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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (608)

Search Parameters:
Keywords = molecular electron density theory

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 5051 KiB  
Article
Negative Solvatochromism of the Intramolecular Charge Transfer Band in Two Structurally Related Pyridazinium—Ylids
by Mihaela Iuliana Avădănei, Antonina Griţco-Todiraşcu and Dana Ortansa Dorohoi
Symmetry 2024, 16(11), 1531; https://doi.org/10.3390/sym16111531 - 15 Nov 2024
Viewed by 62
Abstract
Two charge transfer compounds based on pyridazinium ylids were studied by electronic absorption spectroscopy in binary and ternary solutions, with the purpose of evaluating their descriptors of the first singlet excited state and to estimate the strength of the intermolecular interactions in protic [...] Read more.
Two charge transfer compounds based on pyridazinium ylids were studied by electronic absorption spectroscopy in binary and ternary solutions, with the purpose of evaluating their descriptors of the first singlet excited state and to estimate the strength of the intermolecular interactions in protic solvents. The molecular descriptors of the excited state were comparatively estimated using the variational method and the Abe model of diluted binary solutions. Analysis of electronic properties using density functional theory was performed for several key solvents, in order to understand the solvatochromic behavior. The DFT calculations revealed that, in the polar and strongly interacting solvents, the carbanion and the terminal group become a stronger electron acceptor. The bathochromic shift of the ICT band was confirmed using DFT calculus. The ability of the two ylids to recognize and discriminate the solvents was analyzed with principal component analysis and with cluster analysis. Although the study was performed in 24 solvents, the results showed that the ylids were most sensitive to alcohols, so they can be a useful tool to identify and classify different types of low-alcoholic solvents. Full article
(This article belongs to the Collection Feature Papers in Chemistry)
Show Figures

Figure 1

Figure 1
<p>Energy minimized structures in the ground state of PPPyNiP and PyNiP in vacuo and with the corresponding atomic charges. Level of theory: DFT-ω-B97X-D/STO-3G/def2SV (carbon black; hydrogen grey; nitrogen blue; oxygen red).</p>
Full article ">Figure 2
<p>Plot of the Abe parameters B vs. A for (<b>a</b>) PPPyNiP and (<b>b</b>) PPNiP. The numbers correspond to the current numbering of the solvents listed in <a href="#symmetry-16-01531-t005" class="html-table">Table 5</a>.</p>
Full article ">Figure 3
<p>Graphical representation of <math display="inline"><semantics> <mrow> <mi>l</mi> <mi>n</mi> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> <mrow> <mn>1</mn> <mo>−</mo> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math> vs. <math display="inline"><semantics> <mrow> <mi>l</mi> <mi>n</mi> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <msub> <mrow> <mi>x</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> <mrow> <mn>1</mn> <mo>−</mo> <msub> <mrow> <mi>x</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math> for the binary solvents MeOH (1) + DMF (2) and MeOH (1) + DMSO (2) for: (<b>a</b>) PPPyNiP; (<b>b</b>) PyNiP.</p>
Full article ">Figure 4
<p>Graphical representation of <math display="inline"><semantics> <mrow> <mi>l</mi> <mi>n</mi> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> <mrow> <mn>1</mn> <mo>−</mo> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math> vs. <math display="inline"><semantics> <mrow> <mi>l</mi> <mi>n</mi> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <msub> <mrow> <mi>x</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> <mrow> <mn>1</mn> <mo>−</mo> <msub> <mrow> <mi>x</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </mfrac> </mstyle> </mrow> </semantics></math> for the binary solvents water (W) (1) + MeOH (2) and W (1) + EtOH (2) for (<b>a</b>) PPPyNiP and (<b>b</b>) PyNiP.</p>
Full article ">Figure 5
<p>The molecular electrostatic potential of PPPyNiP and PyNiP (isosurface value = 0.002).</p>
Full article ">Figure 6
<p>Energy diagram of the predicted frontier molecular orbitals of PPPyNiP (<b>a</b>) and PyNiP (<b>b</b>) and distribution of the electron density calculated for vacuo and three representative types of solvents: cyclohexane (CHX), methanol (MeOH), and dimethylsulfoxide (DMSO). Yellow regions correspond to the positive orbital phase, and the blue and red regions correspond to the negative orbital phase, respectively. Level of theory: DFT-ω-B97X-D/STO-3G/def2SV/IEF-PCM.</p>
Full article ">Figure 7
<p>The natural transition orbitals, HONTO and LUNTO, for the ICT transition of PPPyNiP, and the corresponding electronic density difference.</p>
Full article ">Figure 8
<p>The natural transition orbitals, HONTO and LUNTO, for the ICT transition of PyNiP, and the electronic density difference.</p>
Full article ">Figure 9
<p>PCA score plots for the response of PPPyNiP (<b>a</b>) and PyNiP (<b>b</b>) to twenty-four solvents. The first two factors were used, which describe around 85% of the total variance.</p>
Full article ">Figure 10
<p>2D <span class="html-italic">Louvain</span> clustering results for PPPyNiP in the principal component space.</p>
Full article ">Figure 11
<p>2D <span class="html-italic">Louvain</span> clustering results for PyNiP in the principal component space.</p>
Full article ">
14 pages, 10000 KiB  
Article
High-Efficiency Triple-Junction Polymer Solar Cell: A Theoretical Approach
by Fazli Sattar, Xiaozhuang Zhou and Zakir Ullah
Molecules 2024, 29(22), 5370; https://doi.org/10.3390/molecules29225370 - 14 Nov 2024
Viewed by 204
Abstract
This study presents the theoretical design and evaluation of a triple-junction polymer solar cell architecture, incorporating oligomers of PDCBT, PPDT2FBT, and PDPP3T as donor materials and PC71BM as the electron acceptor. Using density functional theory (DFT) simulations and time-dependent DFT (TD-DFT) [...] Read more.
This study presents the theoretical design and evaluation of a triple-junction polymer solar cell architecture, incorporating oligomers of PDCBT, PPDT2FBT, and PDPP3T as donor materials and PC71BM as the electron acceptor. Using density functional theory (DFT) simulations and time-dependent DFT (TD-DFT) methods, the investigation covers essential photovoltaic parameters, including molecular geometries, UV-Vis spectra, and charge transport properties. The device is structured to maximize solar energy absorption across the spectrum, featuring front, middle, and back junctions with band gaps of 1.9 eV, 1.63 eV, and 1.33 eV, respectively. Each layer targets different regions of the solar spectrum, optimizing light harvesting and charge separation. This innovative multi-junction design offers a promising pathway to enhanced power conversion efficiencies in polymer solar cells, advancing the integration of renewable energy technologies. Full article
Show Figures

