true"> <mi>d</mi> <mo stretchy="false">˜</mo> </mover> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>1</mn> <mover accent="true"> <mi>f</mi> <mo stretchy="false">˜</mo> </mover> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>2</mn> <mover accent="true"> <mi>p</mi> <mo stretchy="false">˜</mo> </mover> </mrow> </semantics></math> for <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Sn</mi> <none/> <none/> <mprescripts/> <none/> <mn>132</mn> </mmultiscripts> </semantics></math>. The SO splitting <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>E</mi> <mi>SO</mi> </msub> <mo>=</mo> <mrow> <mo stretchy="false">(</mo> <msub> <mi>E</mi> <mrow> <mi>l</mi> <mi>j</mi> <mo>=</mo> <mi>l</mi> <mo>−</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> <mo>−</mo> <msub> <mi>E</mi> <mrow> <mi>l</mi> <mi>j</mi> <mo>=</mo> <mi>l</mi> <mo>+</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> <mo stretchy="false">)</mo> </mrow> <mo>/</mo> <mrow> <mo stretchy="false">(</mo> <mn>2</mn> <mi>l</mi> <mo>+</mo> <mn>1</mn> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> is also given for <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>1</mn> <mi>p</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>1</mn> <mi>d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>1</mn> <mi>f</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>1</mn> <mi>g</mi> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>2</mn> <mi>p</mi> </mrow> </semantics></math> <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>2</mn> <mi>d</mi> </mrow> </semantics></math> pairs as a function of <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>E</mi> <mo stretchy="false">¯</mo> </mover> <mi>SO</mi> </msub> <mo>=</mo> <mrow> <mo stretchy="false">(</mo> <msub> <mi>E</mi> <mrow> <mi>l</mi> <mi>j</mi> <mo>=</mo> <mi>l</mi> <mo>−</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> <mo>+</mo> <msub> <mi>E</mi> <mrow> <mi>l</mi> <mi>j</mi> <mo>=</mo> <mi>l</mi> <mo>+</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> <mo stretchy="false">)</mo> </mrow> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math>. The results are obtained by the RHF with PKO1 (filled symbols) and the RMF with PKDD (open symbols), respectively. The figure is taken from Ref. [<a href="#B100-symmetry-16-00631" class="html-bibr">100</a>].</p>
Full article ">Figure 3
<p>The PCB and PSOP scaled by the factor <math display="inline"><semantics> <mrow> <msup> <mi>F</mi> <mn>2</mn> </msup> <mo>/</mo> <mrow> <mo stretchy="false">(</mo> <msup> <mi>V</mi> <mi>D</mi> </msup> <mo>−</mo> <mi>E</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> for the PS partner <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>2</mn> <mover accent="true"> <mi>p</mi> <mo stretchy="false">˜</mo> </mover> </mrow> </semantics></math> in <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Sn</mi> <none/> <none/> <mprescripts/> <none/> <mn>132</mn> </mmultiscripts> </semantics></math>. The PCB contributions are shown by the red lines, while the PSOP are shown by the blue shadows. The RHF with PKO1 is used for the calculations. The figure is taken from Ref. [<a href="#B100-symmetry-16-00631" class="html-bibr">100</a>].</p>
Full article ">Figure 4
<p>The functions <math display="inline"><semantics> <mrow> <mrow> <msubsup> <mi>V</mi> <mrow> <mi>PSO</mi> </mrow> <mi mathvariant="normal">E</mi> </msubsup> <msup> <mi>F</mi> <mn>2</mn> </msup> </mrow> <mo>/</mo> <mrow> <mo stretchy="false">(</mo> <msup> <mi>V</mi> <mi mathvariant="normal">D</mi> </msup> <mo>−</mo> <mi>E</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mi>V</mi> <mn>1</mn> <mi mathvariant="normal">E</mi> </msubsup> <mi>F</mi> <msup> <mi>F</mi> <mo>′</mo> </msup> <mo>/</mo> <mrow> <mo stretchy="false">(</mo> <msup> <mi>V</mi> <mi mathvariant="normal">D</mi> </msup> <mo>−</mo> <mi>E</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> given by the exchange (Fock) terms of the RHF with PKO1 for the PS partner <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>2</mn> <mover accent="true"> <mi>p</mi> <mo stretchy="false">˜</mo> </mover> </mrow> </semantics></math> of <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Sn</mi> <none/> <none/> <mprescripts/> <none/> <mn>132</mn> </mmultiscripts> </semantics></math>. The figure is taken from Ref. [<a href="#B100-symmetry-16-00631" class="html-bibr">100</a>].</p>
Full article ">Figure 5
<p>The radial wave functions <span class="html-italic">G</span> and <span class="html-italic">F</span> (<b>left panels</b>), and the non-local terms <span class="html-italic">X</span> and <span class="html-italic">Y</span> (<b>right panels</b>) given by the RHF with PKO1 for the PS partner <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>2</mn> <mover accent="true"> <mi>p</mi> <mo stretchy="false">˜</mo> </mover> </mrow> </semantics></math> in <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Sn</mi> <none/> <none/> <mprescripts/> <none/> <mn>132</mn> </mmultiscripts> </semantics></math>. The figure is taken from Ref. [<a href="#B100-symmetry-16-00631" class="html-bibr">100</a>].</p>
Full article ">Figure 6
<p>Proton SPE of <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Pb</mi> <none/> <none/> <mprescripts/> <none/> <mn>208</mn> </mmultiscripts> </semantics></math>. The results are calculated by RHF with PKA1 and PKO3, and RMF with DD-ME2. The experimental data are from Ref. [<a href="#B147-symmetry-16-00631" class="html-bibr">147</a>].</p>
Full article ">Figure 7
<p>Proton SPE for <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Gd</mi> <none/> <none/> <mprescripts/> <none/> <mn>146</mn> </mmultiscripts> </semantics></math>, calculated by RHF with PKA1 [<a href="#B48-symmetry-16-00631" class="html-bibr">48</a>], Gogny HFB with D1S [<a href="#B148-symmetry-16-00631" class="html-bibr">148</a>], Skyrme HF with SLy4 [<a href="#B149-symmetry-16-00631" class="html-bibr">149</a>], and RMF with PKDD [<a href="#B115-symmetry-16-00631" class="html-bibr">115</a>]. The figure is taken from Ref. [<a href="#B49-symmetry-16-00631" class="html-bibr">49</a>]. Reprinted with permission from Ref. [<a href="#B49-symmetry-16-00631" class="html-bibr">49</a>]. Copyright 2009 Elsevier.</p>
Full article ">Figure 8
<p>(<b>a</b>) SO splittings <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>E</mi> <mrow> <mn>2</mn> <mi>d</mi> </mrow> </msub> </mrow> </semantics></math> of proton <math display="inline"><semantics> <mrow> <mi>π</mi> <mn>2</mn> <mi>d</mi> </mrow> </semantics></math> partner extracted from the calculations of RHF with PKA1, Gogny HFB with D1S, Skyrme HF with SLy4, and RMF with PKDD, in comparison with experimental data [<a href="#B47-symmetry-16-00631" class="html-bibr">47</a>]. (<b>b</b>) Detailed contributions of <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>E</mi> <mrow> <mn>2</mn> <mi>d</mi> </mrow> </msub> </mrow> </semantics></math> from different channels given by the RHF calculations with PKA1. The figure is taken from Ref. [<a href="#B49-symmetry-16-00631" class="html-bibr">49</a>]. Reprinted with permission from Ref. [<a href="#B49-symmetry-16-00631" class="html-bibr">49</a>]. Copyright 2009 Elsevier.</p>
Full article ">Figure 9
<p>Proton shell gaps (MeV) (<b>a</b>) and the splittings of neighboring PS partners <math display="inline"><semantics> <msubsup> <mi>E</mi> <mrow> <mi>PSO</mi> </mrow> <mi>π</mi> </msubsup> </semantics></math> (MeV) (<b>b</b>) for the traditional magic nuclei <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Ca</mi> <none/> <none/> <mprescripts/> <none/> <mn>48</mn> </mmultiscripts> </semantics></math>, <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Zr</mi> <none/> <none/> <mprescripts/> <none/> <mn>90</mn> </mmultiscripts> </semantics></math>, <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Sn</mi> <none/> <none/> <mprescripts/> <none/> <mn>132</mn> </mmultiscripts> </semantics></math>, and <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Pb</mi> <none/> <none/> <mprescripts/> <none/> <mn>208</mn> </mmultiscripts> </semantics></math>, and the superheavy one <math display="inline"><semantics> <mmultiscripts> <mn>126</mn> <none/> <none/> <mprescripts/> <none/> <mn>310</mn> </mmultiscripts> </semantics></math>. The results are calculated by PKA1, PKO3, DD-ME2, PK1 [<a href="#B115-symmetry-16-00631" class="html-bibr">115</a>], and NL3* [<a href="#B150-symmetry-16-00631" class="html-bibr">150</a>]. The shadow areas denote the spreading of the PSO splittings given by PKA1 and the other selected models. Experimental data are taken from Ref. [<a href="#B147-symmetry-16-00631" class="html-bibr">147</a>]. The figure is taken from Ref. [<a href="#B104-symmetry-16-00631" class="html-bibr">104</a>]. Reprinted with permission from Ref. [<a href="#B104-symmetry-16-00631" class="html-bibr">104</a>]. Copyright 2019 American Physical Society.</p>
Full article ">Figure 10
<p>Proton (<math display="inline"><semantics> <mi>π</mi> </semantics></math>) PSO splittings <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msubsup> <mi>E</mi> <mrow> <mi>PSO</mi> </mrow> <mi>π</mi> </msubsup> </mrow> </semantics></math> (MeV) (<b>a</b>) in <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Pb</mi> <none/> <none/> <mprescripts/> <none/> <mn>208</mn> </mmultiscripts> </semantics></math> as functions of pseudo-orbit <math display="inline"><semantics> <msup> <mi>l</mi> <mo>′</mo> </msup> </semantics></math>, and the sum contributions from kinetic energy, <math display="inline"><semantics> <mi>σ</mi> </semantics></math>, and <math display="inline"><semantics> <mi>ω</mi> </semantics></math> potential energies <math display="inline"><semantics> <msub> <mi>E</mi> <mrow> <mrow> <mi>kin</mi> <mo>.</mo> </mrow> <mo>+</mo> <mi>σ</mi> <mo>+</mo> <mi>ω</mi> </mrow> </msub> </semantics></math> (<b>b</b>). The results are extracted from the calculations with PKA1, PKO3, and DD-ME2. Meson–nucleon coupling constants, namely, (<b>c</b>) the isoscalar <math display="inline"><semantics> <msub> <mi>g</mi> <mi>σ</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>g</mi> <mi>ω</mi> </msub> </semantics></math>, (<b>d</b>) the isovector <math display="inline"><semantics> <msub> <mi>g</mi> <mi>ρ</mi> </msub> </semantics></math>, and (<b>e</b>) <math display="inline"><semantics> <msub> <mi>κ</mi> <mi>ρ</mi> </msub> </semantics></math> [<math display="inline"><semantics> <mrow> <msub> <mi>κ</mi> <mi>ρ</mi> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>f</mi> <mi>ρ</mi> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>/</mo> <msub> <mi>g</mi> <mi>ρ</mi> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>] and <math display="inline"><semantics> <msub> <mi>f</mi> <mi>π</mi> </msub> </semantics></math>, as functions of the density <math display="inline"><semantics> <msub> <mi>ρ</mi> <mi>b</mi> </msub> </semantics></math> (<math display="inline"><semantics> <msup> <mi>fm</mi> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>) for PKA1, PKO3, DD-ME2, and DD-LZ1. The figure is taken from Refs. [<a href="#B104-symmetry-16-00631" class="html-bibr">104</a>,<a href="#B143-symmetry-16-00631" class="html-bibr">143</a>]. Reprinted with permission from Ref. [<a href="#B104-symmetry-16-00631" class="html-bibr">104</a>]. Copyright 2019 American Physical Society. Reprinted with permission from Ref. [<a href="#B143-symmetry-16-00631" class="html-bibr">143</a>]. Copyright 2020 Chinese Physical Society and the Institute of High Energy Physics of the Chinese Academy of Sciences and the Institute of Modern Physics of the Chinese Academy of Sciences and IOP Publishing Ltd.</p>