Figure 1

Figure 1
<p>Optimized geometries of PDCBT, PPDT2FBT, and PDPP3T.</p>
Full article ">Figure 2
<p>UV-visible absorption spectra of PDCBT, PPDT2FBT, and PDPP3T.</p>
Full article ">Figure 3
<p>Experimental UV-Vis spectra of PDCBT-PC<sub>71</sub>BM, PPDT2FBT-PC<sub>71</sub>BM, and PDPP3T-PC<sub>71</sub>BM Complexes.</p>
Full article ">Figure 4
<p>Simulated UV-Vis spectra of PDCBT-PC<sub>71</sub>BM, PPDT2FBT-PC<sub>71</sub>BM, and PDPP3T-PC<sub>71</sub>BM Complexes.</p>
Full article ">Figure 5
<p>MEP plots of (<b>1</b>) PDCBT, (<b>2</b>) PPDT2FBT, and (<b>3</b>) PDPP3T polymer.</p>
Full article ">Figure 6
<p>MEP plots of PDCBT-PC<sub>71</sub>BM, PPDT2FBT-PC<sub>71</sub>BM, and PDPP3T-PC<sub>71</sub>BM Complexes.</p>
Full article ">Figure 7
<p>HOMO–LUMO plots of (<b>1</b>) PDCBT, (<b>2</b>) PPDT2FBT, and (<b>3</b>) PDPP3T polymer.</p>
Full article ">Figure 8
<p>HOMO–LUMO plots and electron density maps of (<b>1</b>) PDCBT-PC<sub>71</sub>BM, (<b>2</b>) PPDT2FBT-PC<sub>71</sub>BM, and (<b>3</b>) PDPP3T-PC<sub>71</sub>BM complexes along with electron density map surface contour.</p>
Full article ">Figure 9
<p>(<b>a</b>) Depicts the device structure of a conventional triple-junction solar cell, showing how different materials are stacked in the design to achieve high efficiency. (<b>b</b>) Displays the energy diagram of the conventional triple-junction solar cell, illustrating the energy levels of each layer and how charge carriers (electrons and holes) move through the system.</p>
Full article ">Figure 10
<p>Energy level diagram of PDPP3T, PPDT2FBT, PDPP3T, and PC<sub>71</sub>BM system.</p>
Full article ">
12 pages, 5062 KiB  
Article
A DFT Study of Band-Gap Tuning in 2D Black Phosphorus via Li+, Na+, Mg2+, and Ca2+ Ions
by Liuhua Mu, Jie Jiang, Shiyu Gao, Xiao-Yan Li and Shiqi Sheng
Int. J. Mol. Sci. 2024, 25(21), 11841; https://doi.org/10.3390/ijms252111841 - 4 Nov 2024
Viewed by 583
Abstract
Black phosphorus (BP) and its two-dimensional derivative (2D-BP) have garnered significant attention as promising anode materials for electrochemical energy storage devices, including next-generation fast-charging batteries. However, the interactions between BP and light metal ions, as well as how these interactions influence BP’s electronic [...] Read more.
Black phosphorus (BP) and its two-dimensional derivative (2D-BP) have garnered significant attention as promising anode materials for electrochemical energy storage devices, including next-generation fast-charging batteries. However, the interactions between BP and light metal ions, as well as how these interactions influence BP’s electronic properties, remain poorly understood. Here, we employed density functional theory (DFT) to investigate the effects of monovalent (Li+ and Na+) and divalent (Mg2+ and Ca2+) ions on the valence electronic structure of 2D-BP. Molecular orbital analysis revealed that the adsorption of divalent cations can significantly reduce the band gap, suggesting an enhancement in charge transfer rates. In contrast, the adsorption of monovalent cations had minimal impact on the band gap, suggesting the preservation of 2D-BP’s intrinsic electrical properties. Energetic and charge analyses indicated that the extent of charge transfer primarily governs the ability of ions to modulate 2D-BP’s electronic structure, especially under high-pressure conditions where ions are in close proximity to the 2D-BP surface. Moreover, charge polarization calculations revealed that, compared with monovalent cations, divalent cations induced greater polarization, disrupting the symmetry of the pristine 2D-BP and further influencing its electronic characteristics. These findings provide a molecular-level understanding of how ion interactions influence 2D-BP’s electronic properties during ion-intercalation processes, where ions are in close proximity to the 2D-BP surface. Moreover, the calculated diffusion barrier results revealed the potential of 2D-BP as an effective anode material for lithium-ion, sodium-ion, and magnesium-ion batteries, though its performance may be limited for calcium-ion batteries. By extending our understanding of interactions between ions and 2D-BP, this work contributes to the design of efficient and reliable energy storage technologies, particularly for the next-generation fast-charging batteries. Full article
(This article belongs to the Special Issue Adsorption Materials and Adsorption Behavior: 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Spatial localization of frontier molecular orbitals in (<b>a</b>) Li<sup>+</sup>@2D-BP and (<b>b</b>) Na<sup>+</sup>@2D-BP complexes. The molecular orbital is plotted for iso-values of ±0.02 atomic units with orange and green denoting opposite signs. Spheres in red, blue, pink, and white represent Li<sup>+</sup>, Na<sup>+</sup>, phosphorus, and hydrogen, respectively.</p>
Full article ">Figure 2
<p>Spatial localization of frontier molecular orbitals in (<b>a</b>) Mg<sup>2+</sup>@2D-BP and (<b>b</b>) Ca<sup>2+</sup>@2D-BP complexes. The molecular orbital is plotted for iso-values of ±0.02 atomic units, with orange and green denoting opposite signs. Spheres in green, orange, pink, and white represent Mg<sup>2+</sup>, Ca<sup>2+</sup>, phosphorus, and hydrogen, respectively.</p>
Full article ">Figure 3
<p>Evolution of frontier molecular orbital energy levels in (<b>a</b>) Li<sup>+</sup>@2D-BP, (<b>b</b>) Na<sup>+</sup>@2D-BP, (<b>c</b>) Mg<sup>2+</sup>@2D-BP, and (<b>d</b>) Ca<sup>2+</sup>@2D-BP complexes as a function of separation distance (<span class="html-italic">H</span>). Upon the introduction of a cation into the system, the energy levels of both the HOMO and LUMO of 2D-BP (denoted as BP-HOMO and BP-LUMO) underwent a significant decrease.</p>
Full article ">Figure 4
<p>Binding energy of (<b>a</b>) Li<sup>+</sup>@2D-BP and Na<sup>+</sup>@2D-BP, as well as (<b>b</b>) Mg<sup>2+</sup>@2D-BP and Ca<sup>2+</sup>@2D-BP complexes as a function of separation distance (<span class="html-italic">H</span>).</p>
Full article ">Figure 5
<p>ADCH atomic charges on 2D-BP and the ion for (<b>a</b>) Li<sup>+</sup>@2D-BP, (<b>b</b>) Na<sup>+</sup>@2D-BP, (<b>c</b>) Mg<sup>2+</sup>@2D-BP, and (<b>d</b>) Ca<sup>2+</sup>@2D-BP complexes as a function of separation distance (<span class="html-italic">H</span>).</p>
Full article ">Figure 6
<p>Polarizing action of an ion adsorbed on the 2D-BP sheet, expressed as <math display="inline"><semantics> <mrow> <mi>ξ</mi> <mo>=</mo> <msup> <mrow> <mo>(</mo> <mrow> <msubsup> <mo stretchy="false">∑</mo> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> </mrow> <mrow> <mi>N</mi> </mrow> </msubsup> <mrow> <msubsup> <mrow> <mi>q</mi> </mrow> <mrow> <mi>i</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msubsup> </mrow> </mrow> <mo>)</mo> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math>, where <math display="inline"><semantics> <msub> <mrow> <mi>q</mi> </mrow> <mrow> <mi>i</mi> </mrow> </msub> </semantics></math> represents the partial charges on each atom in the 2D-BP sheet. The dotted horizontal line corresponds to <span class="html-italic">ξ</span> of the isolated 2D-BP sheet.</p>
Full article ">Figure 7
<p>Diffusion energy barriers for (<b>a</b>) Li<sup>+</sup>/Na<sup>+</sup> and (<b>b</b>) Mg<sup>2+</sup>/Ca<sup>2+</sup> ions on a 2D-BP sheet along two diffusion paths, armchair and zigzag, as illustrated in the inset of panel (<b>a</b>). The diffusion calculations were based on the respective separation distances at the potential energy minimum.</p>
Full article ">
21 pages, 7976 KiB  
Article
The Impact of Helium and Nitrogen Plasmas on Electrospun Gelatin Nanofiber Scaffolds for Skin Tissue Engineering Applications
by Abolfazl Mozaffari, Mazeyar Parvinzadeh Gashti, Farbod Alimohammadi and Mohammad Pousti
J. Funct. Biomater. 2024, 15(11), 326; https://doi.org/10.3390/jfb15110326 - 1 Nov 2024
Viewed by 687
Abstract
This study explores the fabrication of tannic acid-crosslinked gelatin nanofibers via electrospinning, followed by helium and nitrogen plasma treatment to enhance their biofunctionality, which was assessed using fibroblast cells. The nanofibers were characterized using scanning electron microscopy, atomic force microscopy, attenuated total reflection [...] Read more.
This study explores the fabrication of tannic acid-crosslinked gelatin nanofibers via electrospinning, followed by helium and nitrogen plasma treatment to enhance their biofunctionality, which was assessed using fibroblast cells. The nanofibers were characterized using scanning electron microscopy, atomic force microscopy, attenuated total reflection Fourier transform infrared spectroscopy, X-ray diffraction, and water contact angle measurements before and after treatment. Helium and nitrogen gas plasma were employed to modify the nanofiber surfaces. Results indicated that helium and nitrogen plasma treatment significantly increased the hydrophilicity and biofunctionality of the nanofibers by 5.1° ± 0.6 and 15.6° ± 2.2, respectively, making them more suitable for human skin fibroblast applications. To investigate the impact of plasma treatment on gelatin, we employed a computational model using density functional theory with the B3LYP/6-31+G(d) method. This model represented gelatin as an amino acid chain composed of glycine, hydroxyproline, and proline, interacting with plasma particles. Vibrational analysis of these systems was used to interpret the vibrational spectra of untreated and plasma-treated gelatin. To further correlate with experimental findings, molecular dynamics simulations were performed on a system of three interacting gelatin chains. These simulations explored changes in amino acid bonding. The computational results align with experimental observations. Comprehensive analyses confirmed that these treatments improved hydrophilicity and biofunctionality, supporting the use of plasma-treated gelatin nanofibers in skin tissue engineering applications. Gelatin’s natural biopolymer properties and the versatility of plasma surface modification techniques underscore its potential in regenerating cartilage, skin, circulatory tissues, and hamstrings. Full article
(This article belongs to the Collection Feature Papers in Biomaterials for Healthcare Applications)
Show Figures

Figure 1

Figure 1
<p>Optimized structure of a triple-helical collagen model composed of three 30-residue chains.</p>
Full article ">Figure 2
<p>Optimized geometry of Gly-Pro-Hyp (GPH) at the B3LYP/6-31+G(d) level of theory. Color code: carbon (green), oxygen (red), nitrogen (blue), hydrogen (cyan).</p>
Full article ">Figure 3
<p>Optimized structure of plasma-treated GPH at B3LYP/6-31+G(d) with helium (<b>a</b>) and nitrogen (<b>b</b>). The occupied volumes are 293.98 Å<sup>3</sup> and 378.24 Å<sup>3</sup>, respectively. Atoms are color-coded as follows: carbon (green), oxygen (red), nitrogen (blue), and hydrogen (cyan).</p>
Full article ">Figure 4
<p>SEM images of electrospun gelatin nanofibers: plasma untreated at ×5000 (<b>A</b>), plasma untreated at ×10,000 (<b>B</b>), helium plasma treatment at ×5000 (<b>C</b>), helium plasma treatment at ×10,000 (<b>D</b>), nitrogen plasma treatment at ×5000 (<b>E</b>), and nitrogen plasma treatment at ×10,000 (<b>F</b>).</p>
Full article ">Figure 5
<p>AFM image of untreated gelatin nanofibers (<b>A</b>), helium plasma-treated gelatin nanofibers (<b>B</b>), and nitrogen plasma-treated gelatin nanofibers (<b>C</b>).</p>
Full article ">Figure 6
<p>ATR-FTIR waveforms of untreated gelatin nanofibers, helium plasma-treated gelatin nanofibers, and nitrogen plasma-treated gelatin nanofibers.</p>
Full article ">Figure 7
<p>Calculated IR spectra of raw, helium plasma-, and nitrogen plasma-treated samples at B3LYP/6-31+G(d).</p>
Full article ">Figure 8
<p>Comparison of optimized geometries for untreated (<b>top</b>) and nitrogen-treated (<b>bottom</b>) gelatin using the Amber99 force field. Atoms are color-coded: carbon (green), oxygen (red), nitrogen (blue), and hydrogen (white). Hydrogen bonds (polar contacts) are indicated by dotted lines.</p>
Full article ">Figure 9
<p>XRD spectra of untreated, helium plasma-treated, and nitrogen plasma-treated electrospun gelatin nanofibers.</p>
Full article ">Figure 10
<p>Images of fibroblast cells on a gelatin scaffold: (<b>A</b>) a SEM image of the helium plasma-treated gelatin scaffold; (<b>B</b>) a SEM image of the nitrogen plasma-treated gelatin scaffold; (<b>C</b>) an inverted optical microscope image of the helium plasma-treated gelatin scaffolds at 250× magnification; and (<b>D</b>) an inverted optical microscope image of the nitrogen plasma-treated gelatin scaffold at 250× magnification.</p>
Full article ">
24 pages, 6781 KiB  
Article
A Structure and Magnetism Study of {MnII3MnIVLnIII3} Coordination Complexes with Ln = Dy, Yb
by Victoria Mazalova, Tatiana Asanova, Igor Asanov and Petra Fromme
Inorganics 2024, 12(11), 286; https://doi.org/10.3390/inorganics12110286 - 31 Oct 2024
Viewed by 444
Abstract
We report the research results of polynuclear complexes consisting of 3d-4f mixed-metal cores that are maintained by acetate ligands and multidentate Schiff base ligands with structurally exposed thioether groups. The presence of the latter at the periphery of these neutral compounds enables their [...] Read more.
We report the research results of polynuclear complexes consisting of 3d-4f mixed-metal cores that are maintained by acetate ligands and multidentate Schiff base ligands with structurally exposed thioether groups. The presence of the latter at the periphery of these neutral compounds enables their anchoring onto substrate surfaces. Specifically, we investigated the electronic and magnetic properties as well as the structural arrangement in {MnII3MnIVLnIII3} with Ln = Dy, Yb coordination complexes using various complementary methods. We studied the electronic and atomic structure of the target compounds using the XAS and XES techniques. The molecular structures of the compounds were determined using density functional theory, and the magnetic data were obtained as a function of the magnetic field. Using the XMCD method, we followed the changes in the electronic and magnetic properties of adsorbed magnetic compounds induced by the reaction of ligands through interaction with the substrate. The complexes show antiferromagnetic exchange interactions between Mn and Ln ions. The spectroscopic analyses confirmed the structural and electronic integrity of complexes in organic solution. This study provides important input for a full understanding of the dependence of the magnetic properties and the molecule–substrate interaction of single adsorbed molecules on the type of ligands. It highlights the importance of chemical synthesis for controlling and tailoring the magnetic properties of metalorganic molecules for their use as optimized building blocks of future molecular spin electronics. Full article
(This article belongs to the Section Coordination Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular structure of the {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} coordination complex. Color code: C of acetates, green; C of the Schiff base ligand L∙SMe<sup>–</sup>, gray; Dy, turquoise; Mn<sup>II</sup>, pink; Mn<sup>IV</sup>, purple; N, blue; O, red and S, yellow. H atoms and solvent molecules are omitted for clarity.</p>
Full article ">Figure 2
<p>(<b>a</b>) Theoretical Mn K-edge XANES spectra of four manganese atoms of the {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>D<sup>III</sup><sub>3</sub>} coordination complex; (<b>b</b>) comparison of the experimental Mn K-edge spectrum with a total spectrum of all four nonequivalent Mn atoms; (<b>c</b>) Fourier transforms of Mn K-edge k<sup>3</sup> weighed EXAFS. Mn1…Mn4 are the numbers of the Mn atoms according to <a href="#inorganics-12-00286-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Comparison of the experimental Mn L<sub>2,3</sub> XAS spectra of {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} and {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Yb<sup>III</sup><sub>3</sub>} coordination complexes with ones for Mn(acac)<sub>2</sub>, Mn(acac)<sub>3</sub>, and MnO<sub>2</sub> reference compounds.</p>
Full article ">Figure 4
<p>(<b>left</b>): partial density of states for Mn<sup>2+</sup> and Mn<sup>4+</sup> atoms ((<b>top</b>) and (<b>middle</b>) panels) and Dy atoms ((<b>bottom</b>) panel); (<b>right</b>): the experimental Mn L<sub>2,3</sub> XAS spectra of the {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} coordination complex measured at the PEAXIS (green line) and 4IDC (black line) beamlines (<b>top</b>) and the calculated total density of states spectra for Mn and Dy atoms (<b>bottom</b>). The Fermi level is set at 0 eV.</p>
Full article ">Figure 5
<p>(<b>left</b>): partial density of states averaged for all four Mn atoms in the molecule. The position of the LUMO level is marked with vertical lines; (<b>right</b>): total density of states for Mn and Dy atoms in the molecule. The experimental spectrum (<b>top right</b>) is shifted to the LUMO level. The Fermi level is marked with a vertical line.</p>
Full article ">Figure 6
<p>DFT calculated HOMO-1, HOMO, LUMO, and LUMO + 1 molecular orbitals (an isosurface value is 0.01). Blue and red isosurfaces refer to the positive and negative spins, respectively.</p>
Full article ">Figure 7
<p>The assignment of atoms in the {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} coordination complex (<b>left</b>) and the spin density distribution (<b>right</b>) corresponding to the data presented in <a href="#inorganics-12-00286-t001" class="html-table">Table 1</a>. Blue and red surfaces on the right picture refer to the positive and negative spin density, respectively.</p>
Full article ">Figure 8
<p>(<b>a</b>) XAS spectrum of the {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} coordination complex with excitation energies for RXES measurements marked by the red vertical lines; (<b>b</b>) Selected RXES spectra measured at various excitation energies; (<b>c</b>) The reconstructed RXES map; (<b>d</b>) RXES spectra plotted as energy loss spectra by setting the energy of the elastic peak to 0 eV.</p>
Full article ">Figure 9
<p>Left panel: Mn L<sub>2,3</sub>-edges XAS (<b>a</b>), XMCD (<b>b</b>), and Dy M<sub>4,5</sub>-edges XAS (<b>c</b>); XMCD (<b>d</b>) of the {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} complexes. Labels µ<sup>+</sup> and µ<sup>−</sup> stand for the left and right directions of the photon helicity. Positive direction of the applied magnetic field is collinear with the photon propagation direction. XMCD spectra measured at the 5T magnetic field and T = 6 K.</p>
Full article ">Figure 10
<p>Mn L<sub>2,3</sub>-edges XAS (<b>a</b>), XMCD (<b>b</b>), and Dy M<sub>4,5</sub>-edges XAS (<b>c</b>); XMCD (<b>d</b>) of the {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} coordination complexes deposited on the Au substrate. Labels µ<sup>+</sup> and µ<sup>−</sup> stand for the left and right directions of the photon helicity. In (<b>b</b>,<b>d</b>), the measured spectra are indicated in black and their averaged values in blue.</p>
Full article ">Figure 11
<p>Comparison of Mn L<sub>2,3</sub>-edge (<b>left panel</b>) and Dy M<sub>4,5</sub>-edge (<b>right panel</b>) XAS (<b>a</b>,<b>c</b>) and XMCD (<b>b</b>,<b>d</b>) spectra of free and deposited on the Au substrate {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} coordination complexes.</p>
Full article ">Figure 12
<p>Mn L<sub>2,3</sub>-edge XAS and XMCD of the {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Yb<sup>III</sup><sub>3</sub>} coordination complexes deposited on the Au substrate measured at the magnetic field of 5T (<b>left panel</b>) and −5T (<b>right panel</b>). Labels (<b>a</b>,<b>c</b>) and (<b>b</b>,<b>d</b>) correspond to the XAS and XMCD spectra, respectively.</p>
Full article ">Figure 13