Full article ">Figure 11
<p>Schematic diagrams of the matter density <math display="inline"><semantics> <msub> <mi>ρ</mi> <mi>b</mi> </msub> </semantics></math> and probability densities of the <span class="html-italic">s</span>-, <span class="html-italic">p</span>-, <span class="html-italic">d</span>-, and <span class="html-italic">f</span>-orbits.</p>
Full article ">Figure 12
<p>Panel (<b>a</b>) shows the correlation between the critical parameters <math display="inline"><semantics> <msub> <mi>T</mi> <mi>C</mi> </msub> </semantics></math> (MeV) and <math display="inline"><semantics> <msub> <mi>P</mi> <mi>C</mi> </msub> </semantics></math> (MeV <math display="inline"><semantics> <msup> <mi>fm</mi> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>). The solid circles correspond to 20 selected relativistic Lagrangians from Ref. [<a href="#B165-symmetry-16-00631" class="html-bibr">165</a>] and PKA1, and the open circles correspond to the testing sets <math display="inline"><semantics> <mrow> <mi>x</mi> <msub> <mi>κ</mi> <mi>ρ</mi> </msub> </mrow> </semantics></math>, with <span class="html-italic">x</span> = 1.0 to 0.7 (in red) and <math display="inline"><semantics> <mrow> <mi>x</mi> <msubsup> <mi>κ</mi> <mi>ρ</mi> <mo>*</mo> </msubsup> </mrow> </semantics></math> with <span class="html-italic">x</span> = 0.9, 0.8, and 0.7 (in blue). As the references, the solid and dashed lines represent the linear fittings with the Pearson correlation coefficient <span class="html-italic">r</span>. Panel (<b>b</b>) presents LG phase diagrams of thermal nuclear matter at temperature <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> MeV given by the sets <math display="inline"><semantics> <mrow> <mi>x</mi> <msub> <mi>κ</mi> <mi>ρ</mi> </msub> </mrow> </semantics></math> (<span class="html-italic">x</span> = 1.0, 0.9, 0.8, and 0.7), as compared to the ones given by RHF Lagrangians PKA1 and PKO3. The figure is taken from Ref. [<a href="#B166-symmetry-16-00631" class="html-bibr">166</a>]. Reprinted with permission from Ref. [<a href="#B166-symmetry-16-00631" class="html-bibr">166</a>]. Copyright 2021 American Physical Society.</p>
Full article ">Figure 13
<p>Two-proton shell gap <math display="inline"><semantics> <msub> <mi>δ</mi> <mrow> <mn>2</mn> <mi>p</mi> </mrow> </msub> </semantics></math> for <span class="html-italic">N</span> = 82 and 126 isotonic chains obtained from RH(F)B calculations with the new effective interaction DD-LZ1 (<b>left panel</b>) and PCF-PK1 (<b>right panel</b>). For comparison, the experimental data [<a href="#B168-symmetry-16-00631" class="html-bibr">168</a>] and calculated results of PKA1, PKO1, DD-ME2, PC-PK1 [<a href="#B169-symmetry-16-00631" class="html-bibr">169</a>], DD-PC1 [<a href="#B170-symmetry-16-00631" class="html-bibr">170</a>], and DD-ME<math display="inline"><semantics> <mi>δ</mi> </semantics></math> [<a href="#B123-symmetry-16-00631" class="html-bibr">123</a>] are also given. The results of DD-LZ1 are taken from Ref. [<a href="#B143-symmetry-16-00631" class="html-bibr">143</a>], and those of PCF-PK1 are taken from Ref. [<a href="#B167-symmetry-16-00631" class="html-bibr">167</a>]. Reprinted with permission from Ref. [<a href="#B143-symmetry-16-00631" class="html-bibr">143</a>]. Copyright 2020 Chinese Physical Society and the Institute of High Energy Physics of the Chinese Academy of Sciences and the Institute of Modern Physics of the Chinese Academy of Sciences and IOP Publishing Ltd. Reprinted with permission from Ref. [<a href="#B167-symmetry-16-00631" class="html-bibr">167</a>]. Copyright 2022 American Physical Society.</p>
Full article ">Figure 14
<p>(<b>a</b>) Two-body interaction matrix elements <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>a</mi> <mi>b</mi> </mrow> </msub> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mi>π</mi> <mn>2</mn> <msub> <mi>d</mi> <mrow> <mn>5</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math> (filled symbols) and <math display="inline"><semantics> <mrow> <mi>π</mi> <mn>1</mn> <msub> <mi>g</mi> <mrow> <mn>7</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math> (open symbols), and <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mi>ν</mi> <mn>2</mn> <msub> <mi>g</mi> <mrow> <mn>9</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>ν</mi> <mn>1</mn> <msub> <mi>i</mi> <mrow> <mn>13</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>. The figure is taken from Ref. [<a href="#B109-symmetry-16-00631" class="html-bibr">109</a>]. Reprinted with permission from Ref. [<a href="#B109-symmetry-16-00631" class="html-bibr">109</a>]. Copyright 2010 American Physical Society. (<b>b</b>) Schematic diagrams of the proton–neutron configurations for the existence (I) and elimination (II) of <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <mn>58</mn> </mrow> </semantics></math> artificial shell.</p>
Full article ">Figure 15
<p>(<b>a</b>) Neutron and proton densities; (<b>b</b>) neutron canonical SPE, occupation probability (<span class="html-italic">x</span>-axis error bars), and Fermi energy <math display="inline"><semantics> <msub> <mi>E</mi> <mi>F</mi> </msub> </semantics></math> (open circles). Results are calculated by RHFB with PKA1 plus the Gogny pairing force D1S. The figure is taken from Ref. [<a href="#B109-symmetry-16-00631" class="html-bibr">109</a>]. Reprinted with permission from Ref. [<a href="#B109-symmetry-16-00631" class="html-bibr">109</a>]. Copyright 2010 American Physical Society.</p>
Full article ">Figure 16