<p>Comparison of Mn L<sub>2,3</sub>-edges (<b>left panel</b>) and Dy M<sub>4,5</sub>-edges (<b>right panel</b>) XAS (<b>a</b>,<b>c</b>) and XMCD (<b>b</b>,<b>d</b>) spectra of {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Dy<sup>III</sup><sub>3</sub>} and {Mn<sup>II</sup><sub>3</sub>Mn<sup>IV</sup>Yb<sup>III</sup><sub>3</sub>} coordination complexes deposited on the Au substrate.</p>
Full article ">
15 pages, 2781 KiB  
Article
An MEDT Study of the Reaction Mechanism and Selectivity of the Hetero-Diels–Alder Reaction Between 3-Methylene-2,4-Chromandione and Methyl Vinyl Ether
by Abderrazzak Bouhaoui, Aziz Moumad, Luis R. Domingo and Latifa Bouissane
Molecules 2024, 29(21), 5109; https://doi.org/10.3390/molecules29215109 - 29 Oct 2024
Viewed by 397
Abstract
The hetero-Diels–Alder (HDA) reaction between the ambident heterodiene 3-methylene-2,4-chromandione (MCDO) and non-symmetric methyl vinyl ether (MVE) is investigated using the molecular electron density theory (MEDT) at the B3LYP/6-311G(d,p) computational level. The aim of this study is to gain insight into its molecular mechanism [...] Read more.
The hetero-Diels–Alder (HDA) reaction between the ambident heterodiene 3-methylene-2,4-chromandione (MCDO) and non-symmetric methyl vinyl ether (MVE) is investigated using the molecular electron density theory (MEDT) at the B3LYP/6-311G(d,p) computational level. The aim of this study is to gain insight into its molecular mechanism and to elucidate the factors that control the selectivity found experimentally. DFT-based reactivity indices reveal that MCDO exhibits strong electrophilic characteristics, while MVE displays a strong nucleophilic character. Meanwhile, the Parr function explains the ortho regioselectivity of this HDA reaction. The highly polar nature of this HDA reaction, supported by the high global electron density transfer (GEDT) taking place at the transition state structures (TSs), accounts for the very low activation energy associated with the most favorable TS-4on. The ambident nature of MCDO allows for the formation of two constitutional isomeric cycloadducts. In the case of MVE, pseudocyclic selectivity is attained using a thermodynamic control. This polar HDA reaction displays an endo stereoselectivity and a complete ortho regioselectivity. A comparative relative interacting atomic energy (RIAE) analysis of the two diastereomeric structures TS-4on and TS-6on indicates a high degree of likeness, which explains the low pseudocyclic selectivity under kinetic control. Full article
(This article belongs to the Special Issue Feature Papers in Computational and Theoretical Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>B3LYP/6-311G(d,p) ELF basin attractor positions, populations of the most relevant valence basins, and natural atomic charges of MCDO <b>1</b> and MVE <b>15</b>. Valence basin populations and natural atomic charges are given as average number of electrons, e. Negative charges are shown in red, and positive charges are shown in blue.</p>
Full article ">Figure 2
<p>3D representations of the Mulliken atomic spin densities of the radical anion of MCDO <b>1</b> and of the radical cation of MVE <b>15</b>, together with the electrophilic <math display="inline"><semantics> <mrow> <msubsup> <mi>P</mi> <mi>k</mi> <mo>+</mo> </msubsup> </mrow> </semantics></math> Parr functions of MCDO <b>1</b> and the nucleophilic <math display="inline"><semantics> <mrow> <msubsup> <mi>P</mi> <mi>k</mi> <mo>−</mo> </msubsup> <mrow> <mtext> </mtext> <mi>Parr</mi> </mrow> </mrow> </semantics></math> functions of MVE <b>15</b>.</p>
Full article ">Figure 3
<p>B3LYP/6-311G(d,p) Gibbs free energy profiles, ΔG in kcal·mol<sup>−1</sup>, computed at 101.1 °C in dioxane for the more favorable regioisomeric <span class="html-italic">ortho</span> reaction paths associated with HDA reaction of MCDO <b>1</b> with MVE <b>15</b>.</p>
Full article ">Figure 4
<p>B3LYP/6-311G(d,p) geometries of the most favorable regioisomeric <span class="html-italic">ortho</span> TSs associated with HDA reaction of MCDO <b>1</b> with MVE <b>15</b>. Values in dioxane are given in parenthesis. Distances are given in angstroms Å.</p>
Full article ">Figure 5
<p>View of the diastereomeric <b>TS-4on</b> and <b>TS-6ox</b> along the axis formed by the interacting C1 and C8 carbons. The C2–C1–C8–C9 dihedral angles are given in degrees.</p>
Full article ">Figure 6
<p>ELF basin attractor positions and populations of the most relevant valence basins of <b>TS-4on</b> and <b>TS-6on</b>. Valence basin populations are given in average number of electrons, e.</p>
Full article ">Figure 7
<p>Positions of the CPs in <b>TS-4on</b> and <b>TS-6on,</b> with (3, +1) in gray and (3, −1) in pink.</p>
Full article ">Figure 8
<p>Graphical representation of the sets of the ξ<math display="inline"><semantics> <mrow> <msubsup> <mi>E</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> <mi>a</mi> <mi>l</mi> </mrow> <mrow> <mi>E</mi> <mi>t</mi> </mrow> </msubsup> </mrow> </semantics></math>, ξ<math display="inline"><semantics> <mrow> <msubsup> <mi>E</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> <mi>a</mi> <mi>l</mi> </mrow> <mrow> <mi>M</mi> <mi>C</mi> <mi>D</mi> <mi>O</mi> </mrow> </msubsup> </mrow> </semantics></math>, and ξ<math display="inline"><semantics> <mrow> <msubsup> <mi>E</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> <mi>a</mi> <mi>l</mi> </mrow> <mrow> <mi>M</mi> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mo>+</mo> <mi>E</mi> <mi>t</mi> </mrow> </msubsup> </mrow> </semantics></math> energies in that order for the TSs associated with the HDA reactions of MCDO <b>1</b> with ethylene <b>8</b> and MVE <b>15</b>. The ξ<math display="inline"><semantics> <mrow> <msubsup> <mi>E</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> <mi>a</mi> <mi>l</mi> </mrow> <mrow> <mi>M</mi> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mo>+</mo> <mi>E</mi> <mi>t</mi> </mrow> </msubsup> </mrow> </semantics></math> energies correspond to the RIAE activation energies of these HDA reactions. ξ<math display="inline"><semantics> <mrow> <msubsup> <mi>E</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> <mi>a</mi> <mi>l</mi> </mrow> <mi>X</mi> </msubsup> </mrow> </semantics></math> energies of the Et and MCDO frameworks at the four TSs are shown in blue and red, respectively. The black bar represents the ξ<math display="inline"><semantics> <mrow> <msubsup> <mi>E</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> <mi>a</mi> <mi>l</mi> </mrow> <mrow> <mi>M</mi> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mo>+</mo> <mi>E</mi> <mi>t</mi> </mrow> </msubsup> </mrow> </semantics></math> energies. The GEDT values at the TSs are given in pink. ξ<math display="inline"><semantics> <mrow> <msubsup> <mi>E</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> <mi>a</mi> <mi>l</mi> </mrow> <mi>X</mi> </msubsup> </mrow> </semantics></math> energies are given in kcal mol<sup>−1</sup>, and the GEDT is given in average number of electrons, e.</p>
Full article ">Scheme 1
<p>HDA reactions of MCDO <b>1</b> with olefins <b>2</b> and <b>3</b>.</p>
Full article ">Scheme 2
<p>HDA reactions of the ambident compound <b>7</b> with ethylene <b>8</b>.</p>
Full article ">Scheme 3
<p>Competitive HDA and Alder-ene reactions in the Lewis acid catalyzed reaction between isoprene <b>11</b> and formaldehyde <b>12</b>.</p>
Full article ">Scheme 4
<p>The eight competitive reaction paths associated with the HDA reaction of MCDO <b>1</b> with MVE <b>15</b>.</p>
Full article ">
13 pages, 1570 KiB  
Article
In Silico Design of Novel Piperazine-Based mTORC1 Inhibitors Through DFT, QSAR and ADME Investigations
by El Mehdi Karim, Oussama Abchir, Hassan Nour, Ossama Daoui, Souad El Khattabi, Farhan Siddique, M’Hammed El Kouali, Mohammed Talbi, Abdelkbir Errougui and Samir Chtita
Biophysica 2024, 4(4), 517-529; https://doi.org/10.3390/biophysica4040034 - 24 Oct 2024
Viewed by 484
Abstract
Mammalian target of rapamycin complex 1 (mTORC1) is an important and promising alternative biological target for the treatment of different types of cancer including breast, lung and renal cell carcinoma. This study contributed to the development of mathematical models highlighting the quantitative structure-activity [...] Read more.
Mammalian target of rapamycin complex 1 (mTORC1) is an important and promising alternative biological target for the treatment of different types of cancer including breast, lung and renal cell carcinoma. This study contributed to the development of mathematical models highlighting the quantitative structure-activity relationship of a series of piperazine derivatives reported as mTORC1 inhibitors. Various molecular descriptors were calculated using Gaussian 09, Chemsketch, and ChemOffice software. The density funcional theory (DFT) method at the level B3LYP/6-31G+(d, p) was applied to determine the structural, electronic and energetic parameters associated with the studied molecules. The predictive ability of the built models, which is obtained by two methods (MLR and MNLR), showed that the built models are statistically significant. The QSAR modeling results revealed that the six molecular descriptors of lowest unoccupied molecular orbital energy (ELUMO), electrophilicity index (ω), molar refractivity (MR), aqueous solubility (Log S), topological polar surface area (PSA), and refractive index (n) significantly correlated to the biological inhibitory activity of piperazine derivatives. Using QSAR models and in silico pharmacokinetic profiles predictions, five new candidate compounds are selected as potential inhibitors against cancer. Full article
(This article belongs to the Special Issue Molecular Structure and Simulation in Biological System 3.0)
Show Figures