<p>Energy difference <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msubsup> <mi>E</mi> <mrow> <mi>s</mi> <mi>d</mi> </mrow> <mi>π</mi> </msubsup> <mo>=</mo> <msub> <mi>ε</mi> <mrow> <mi>π</mi> <mn>2</mn> <msub> <mi>s</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> </mrow> </msub> <mo>−</mo> <msub> <mi>ε</mi> <mrow> <mi>π</mi> <mn>1</mn> <msub> <mi>d</mi> <mrow> <mn>3</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> </mrow> </msub> </mrow> </semantics></math> of Ca isotopes. The experimental values of <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Ca</mi> <none/> <none/> <mprescripts/> <none/> <mn>40</mn> </mmultiscripts> </semantics></math> and <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">Ca</mi> <none/> <none/> <mprescripts/> <none/> <mn>48</mn> </mmultiscripts> </semantics></math> are taken from Ref. [<a href="#B147-symmetry-16-00631" class="html-bibr">147</a>] (up triangle), Ref. [<a href="#B174-symmetry-16-00631" class="html-bibr">174</a>] (down triangle), Ref. [<a href="#B175-symmetry-16-00631" class="html-bibr">175</a>] (left triangle), and Ref. [<a href="#B176-symmetry-16-00631" class="html-bibr">176</a>] (right triangle). The figure is taken from Ref. [<a href="#B110-symmetry-16-00631" class="html-bibr">110</a>].</p>
Full article ">Figure 17
<p>Canonical proton SPE along the <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>28</mn> </mrow> </semantics></math> isotonic chain calculated by RHFB with PKA1 (<b>a</b>) and RHB with DD-ME2 (<b>b</b>). The lengths of thick bars correspond with the occupation probabilities of the proton orbits and the filled stars denote the experimental data. Charge distributions of <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>28</mn> </mrow> </semantics></math> isotones calculated by RHFB with PKA1 and PKO3, and by RHB with DD-ME2 (<b>c</b>). The figure is taken from Ref. [<a href="#B31-symmetry-16-00631" class="html-bibr">31</a>]. Reprinted with permission from Ref. [<a href="#B31-symmetry-16-00631" class="html-bibr">31</a>]. Copyright 2016 American Physical Society.</p>
Full article ">Figure 18
<p>Proton (<b>left panel</b>) and neutron (<b>right panel</b>) SPE of SHN <math display="inline"><semantics> <mmultiscripts> <mn>120</mn> <mn>184</mn> <none/> <mprescripts/> <none/> <mn>304</mn> </mmultiscripts> </semantics></math>. The results are extracted from RHFB calculations with PKO<span class="html-italic">i</span> (<math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> </mrow> </semantics></math>) and PKA1, and compared to the RHB ones with PKDD and DD-ME2. In all cases, the pairing force is derived from the finite-range Gogny force D1S with the strength factor <span class="html-italic">f</span> = 0.9. The figure is taken from Ref. [<a href="#B51-symmetry-16-00631" class="html-bibr">51</a>].</p>
Full article ">Figure 19
<p>The neutron (<math display="inline"><semantics> <mi>ν</mi> </semantics></math>) and proton (<math display="inline"><semantics> <mi>π</mi> </semantics></math>) density distributions of superheavy nuclei calculated by the RHFB model with PKA1 (red line) and PKO3 (blue line).</p>
Full article ">
18 pages, 436 KiB  
Article
Landau Levels versus Hydrogen Atom
by Tekin Dereli, Philippe Nounahon and Todor Popov
Universe 2024, 10(4), 172; https://doi.org/10.3390/universe10040172 - 7 Apr 2024
Cited by 1 | Viewed by 2000
Abstract
The Landau problem and harmonic oscillator in the plane share a Hilbert space that carries the structure of Dirac’s remarkable so(2,3) representation. We show that the orthosymplectic algebra osp(1|4) is [...] Read more.
The Landau problem and harmonic oscillator in the plane share a Hilbert space that carries the structure of Dirac’s remarkable so(2,3) representation. We show that the orthosymplectic algebra osp(1|4) is the spectrum generating algebra for the Landau problem and, hence, for the 2D isotropic harmonic oscillator. The 2D harmonic oscillator is in duality with the 2D quantum Coulomb–Kepler systems, with the osp(1|4) symmetry broken down to the conformal symmetry so(2,3). The even so(2,3) submodule (coined Rac) generated from the ground state of zero angular momentum is identified with the Hilbert space of a 2D hydrogen atom. An odd element of the superalgebra osp(1|4) creates a pseudo-vacuum with intrinsic angular momentum 1/2 from the vacuum. The odd so(2,3)-submodule (coined Di) built upon the pseudo-vacuum is the Hilbert space of a magnetized 2D hydrogen atom: a quantum system of a dyon and an electron. Thus, the Hilbert space of the Landau problem is a direct sum of two massless unitary so(2,3) representations, namely, the Di and Rac singletons introduced by Flato and Fronsdal. Full article
Show Figures

Figure 1

Figure 1
<p>Hilbert space <math display="inline"><semantics> <mrow> <mi mathvariant="bold">Dirac</mi> <mo>=</mo> <mi mathvariant="bold">Di</mi> <mo>⊕</mo> <mi mathvariant="bold">Rac</mi> </mrow> </semantics></math> of Landau levels: odd <math display="inline"><semantics> <mi mathvariant="bold">Di</mi> </semantics></math> (red) and even <math display="inline"><semantics> <mi mathvariant="bold">Rac</mi> </semantics></math> (blue) states.</p>
Full article ">
14 pages, 3587 KiB  
Article
Pseudo-Polymorphism in 2-Pyridylmethoxy Cone Derivatives of p-tert-butylcalix[4]arene and p-tert-butylhomooxacalix[n]arenes
by Siddharth Joshi, Neal Hickey, Paula M. Marcos and Silvano Geremia
Crystals 2024, 14(4), 343; https://doi.org/10.3390/cryst14040343 - 3 Apr 2024
Viewed by 1240
Abstract
This paper investigates pseudo-polymorphism in 2-pyridylmethoxy derivatives of p-tert-butylcalix[4]arene (PyC4), p-tert-butyldihomooxa-calix[4]arenes (PyHOC4), and p-tert-butylhexahomotrioxacalix[3]arenes (PyHO3C3), presenting 11 crystal structures with 15 crystallographically independent molecules. The macrocycle of PyC4 is smaller and less flexible with [...] Read more.