Figure 1

Figure 1
<p>Structures of currently available natural mTORC1 inhibitors.</p>
Full article ">Figure 2
<p>The basic skeleton of piperazine derivatives.</p>
Full article ">Figure 3
<p>Correlation between observed activity values and predicted activity values via the MLR model.</p>
Full article ">Figure 4
<p>Correlation between observed activity values and predicted activity values via the MNLR model.</p>
Full article ">Figure 5
<p>William’s plot of leverage-normalized residuals for the MLR pIC<sub>50</sub> model (with <span class="html-italic">h</span>* = 0.58 and residual limits y = ±2.5).</p>
Full article ">Figure 6
<p>3D structures of selected molecules (2, 4, 13, 25, and 71) by evaluating pharmacokinetic properties ADME.</p>
Full article ">
19 pages, 2336 KiB  
Article
Crystallographic and Optical Spectroscopic Study of Metal–Organic 2D Polymeric Crystals of Silver(I)– and Zinc(II)–Squarates
by Bojidarka Ivanova
Crystals 2024, 14(10), 905; https://doi.org/10.3390/cryst14100905 - 18 Oct 2024
Viewed by 533
Abstract
Metal–organic framework materials, as innovative functional materials for nonlinear optical technologies, feature linear and nonlinear optical responses, such as a laser damage threshold, outstanding mechanical properties, thermal stability, and optical transparency. Their non-centrosymmetric crystal structure induces a higher-order nonlinear optical response, which guarantees [...] Read more.
Metal–organic framework materials, as innovative functional materials for nonlinear optical technologies, feature linear and nonlinear optical responses, such as a laser damage threshold, outstanding mechanical properties, thermal stability, and optical transparency. Their non-centrosymmetric crystal structure induces a higher-order nonlinear optical response, which guarantees technological applications. ZnII– and AgI–squarate complexes are attractive templates for these purposes due to their good crystal growth, optical transparency, high thermal stability, etc. However, the space group type of the catena-((μ2-squarato)-tetra-aqua-zinc(II)) complex ([Zn(C4O4)(H2O)4]) is debatable, (1) showing centro- and non-centrosymmetric monoclinic C2/c and Cc phases. The same is valid for the catena-((μ3-squarato)-(μ2-aqua)-silver(I)) complex (Ag2C4O4), (2) exhibiting, so far, only a C2/c phase. This study is the first to report new crystallographic data on (1) and (2) re-determined at different temperatures (293(2) and 300(2)K) and the non-centrosymmetric Cc phase of (2), having different numbers of molecules per unit cell compared with the C2/c phase. There are high-resolution crystallographic measurements of single crystals, experimental electronic absorption, and vibrational spectroscopic data, together with ultra-high-resolution mass spectrometric ones. The experimental results are supported for theoretical optical and nonlinear optical properties obtained via high-accuracy static computational methods and molecular dynamics, using density functional theory as well as chemometrics. Full article
(This article belongs to the Special Issue Exploring the Frontier of MOFs through Crystallographic Studies)
Show Figures

Figure 1

Figure 1
<p>Pluton plots of crystals (<b>1</b>) and (<b>2</b>).</p>
Full article ">Figure 2
<p>Unit cell content of crystals (1) (<b>A</b>) and (2) (<b>B</b>) viewed from different perspectives; solvent accessible surfaces.</p>
Full article ">Figure 3
<p>Experimental and theoretical (M062X/LANL2DZ) spectra of crystal of (2) and its optimized molecular geometry at various level of theory; the experimental spectra were measured in solvent water.</p>
Full article ">
11 pages, 11784 KiB  
Article
Rational Design of High-Performance Photocontrolled Molecular Switches Based on Chiroptical Dimethylcethrene: A Theoretical Study
by Li Han, Mei Wang, Yifan Zhang, Bin Cui and Desheng Liu
Molecules 2024, 29(20), 4912; https://doi.org/10.3390/molecules29204912 - 17 Oct 2024
Viewed by 457
Abstract
The reversible photo-induced conformation transition of a single molecule with a [5]helicene backbone has garnered considerable interest in recent studies. Based on such a switching process, one can build molecular photo-driven switches for potential applications of nanoelectronics. But the achievement of high-performance reversible [...] Read more.
The reversible photo-induced conformation transition of a single molecule with a [5]helicene backbone has garnered considerable interest in recent studies. Based on such a switching process, one can build molecular photo-driven switches for potential applications of nanoelectronics. But the achievement of high-performance reversible single-molecule photoswitches is still rare. Here, we theoretically propose a 13,14-dimethylcethrene switch whose photoisomerization between the ring-closed and ring-open forms can be triggered by ultraviolet (UV) and visible light irradiation. The electronic structure transitions and charge transport characteristics, concurrent with the photo-driven electrocyclization of the molecule, are calculated by the non-equilibrium Green’s function (NEGF) in combination with density functional theory (DFT). The electrical conductivity bears great diversity between the closed and open configurations, certifying the switching behavior and leading to a maximum on–off ratio of up to 103, which is considerable in organic junctions. Further analysis confirms the evident switching behaviors affected by the molecule–electrode interfaces in molecular junctions. Our findings are helpful for the rational design of organic photoswitches at the single-molecule level based on cethrene and analogous organic molecules. Full article
Show Figures