This paper investigates pseudo-polymorphism in 2-pyridylmethoxy derivatives of p-tert-butylcalix[4]arene (PyC4), p-tert-butyldihomooxa-calix[4]arenes (PyHOC4), and p-tert-butylhexahomotrioxacalix[3]arenes (PyHO3C3), presenting 11 crystal structures with 15 crystallographically independent molecules. The macrocycle of PyC4 is smaller and less flexible with respect to those of PyHOC4 and PyHO3C3, and in solution, the cone conformation of these three molecules exhibits different point symmetries: C4, Cs, and C3, respectively. A correlation is observed between the macrocycle’s structural rigidity and the number of pseudo-polymorphs formed. The more rigid PyC4 displays a higher number (six) of pseudo-polymorphs compared to PyHOC4 and PyHO3C3, which exhibit a smaller number of crystalline forms (three and two, respectively). The X-ray structures obtained show that the conformation of the macrorings is primarily influenced by the presence of an acetonitrile guest molecule within the cavity, with limited impact from crystal packing and intermolecular co-crystallized solvent molecules. Notably, both calix[4]arene derivatives produce a host–guest complex with acetonitrile, while the most flexible and less aromatic PyHO3C3 does not give crystals with acetonitrile as the guest. Intertwined 1D and 2D solvent channel networks were observed in the PyHOC4-hexane and in the PyHO3C3-H2O-MeOH crystal structures, respectively, while the other pseudopolymorphs of PyHOC4 and PyHO3C3 and all PyC4 crystal forms exhibit closely packed crystal structures without open channels. Full article
(This article belongs to the Section Biomolecular Crystals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of the three macrocycles.</p>
Full article ">Figure 2
<p>Six asymmetric units of various <b>PyC4</b> structures obtained in different solvents: (<b>a</b>) <b>PyC4-MeOH-α</b>; (<b>b</b>) <b>PyC4-H<sub>2</sub>O-α</b>; (<b>c</b>) <b>PyC4-MeOH-β</b>; (<b>d</b>) <b>PyC4⸦MeCN-MeOH</b>; (<b>e</b>) <b>PyC4⸦MeCN-H<sub>2</sub>O;</b> and (<b>f</b>) <b>PyC4</b>. H-bonds between calixarene and co-crystallized solvent molecules are represented by green dashed lines. Hydrogen atoms have been omitted and only one position of the disorder groups is shown for better clarity. Atoms are in CPK colors.</p>
Full article ">Figure 3
<p>Asymmetric units of three different <b>PyHOC4</b> structures obtained in different solvents: (<b>a</b>) <b>PyHOC4-Hexane</b>; (<b>b</b>) <b>PyHOC4-DMSO;</b> and (<b>c</b>) <b>PyHOC4⸦MeCN-MeOH</b>. Hydrogen atoms have been omitted and only one position of the disorder groups is shown for better clarity. Atoms are in CPK colors.</p>
Full article ">Figure 4
<p>Crystal packing of <b>PyHOC4-hexane</b> (<b>a</b>) viewed along the crystallographic <span class="html-italic">b</span> axis and (<b>b</b>) viewed along the crystallographic <span class="html-italic">a</span> axis. The surface of the solvent accessible volume calculated with a 1.2 Å probe evidences the intertwined 1D channel network developed along the crystallographic <span class="html-italic">a</span> axis. Atoms in CPK colors.</p>
Full article ">Figure 5
<p>Asymmetric unit of two different <b>PyHO3C3</b> structures: (<b>a</b>) the anhydrous form <b>PyHO3C3</b> and (<b>b</b>) the hydrate form <b>PyHO3C3-H<sub>2</sub>O-MeOH</b>. H-bonds between calixarene and co-crystallized solvent molecules are represented by green dashed lines. Hydrogen atoms have been omitted and only one position of the disorder groups is shown for better clarity. Atoms are in CPK colors.</p>
Full article ">Figure 6
<p>Crystal packing of <b>PyHO3C3-H<sub>2</sub>O-MeOH</b> (<b>a</b>) viewed along the crystallographic <span class="html-italic">a</span> axis and (<b>b</b>) viewed along the crystallographic <span class="html-italic">a</span> axis. The surface of the solvent accessible volume calculated with a 1.2 Å probe evidences the 2D channel network developed in the (100) crystallographic plane. Atoms in CPK colors.</p>
Full article ">Figure 7
<p>Effect of acetonitrile guest complexation on the macrocycle conformation. The cups are viewed perpendicular to the mean planes of the bridging methylene carbon atoms: (<b>a</b>) <b>PyC4</b>, (<b>b</b>) <b>PyC4⸦MeCN</b>, (<b>c</b>) <b>PyHOC4</b>, (<b>d</b>) <b>PyHOC4⸦MeCN</b>, (<b>e</b>) <b>PyHO3C3</b>. Atoms in CPK colors.</p>
Full article ">
16 pages, 1048 KiB  
Article
On the Breaking of the U(1) Peccei–Quinn Symmetry and Its Implications for Neutrino and Dark Matter Physics
by Osvaldo Civitarese
Symmetry 2024, 16(3), 364; https://doi.org/10.3390/sym16030364 - 18 Mar 2024
Viewed by 1105
Abstract
The Standard Model of electroweak interactions is based on the fundamental SU(2)weak × U(1)elect representation. It assumes massless neutrinos and purely left-handed massive W± and Z0 bosons to which one should add the massless photon. The existence, [...] Read more.
The Standard Model of electroweak interactions is based on the fundamental SU(2)weak × U(1)elect representation. It assumes massless neutrinos and purely left-handed massive W± and Z0 bosons to which one should add the massless photon. The existence, verified experimentally, of neutrino oscillations poses a challenge to this scheme, since the oscillations take place between at least three massive neutrinos belonging to a mass hierarchy still to be determined. One should also take into account the possible existence of sterile neutrino species. In a somehow different context, the fundamental nature of the strong interaction component of the forces in nature is described by the, until now, extremely successful representation based on the SU(3)strong group which, together with the confining rule, give a description of massive hadrons in terms of quarks and gluons. To this is added the minimal U(1) Higgs group to give mass to the otherwise massless generators. This representation may also be challenged by the existence of both dark matter and dark energy, of still unknown composition. In this note, we shall discuss a possible connection between these questions, namely the need to extend the SU(3)strong × SU(2)weak × U(1)elect to account for massive neutrinos and dark matter. The main point of it is related to the role of axions, as postulated by Roberto Peccei and Helen Quinn. The existence of neutral pseudo-scalar bosons, that is, the axions, has been proposed long ago by Peccei and Quinn to explain the suppression of the electric dipole moment of the neutron. The associated U(1)PQ symmetry breaks at very high energy, and it guarantees that the interaction of other particles with axions is very weak. We shall review the axion properties in connection with the apparently different contexts of neutrino and dark matter physics. Full article
(This article belongs to the Special Issue Role of Symmetries in Nuclear Physics)
Show Figures

Figure 1

Figure 1
<p>Interaction between neutrinos and particles belonging to the environment.</p>
Full article ">Figure 2
<p>Real (upper plot) and imaginary (lower plot) matrix elements of the neutrino density matrix as a function of time for the three-flavor scheme. We used the following parameters: <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>λ</mi> <mrow> <mi>c</mi> <mi>o</mi> <mi>u</mi> <mi>p</mi> </mrow> </msub> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics></math> for the Gaussian function and the coupling to the environment. The relative time scale (<span class="html-italic">t</span> divided by <span class="html-italic">ℏ</span> ) is expressed in units of 10<sup>7</sup>.</p>
Full article ">Figure 3
<p>Diagrams showing the double beta decay transitions, for the case of zero neutrino double beta decay channels, and an equivalent process mediated by W bosons. The neutrinoless double beta decay process is not allowed by the Standard Model, since it implies that the neutrino is a Majorana particle.</p>
Full article ">Figure 4
<p>View of the close connection existing between experimental, nuclear, and particle physics focusing on the determination of neutrino properties, as well as on the non-trivial extensions of the minimal Standard Model.</p>
Full article ">
13 pages, 4944 KiB  
Article
Synthesis and Dynamic Behavior of Ce(IV) Double-Decker Complexes of Sterically Hindered Phthalocyanines
by Jeevithra Dewi Subramaniam, Toshio Nishino, Kazuma Yasuhara and Gwénaël Rapenne
Molecules 2024, 29(4), 888; https://doi.org/10.3390/molecules29040888 - 17 Feb 2024
Cited by 2 | Viewed by 1333
Abstract
Phthalocyanines and their double-decker complexes are interesting in designing rotative molecular machines, which are crucial for the development of molecular motors and gears. This study explores the design and synthesis of three bulky phthalocyanine ligands functionalized at the α-positions with phenothiazine or carbazole [...] Read more.
Phthalocyanines and their double-decker complexes are interesting in designing rotative molecular machines, which are crucial for the development of molecular motors and gears. This study explores the design and synthesis of three bulky phthalocyanine ligands functionalized at the α-positions with phenothiazine or carbazole fragments, aiming to investigate dynamic rotational motions in these sterically hindered molecular complexes. Homoleptic and heteroleptic double-decker complexes were synthesized through the complexation of these ligands with Ce(IV). Notably, CeIV(Pc2)2 and CeIV(Pc3)2, both homoleptic complexes, exhibited blocked rotational motions even at high temperatures. The heteroleptic CeIV(Pc)(Pc3) complex, designed to lower symmetry, demonstrated switchable rotation along the pseudo-C4 symmetry axis upon heating the solution. Variable-temperature 1H-NMR studies revealed distinct dynamic behaviors in these complexes. This study provides insights into the rotational dynamics of sterically hindered double-decker complexes, paving the way for their use in the field of rotative molecular machines. Full article