Figure 1

Figure 1
<p>A schematic diagram of (<b>a</b>) the photoisomerization process of 13,14-dimethylcethrene and (<b>b</b>) the corresponding 45° top views and (<b>c</b>) side views.</p>
Full article ">Figure 2
<p>A schematic diagram of (<b>a</b>) molecular junctions based on 13,14-dimethylcethrene with Au(111) electrodes. Each junction is composed of three distinct components, i.e., the left electrode, the central scattering region, and the right electrode. (<b>b</b>) Forty-five-degree top views of ring-<span class="html-italic">closed</span> and ring-<span class="html-italic">open</span> isomers with different anchor attachment sites denoted as I(II)-<span class="html-italic">closed</span> and I(II)-<span class="html-italic">open</span>, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) The current–voltage (I–V) curves of the molecular junctions based on 13,14-dimethylcethrene with different molecule–electrode interfaces. The solid black and green lines represent the I–V curves of the closed and open configurations with connection I. The solid blue and red lines represent the I–V curves of the closed and open configurations with connection II. (<b>b</b>) The <span class="html-italic">on–off</span> ratio curves of the molecular junctions. The solid red and blue lines represent the <span class="html-italic">on–off</span> ratio curves of the I-configuration and II-configuration, respectively. The inset in the top right corner presents an enlarged depiction of the <span class="html-italic">on–off</span> ratio curve for the I-configuration.</p>
Full article ">Figure 4
<p>The bias-dependent transmission spectra of the (<b>a</b>) closed and (<b>b</b>) open configurations consist of connection I. (<b>c</b>,<b>d</b>) The same case consists of connection II. The white dotted line indicates the bias window.</p>
Full article ">Figure 5
<p>The bias-dependent projected density of states (PDOS) spectra of the (<b>a</b>) closed and (<b>b</b>) open configurations consists of connection I. (<b>c</b>,<b>d</b>) The same case consists of connection II. The projection subspace is segmented into three parts, i.e., the left electrode (green line), the molecule (red line), and the right electrode (blue line). The magenta line indicates the bias window.</p>
Full article ">Figure 6
<p>The zero-bias transmission spectra of the (<b>a</b>) I-configuration and (<b>c</b>) II-configuration based on 13,14-dimethylcethrene. The solid red and blue lines represent the transmission spectra of the closed and open configurations, respectively. The Fermi level is set to zero, and the magenta line indicates the bias window. The red and blue solid down-pointing triangles denote the MPSH eigenvalues of the <span class="html-italic">closed</span> and <span class="html-italic">open</span> molecules. H and L denote HOMO and LUMO. (<b>b</b>,<b>d</b>) The bias-dependent spatial distribution of the frontier molecular orbitals in the I-configuration and II-configuration. The isovalue is set to 0.035 for all plots.</p>
Full article ">
20 pages, 5554 KiB  
Article
Syn-Propanethial S-Oxide as an Available Natural Building Block for the Preparation of Nitro-Functionalized, Sulfur-Containing Five-Membered Heterocycles: An MEDT Study
by Mikołaj Sadowski, Ewa Dresler, Karolina Zawadzińska, Aneta Wróblewska and Radomir Jasiński
Molecules 2024, 29(20), 4892; https://doi.org/10.3390/molecules29204892 - 15 Oct 2024
Viewed by 941
Abstract
The regio- and stereoselectivity and the molecular mechanisms of the [3 + 2] cycloaddition reactions between Syn-propanethial S-oxide and selected conjugated nitroalkenes were explored theoretically in the framework of the Molecular Electron Density Theory. It was found that cycloadditions with the participation [...] Read more.
The regio- and stereoselectivity and the molecular mechanisms of the [3 + 2] cycloaddition reactions between Syn-propanethial S-oxide and selected conjugated nitroalkenes were explored theoretically in the framework of the Molecular Electron Density Theory. It was found that cycloadditions with the participation of nitroethene as well as its methyl- and chloro-substituted analogs can be realized via a single-step mechanism. On the other hand, [3 + 2] cycloaddition reactions between Syn-propanethial S-oxide and 1,1-dinitroethene can proceed according to a stepwise mechanism with a zwitterionic intermediate. Finally, we evaluated the affinity of model reaction products for several target proteins: cytochrome P450 14α-sterol demethylase CYP51 (RSCB Database PDB ID: 1EA1), metalloproteinase gelatinase B (MMP-9; PDB ID: 4XCT), and the inhibitors of cyclooxygenase COX-1 (PDB:3KK6) and COX-2 (PDB:5KIR). Full article
(This article belongs to the Special Issue Heterocyclic Compounds: Synthesis, Application and Theoretical Study)
Show Figures

Figure 1

Figure 1
<p>Topology of molecule <b>1</b> ELF, as rendered at an isovalue of 0.8. Core basins are given in magenta, protonated basins in cyan, disynaptic basins in green, and monosynaptic basins in red.</p>
Full article ">Figure 2
<p>Positions and populations of ELF attractors of unprotonated basins in molecule <b>1</b>.</p>
Full article ">Figure 3
<p>Views of critical structures for the [3 + 2] cycloaddition between <span class="html-italic">Syn</span>-propanethial S-oxide (<b>1</b>) and nitroethene <b>2a</b> in toluene solution according to wB97XD/6-31G(d) (PCM) calculations.</p>
Full article ">Figure 4
<p>Visualization of ELF basins of system <b>IA</b> of reaction <b>1</b> + <b>2d</b> at an ELF isovalue of 0.8, as computed by wB97XD/6-31G(d) in toluene (PCM). Core basins are given in magenta, protonated basins in cyan, disynaptic basins in green, and monosynaptic basins in red.</p>
Full article ">Figure 5
<p>Populations of ELF attractors in intermediate <b>IA</b> of reaction <b>1</b> + <b>2d</b> and populations of significant basins, as computed by wB97XD/6-31G(d)) in toluene (PCM).</p>
Full article ">Figure 6
<p>The natural charges of the atoms in intermediate <b>IA</b> of reaction <b>1</b> + <b>2d</b> (WB97XD/6-31G(d) in toluene (PCM)). Charges greater than 0.2 are given in red, those less than −0.2 are given in blue, and those between or equal to −0.2 and 0.2 are given in black.</p>
Full article ">Figure 7
<p>Views of critical structures for path A of the [3 + 2] cycloaddition between <span class="html-italic">Syn</span>-propanethial S-oxide (<b>1</b>) and 1,1-dinitroethene <b>2d</b> in toluene solution, according to wB97XD/6-31G(d) (PCM) calculations.</p>
Full article ">Figure 8
<p>Visualization of hydrogen bond formation between ligand and protein for (<b>A</b>) <b>4a</b>, (<b>B</b>) <b>3a</b>, and COX-2 (PDB:5kir).</p>
Full article ">Figure 9
<p>Visualization of hydrogen bond formation between ligand and protein for (<b>A</b>) <b>6a</b>, (<b>B</b>) <b>5a</b>, and CYP51 (PDB:1ea1).</p>
Full article ">Figure 10
<p>Visualization of (<b>A</b>) electrostatic and (<b>B</b>) hydrophobic–hydrophilic potential of CYP51 (PDB:1EA1) with <b>6a</b> in the pocket.</p>
Full article ">Scheme 1
<p>Examples of bioactive 1,3-oxathiolane derivatives.</p>
Full article ">Scheme 2
<p>Single-step and stepwise mechanisms of [3 + 2] cycloaddition reactions.</p>
Full article ">Scheme 3
<p>Theoretically possible regio- and stereoisomeric paths of the [3 + 2] cycloaddition between <span class="html-italic">Syn</span>-propanethial S-oxide (<b>1</b>) and 1-R-1-nitroethenes (<b>2a</b>–<b>d</b>).</p>
Full article ">Scheme 4
<p>Geometry of the CSO group of oxide <b>1</b> calculated at the ground state.</p>
Full article ">Scheme 5
<p>Natural charges of atoms (in electrons) in the S-oxide Xet, as computed at the ground state in the gaseous phase (B3LYP/6-31G(d)); charges &gt; 0.2 e are given in red, and charges &lt; −0.2 are given in blue.</p>
Full article ">Scheme 6
<p>Parr function values for radical anion and radical cation of molecule <b>1</b>.</p>
Full article ">
17 pages, 8670 KiB  
Article
A New Insight into the Molecular Mechanism of the Reaction between 2-Methoxyfuran and Ethyl (Z)-3-phenyl-2-nitroprop-2-enoate: An Molecular Electron Density Theory (MEDT) Computational Study
by Mikołaj Sadowski, Ewa Dresler, Aneta Wróblewska and Radomir Jasiński
Molecules 2024, 29(20), 4876; https://doi.org/10.3390/molecules29204876 - 14 Oct 2024
Viewed by 606
Abstract
The molecular mechanism of the reaction between 2-methoxyfuran and ethyl (Z)-3-phenyl-2-nitroprop-2-enoate was investigated using wb97xd/6-311+G(d,p)(PCM) quantum chemical calculations. It was found that the most probable reaction mechanism is fundamentally different from what was previously postulated. In particular, six possible zwitterionic intermediates [...] Read more.
The molecular mechanism of the reaction between 2-methoxyfuran and ethyl (Z)-3-phenyl-2-nitroprop-2-enoate was investigated using wb97xd/6-311+G(d,p)(PCM) quantum chemical calculations. It was found that the most probable reaction mechanism is fundamentally different from what was previously postulated. In particular, six possible zwitterionic intermediates were detected on the reaction pathway. Their formation is determined by the nature of local nucleophile/electrophile interactions. Additionally, the channel involving the formation of the exo-nitro Diels–Alder cycloadduct was completely ruled out. Finally, the electronic nature of the five- and six-membered nitronates as potential TACs was evaluated. Full article
Show Figures

Figure 1

Figure 1
<p>Views of key structures for the formation of zwitterionic intermediates via the <span class="html-italic">endo</span>-attack of 2-methoxyfuran (<b>1</b>) on the ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>) molecule according to wb97xd/6-311+G(d,p) (PCM) calculations.</p>
Full article ">Figure 2
<p>Views of key structures for the formation of zwitterionic intermediates via the <span class="html-italic">exo</span>-attack of 2-methoxyfuran (<b>1</b>) on the ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>) molecule according to wb97xd/6-311+G(d,p) (PCM) calculations.</p>
Full article ">Figure 3
<p>Topology of <b>I<span class="html-italic">endo</span>1</b> ELF, rendered at an isovalue of 0.8. Core basins are shown in magenta, protonated basins in cyan, disynaptic basins in green, and monosynaptic basins in red. Parts of the function significant for intermediate-type identification are depicted as solid, while the rest are translucent.</p>
Full article ">Figure 4
<p>Positions and populations of significant ELF attractors of unprotonated basins in <b>I<span class="html-italic">endo</span>1</b>.</p>
Full article ">Figure 5
<p>Views of key structures for the formation of DA adducts from isomeric zwitterions formed in the reaction between 2-methoxyfuran (<b>1</b>) and ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>) according to wb97xd/6-311+G(d,p) (PCM) calculations.</p>
Full article ">Figure 6
<p>Views of key structures for the formation of HDA adducts from isomeric zwitterions formed in the reaction between 2-methoxyfuran (<b>1</b>) and ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>) according to wb97xd/6-311+G(d,p) (PCM) calculations.</p>
Full article ">Figure 7
<p>Views of key structures for the formation of Michael-type adducts from isomeric zwitterions formed in the reaction between 2-methoxyfuran (<b>1</b>) and ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>) according to wb97xd/6-311+G(d,p) (PCM) calculations.</p>
Full article ">Figure 8
<p>Views of key structures for the formation of the <b><span class="html-italic">Z</span>-4</b> nitronate in the reaction between 2-methoxyfuran (<b>1</b>) and ethyl (Z)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>) according to wb97xd/6-311+G(d,p) (PCM) calculations.</p>
Full article ">Figure 9
<p>Enthalpy profiles for the formation of the <b><span class="html-italic">Z</span>-3</b>, <b><span class="html-italic">E</span>-3</b>, and <b><span class="html-italic">Z</span>-4</b> adducts in the reaction between 2-methoxyfuran (<b>1</b>) and ethyl (Z)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>) according to wb97xd/6-311+G(d,p) (PCM) calculations.</p>
Full article ">Figure 10
<p>Topology of <b><span class="html-italic">Z</span>-4</b> (<b>left</b>) and <b>HDA<span class="html-italic">exo</span></b> (<b>right</b>) ELF, rendered at an isovalue of 0.8. Core basins are shown in magenta, protonated basins in cyan, disynaptic basins in green, and monosynaptic basins in red. Parts of the function, significant for intermediate type identification, are depicted as solid, while the rest are translucent.</p>
Full article ">Figure 11
<p>Positions and populations of significant ELF attractors of unprotonated basins in <b><span class="html-italic">Z</span>–4</b> (<b>left</b>) and <b>HDA<span class="html-italic">exo</span></b> (<b>right</b>).</p>
Full article ">Scheme 1
<p>Experimental results of the reaction between 2-methoxyfuran (<b>1</b>) and ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>).</p>
Full article ">Scheme 2
<p>Postulated mechanisms for the reaction between 2-methoxy (<b>1</b>) and ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>).</p>
Full article ">Scheme 3
<p>Considered paths of the reaction between 2-methoxyfuran (<b>1</b>) and ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>).</p>
Full article ">Scheme 4
<p>Local electronic properties of 2-methoxyfuran (<b>1</b>) and ethyl (<span class="html-italic">Z</span>)-3-phenyl-2-nitroprop-2-enoate (<b>2</b>).</p>
Full article ">Scheme 5
<p>Natural charges of significant atoms in the structure <b>I<span class="html-italic">endo</span>1</b>, calculated using the wb97xd/6-311+G(d,p) (PCM) level of theory.</p>
Full article ">Scheme 6
<p>Natural charges of significant atoms in structures <b><span class="html-italic">Z</span>–4</b> (<b>left</b>) and <b>HDA<span class="html-italic">exo</span></b> (<b>right</b>), as computed using WB97XD/6-311+G(d,p) pcm CH<sub>2</sub>Cl<sub>2</sub>.</p>
Full article ">
20 pages, 824 KiB  
Communication
Oxidation Potential of 2,6-Dimethyl-1,4-dihydropyridine Derivatives Estimated by Structure Descriptors
by Lorentz Jäntschi
Symmetry 2024, 16(10), 1320; https://doi.org/10.3390/sym16101320 - 7 Oct 2024
Viewed by 576
Abstract
Linear relationships, expressing the electrochemical properties of molecules as functions of structure, give insight into the associated electrochemical process and are a tool for prediction. Many biological activities rely on water-based dissociation, making electrochemical properties a bridge between structure and activity. Motivated by [...] Read more.
Linear relationships, expressing the electrochemical properties of molecules as functions of structure, give insight into the associated electrochemical process and are a tool for prediction. Many biological activities rely on water-based dissociation, making electrochemical properties a bridge between structure and activity. Motivated by a previous study, a replica is made here on a different dataset in order to validate/invalidate the previously reported results. There are several methods for obtaining structure-based descriptors. Some of the methods have been devised to account for molecular topology, some to account for molecular geometry, and others to account for both. Two methods are involved here to derive structure-based descriptors and further obtain structure–property relationships (FMPI and ChPE). In order to express structure descriptors, both FMPI and ChPE express first the topology of the molecule, using the heavy atoms identity matrix and the heavy atoms adjacency matrix, both square symmetric matrices in the belief that symmetry is one major factor of molecular stability. A set of 2,6-dimethyl-1,4-dihydropyridine derivatives with oxidation peak potentials and coulometrically determined number of electrons experimental data is subjected to the search for structure–activity relationships. Even if the 2,6-dimethyl-1,4-dihydropyridine is a symmetric compound (of Cs point group), their derivatives are generally not symmetric (9 out of 24 are asymmetric). The dataset is subjected to descriptive and inferential statistics in order to filter out the most relevant structure–activity relationship. The geometry is built using three levels of theory (one from molecular mechanics and two others from density functionals, of which one accounts for the interaction with water as solvent). One challenge of picking one out of two reported measured values is dealt with by calculating the likelihood associated with the two choices. Relevant structure–activity models are extracted and discussed. The use of in vivo (in water, SM8 model) models in geometry optimization (from MMFF94 and B3LYP and to M06 + Water SM8) results in a precision gain, but this is, in most of the cases, not statistically significant, and this can be considered a negative result. Full article
(This article belongs to the Special Issue Symmetry/Asymmetry of Molecules Related to Biological Activity)
Show Figures