(This article belongs to the Special Issue Macrocyclic Compounds: Derivatives and Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic design of the sterically hindered double-decker complex of this work with the axis of rotation of interest (the main axis of rotation is given as a dashed line). The phthalocyanine rings are shown in green, the lanthanoid ion in purple, and the planar bulky substituents in blue.</p>
Full article ">Figure 2
<p>Regioisomers of the Pc ring obtained by cyclic tetramerization of 3-substituted phthalonitrile.</p>
Full article ">Figure 3
<p>The three Pc ligands synthesized by tetramerization of phathalonitrile precursors.</p>
Full article ">Figure 4
<p>The three double-decker complexes synthesized by coordination with a cerium(IV) ion.</p>
Full article ">Figure 5
<p>Possible stereoisomers of double-decker complex formed from a Pc functionalized at the four <math display="inline"><semantics> <mrow> <mi>α</mi> </mrow> </semantics></math>-positions: R,R and S,S enantiomers as well as the R,S meso diastereomer (Cbz = 3,6-di-<span class="html-italic">tert</span>-butyl-carbazole).</p>
Full article ">Figure 6
<p>Side view (<b>a</b>) and top view (<b>b</b>) of the single crystal structure of <b>Ce<sup>IV</sup>(Pc2)<sub>2</sub></b>. Hydrogens are omitted for clarity. Ce is in orange, N in blue and C in grey.</p>
Full article ">Figure 7
<p><sup>1</sup>H-NMR spectra of (<b>a</b>) <b>H<sub>2</sub>Pc2</b> and (<b>b</b>) <b>Ce<sup>IV</sup>(Pc2)<sub>2</sub></b> in CDCl<sub>3</sub> (400 MHz). The full assignments of the signals (indicated with letters in the molecule shown top right) were made with the assistance of COSY (<a href="#app1-molecules-29-00888" class="html-app">Figure S9</a>).</p>
Full article ">Figure 8
<p>VT-<sup>1</sup>H-NMR spectra of <b>Ce<sup>IV</sup>(Pc)(Pc3)</b> in C<sub>2</sub>D<sub>2</sub>Cl<sub>4</sub> (600 MHz): (<b>a</b>) In the aromatic region, the signals corresponding to the α and β protons (green region) are shifted and simplified at higher temperatures. (<b>b</b>) The molecular structure with the color code of the discussed protons; in red and blue are the <span class="html-italic">in</span> and <span class="html-italic">out tert</span>-butyl protons, and in green is the <b>Pc</b> protons. (<b>c</b>) In the aliphatic region, the signals corresponding to the <span class="html-italic">tert</span>-butyl protons are splitted in two groups. While the in signals are not changed, The blue and green arrows correspond to the rotation axis around the pseudo-C<sub>4</sub> symmetry axis (blue-dotted axis) and the C-N bound axis between the phthalocyanine and carbazole (green-dotted axis).</p>
Full article ">Scheme 1
<p>Synthesis of the disymmetric <b>H<sub>2</sub>Pc3</b> ligand based on a statistical condensation reaction.</p>
Full article ">
23 pages, 1002 KiB  
Article
de Broglie, General Covariance and a Geometric Background to Quantum Mechanics
by Basil Hiley and Glen Dennis
Symmetry 2024, 16(1), 67; https://doi.org/10.3390/sym16010067 - 4 Jan 2024
Cited by 1 | Viewed by 2308
Abstract
What is striking about de Broglie’s foundational work on wave–particle dualism is the role played by pseudo-Riemannian geometry in his early thinking. While exploring a fully covariant description of the Klein–Gordon equation, he was led to the revolutionary idea that a variable rest [...] Read more.
What is striking about de Broglie’s foundational work on wave–particle dualism is the role played by pseudo-Riemannian geometry in his early thinking. While exploring a fully covariant description of the Klein–Gordon equation, he was led to the revolutionary idea that a variable rest mass was essential. DeWitt later explained that in order to obtain a covariant quantum Hamiltonian, one must supplement the classical Hamiltonian with an additional energy 2Q from which the quantum potential emerges, a potential that Berry has recently shown also arises in classical wave optics. In this paper, we show how these ideas emerge from an essentially geometric structure in which the information normally carried by the wave function is contained within the algebraic description of the geometry itself, within an element of a minimal left ideal. We establish the fundamental importance of conformal symmetry, in which rescaling of the rest mass plays a vital role. Thus, we have the basis for a radically new theory of quantum phenomena based on the process of mass-energy flow. Full article
Show Figures

Figure 1

Figure 1
<p>Streamlines emanating from two coherent optical sources [<a href="#B12-symmetry-16-00067" class="html-bibr">12</a>].</p>
Full article ">Figure 2
<p>A primitive image of the unfolding and enfolding of an event.</p>
Full article ">Figure 3
<p>The continuous enfolding and unfolding of events.</p>
Full article ">Figure 4
<p>Tower of Clifford algebras.</p>
Full article ">
20 pages, 6247 KiB  
Review
The 3 31 Nucleotide Minihelix tRNA Evolution Theorem and the Origin of Life
by Lei Lei and Zachary Frome Burton
Life 2023, 13(11), 2224; https://doi.org/10.3390/life13112224 - 19 Nov 2023
Cited by 4 | Viewed by 5875
Abstract
There are no theorems (proven theories) in the biological sciences. We propose that the 3 31 nt minihelix tRNA evolution theorem be universally accepted as one. The 3 31 nt minihelix theorem completely describes the evolution of type I and type II tRNAs [...] Read more.
There are no theorems (proven theories) in the biological sciences. We propose that the 3 31 nt minihelix tRNA evolution theorem be universally accepted as one. The 3 31 nt minihelix theorem completely describes the evolution of type I and type II tRNAs from ordered precursors (RNA repeats and inverted repeats). Despite the diversification of tRNAome sequences, statistical tests overwhelmingly support the theorem. Furthermore, the theorem relates the dominant pathway for the origin of life on Earth, specifically, how tRNAomes and the genetic code may have coevolved. Alternate models for tRNA evolution (i.e., 2 minihelix, convergent and accretion models) are falsified. In the context of the pre-life world, tRNA was a molecule that, via mutation, could modify anticodon sequences and teach itself to code. Based on the tRNA sequence, we relate the clearest history to date of the chemical evolution of life. From analysis of tRNA evolution, ribozyme-mediated RNA ligation was a primary driving force in the evolution of complexity during the pre-life-to-life transition. TRNA formed the core for the evolution of living systems on Earth. Full article
(This article belongs to the Special Issue Feature Papers in Origins of Life)
Show Figures

Figure 1

Figure 1
<p>Type II tRNA evolved via RNA ligation and a 9 nt internal deletion within ligated 3′- and 5′-acceptor stems [<a href="#B42-life-13-02224" class="html-bibr">42</a>]. Also, 3 31 nt minihelices (one D loop minihelix (magenta 17 nt core) and two anticodon stem-loop-stem minihelices (blue 17 nt core)) were fused by ligation for minihelix replication. The 93 nt tRNA precursor was processed by an internal 9 nt deletion (see below) within fused 3′-acceptor (yellow) and 5′-acceptor (green) stems. In the type II tRNA structure, the red arrow indicates the fusion of the magenta segment (17 nt D loop minihelix core; UAGCC repeat) and the green segment (5′-acceptor stem fragment; initially GGCGG). Abbreviations: SLS) stem-loop-stem; Ac) anticodon. Molecular graphics were created using ChimeraX [<a href="#B43-life-13-02224" class="html-bibr">43</a>].</p>
Full article ">Figure 2
<p>Type II tRNA resulted from failure to process a 14 nt V loop (initially a 7 nt 3′-acceptor stem ligated to a 7 nt 5′-acceptor stem) rather than by accretion. Colors: green) 5′-acceptor stem and 5′-acceptor stem fragment; magenta) 17 nt D loop core; cyan) 5′-anticodon and T stem; red) anticodon and T loops; cornflower blue) 3′-anticodon and T stem; and yellow) 3′-acceptor stem. Arrow colors: blue) U-turns; red) processing site in evolution; light yellow) discriminator base (D); and gold) site of amino acid placement. The structure is of an unmodified Pyrococcus horikoshii tRNA<sup>Leu</sup> in complex with LeuRS-IA. At the right of the figure are tRNA<sup>Leu</sup> and tRNA<sup>Ser</sup> V loops from Pyrococcus furiosis, an ancient Archaeon. Colors: red) V loop UAG that binds LeuRS-IA in tRNA<sup>Leu</sup> recognition in P. furiosis [<a href="#B44-life-13-02224" class="html-bibr">44</a>]; yellow) unpaired bases just 5′ of the Levitt base; and green) tRNA<sup>Ser</sup> bases at the 3′-end of the V loop. PRE indicates an initial pre-life sequence. Parentheses indicate stems; * indicates unpaired bases.</p>
Full article ">Figure 3
<p>Evolution of type I tRNA via RNA ligation and two related, internal 9 nt deletions. Colors and arrow colors are as in <a href="#life-13-02224-f001" class="html-fig">Figure 1</a> and <a href="#life-13-02224-f002" class="html-fig">Figure 2</a>. G* (OMG) is 2′-O-methyl-G. Also, 9 nt internal deletions generate a magenta-green fusion and a yellow-cyan fusion (red arrows).</p>
Full article ">Figure 4
<p>Type I tRNA. Colors and arrow colors are as in <a href="#life-13-02224-f001" class="html-fig">Figure 1</a> and <a href="#life-13-02224-f002" class="html-fig">Figure 2</a>. G* (OMG) is 2′-O-methyl-G. YYG is Wy-butosine [<a href="#B46-life-13-02224" class="html-bibr">46</a>]. The V loop (3′-As*; yellow) is fused to the cyan (5′-T stem), in slight contrast to type II tRNA processing (<a href="#life-13-02224-f001" class="html-fig">Figure 1</a> and <a href="#life-13-02224-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 5
<p>The 3 31 nt minihelix theorem, and evolution of tRNA world from polymer world and 31 nt minihelix world. The inset describes the 9 nt deletions to generate tRNAs: the more 5′ processing event involves deletion between the blue arrows; the more 3′ processing event (type I tRNA only) involves deletion between the red arrows. Internal deletions were at stem-loop junctions. Colors and arrows are consistent with previous figures. Yellow arrows mark the cornflower blue-yellow junction, indicating the degree of order in tRNA assembly.</p>
Full article ">Figure 6
<p>Features of polymer world. (<b>A</b>) The D loop minihelix core could function as a primitive translational adapter to recognize codon GGC. (<b>B</b>) The anticodon and T stem-loop-stems could function as a translational adapter. The dotted blue line indicates a Hoogsteen A–C pair that stabilizes the U-turn loop. Ligation of 3′-ACCA-Gly converted these sequences into primitive translational adapters in the pre-life world. Parentheses indicate paired bases. Asterisks indicate loop bases.</p>
Full article ">Figure 7
<p>Typical tDNA diagrams for Pyrococcus furiosis (an ancient Archaeon) (<b>A</b>) and Archaea (<b>B</b>) [<a href="#B5-life-13-02224" class="html-bibr">5</a>]. Arrow colors: red) processing positions for evolution of type I tRNAs; and blue) U-turns. Red lines indicate: (1) the Levitt reverse Watson–Crick base pair (G15-CV<sub>5</sub>) (Lbp for Levitt base pair); (2) intercalation of G18 between A57 and A58 (“elbow” contact); and (3) Watson–Crick interaction of G19 and C56 (“elbow” contact) [<a href="#B4-life-13-02224" class="html-bibr">4</a>].</p>
Full article ">Figure 8
<p>Evolution of (β−α)<sub>8</sub> barrels by RNA ligation, translation and pseudosymmetrical folding. β-sheets and α-helices are numbered in the figure.</p>
Full article ">Figure 9
<p>Refolding of a (β−α)<sub>8</sub> barrel generated a (β−α)<sub>8</sub> sheet. β7 lost its β-sheet partners in the refolding. Two views are shown. EST is estradiol. Helices are numbered in the left images. β-sheets are numbered in the right images.</p>
Full article ">Figure 10
<p>Generation of double-Ψ−β-barrels in pre-life. Numbers indicate β-sheets.</p>
Full article ">Figure 11
<p>Evolution of AARS enzymes. Class II AARS are simple homologs of class I AARS. The blue segments include homologous sequences, including a Zn finger. The red segment is unique to class I AARS and directs the distinct class I AARS fold. At the bottom of the figure are two alignments demonstrating homology of GlyRS-IIA (class II), IleRS-IA (class I) and ValRS-IA (class I). + indicates amino acid similarity. Expect values indicate homology of sequences.</p>
Full article ">
Back to TopTop