Figure 1

Figure 1
<p>2,6-Dimethyl-1,4-dihydropyridine.</p>
Full article ">Figure 2
<p>2,6-Dimethyl-1,4-dihydropyridine scaffold.</p>
Full article ">Figure 3
<p>Order statistics on rejecting the hypothesis of samples normality, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>24</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Order statistics on rejecting the hypothesis of samples normality, <math display="inline"><semantics> <mrow> <msup> <mi>n</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>FMPI descriptors predicting <math display="inline"><semantics> <msup> <mi>E</mi> <mrow> <mi>o</mi> <mi>x</mi> </mrow> </msup> </semantics></math> of all (<math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>24</mn> </mrow> </semantics></math>) 2,6-dimethyl-1,4-dihydropyridine derivatives (experimental as a predict function, <math display="inline"><semantics> <mrow> <mi>y</mi> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <msub> <mn>0.49</mn> <mrow> <mo>±</mo> <mn>0.09</mn> </mrow> </msub> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <mi>LFUUBS</mi> <mo>−</mo> <mn>0</mn> <mo>.</mo> <msub> <mn>84</mn> <mrow> <mo>±</mo> <mn>0.37</mn> </mrow> </msub> </mrow> </semantics></math> with rhombs and <math display="inline"><semantics> <mrow> <mi>M</mi> <mn>3</mn> <mo>(</mo> <mi>IJUUFM</mi> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>7</mn> </mrow> </msup> <mo>,</mo> <mi>LFUUBN</mi> <mo>,</mo> <mo>−</mo> <msub> <mn>17</mn> <mrow> <mo>±</mo> <mn>2</mn> </mrow> </msub> <mo>,</mo> <msub> <mn>1.9</mn> <mrow> <mo>±</mo> <mn>0.2</mn> </mrow> </msub> <mo>,</mo> <msub> <mn>3.6</mn> <mrow> <mo>±</mo> <mn>0.4</mn> </mrow> </msub> <mo>,</mo> <mo>−</mo> <msub> <mn>0.38</mn> <mrow> <mo>±</mo> <mn>0.04</mn> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> with squares).</p>
Full article ">Figure 6
<p>ChPE and FMPI descriptors predicting <math display="inline"><semantics> <msup> <mi>E</mi> <mrow> <mi>o</mi> <mi>x</mi> </mrow> </msup> </semantics></math> of M06w stable (<math display="inline"><semantics> <mrow> <msup> <mi>n</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>) 2,6-dimethyl-1,4-dihydropyridine derivatives (experimental as a predict function, <math display="inline"><semantics> <mrow> <mi>y</mi> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <msub> <mn>1.01</mn> <mrow> <mo>±</mo> <mn>0.04</mn> </mrow> </msub> <mo>−</mo> <mn>7</mn> <mo>.</mo> <msub> <mn>3</mn> <mrow> <mo>±</mo> <mn>1.4</mn> </mrow> </msub> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <mrow> <mi>LGCP</mi> <mn>0619</mn> </mrow> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math> with rhombs and <math display="inline"><semantics> <mrow> <mi>M</mi> <mn>3</mn> <mo>(</mo> <mi>REETGM</mi> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>5</mn> </msup> <mo>,</mo> <mi>IFEUDN</mi> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>8</mn> </mrow> </msup> <mo>,</mo> <mo>−</mo> <msub> <mn>4.2</mn> <mrow> <mo>±</mo> <mn>0.6</mn> </mrow> </msub> <mo>,</mo> <mo>−</mo> <msub> <mn>1.9</mn> <mrow> <mo>±</mo> <mn>0.6</mn> </mrow> </msub> <mo>,</mo> <mo>−</mo> <msub> <mn>1.9</mn> <mrow> <mo>±</mo> <mn>0.5</mn> </mrow> </msub> <mo>,</mo> <mo>−</mo> <msub> <mn>3.2</mn> <mrow> <mo>±</mo> <mn>0.6</mn> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> with squares).</p>
Full article ">Figure 7
<p>ChPE descriptors predicting <math display="inline"><semantics> <msub> <mi>n</mi> <mi>e</mi> </msub> </semantics></math> (with <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>1.32</mn> </mrow> </semantics></math>) of M06w stable (<math display="inline"><semantics> <mrow> <msup> <mi>n</mi> <mo>′</mo> </msup> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>) 2,6-dimethyl-1,4-dihydropyridine derivatives: experimental as a predict function with <math display="inline"><semantics> <mrow> <mi>M</mi> <mn>3</mn> <mo>(</mo> <mrow> <mi>IGcP</mi> <mn>0847</mn> </mrow> <mspace width="0.166667em"/> <mi>·</mi> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>2</mn> </msup> <mo>,</mo> <mrow> <mi>ICtN</mi> <mn>0524</mn> </mrow> <mo>,</mo> <msub> <mn>1.10</mn> <mrow> <mo>±</mo> <mn>0.05</mn> </mrow> </msub> <mo>,</mo> <mo>−</mo> <msub> <mn>0.45</mn> <mrow> <mo>±</mo> <mn>0.20</mn> </mrow> </msub> <mo>,</mo> <msub> <mn>0.44</mn> <mrow> <mo>±</mo> <mn>0.15</mn> </mrow> </msub> <mo>,</mo> <mo>−</mo> <msub> <mn>1.2</mn> <mrow> <mo>±</mo> <mn>0.3</mn> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">
16 pages, 1401 KiB  
Article
Quelling the Geometry Factor Effect in Quantum Chemical Calculations of 13C NMR Chemical Shifts with the Aid of the pecG-n (n = 1, 2) Basis Sets
by Yuriy Yu. Rusakov, Valentin A. Semenov and Irina L. Rusakova
Int. J. Mol. Sci. 2024, 25(19), 10588; https://doi.org/10.3390/ijms251910588 - 1 Oct 2024
Viewed by 552
Abstract
A root factor for the accuracy of all quantum chemical calculations of nuclear magnetic resonance (NMR) chemical shifts is the quality of the molecular equilibrium geometry used. In turn, this quality depends largely on the basis set employed at the geometry optimization stage. [...] Read more.
A root factor for the accuracy of all quantum chemical calculations of nuclear magnetic resonance (NMR) chemical shifts is the quality of the molecular equilibrium geometry used. In turn, this quality depends largely on the basis set employed at the geometry optimization stage. This parameter represents the main subject of the present study, which is a continuation of our recent work, where new pecG-n (n = 1, 2) basis sets for the geometry optimization were introduced. A goal of this study was to compare the performance of our geometry-oriented pecG-n (n = 1, 2) basis sets against the other basis sets in massive calculations of 13C NMR shielding constants/chemical shifts in terms of their efficacy in reducing geometry factor errors. The testing was carried out with both large-sized biologically active natural products and medium-sized compounds with complicated electronic structures. The former were treated using the computation protocol based on the density functional theory (DFT) and considered in the theoretical benchmarking, while the latter were treated using the computational scheme based on the upper-hierarchy coupled cluster (CC) methods and were used in the practical benchmarking involving the comparison with experimental NMR data. Both the theoretical and practical analyses showed that the pecG-1 and pecG-2 basis sets resulted in substantially reduced geometry factor errors in the calculated 13C NMR chemical shifts/shielding constants compared to their commensurate analogs, with the pecG-2 basis set being the best of all the considered basis sets. Full article
(This article belongs to the Section Physical Chemistry and Chemical Physics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Compounds used in theoretical analysis (set <b>1</b>). Blue, red, yellow and gray balls represent nitrogen, oxygen, carbon and hydrogen atoms, respectively.</p>
Full article ">Figure 2
<p>MAEs for the <sup>13</sup>C NMR shielding constants calculated for set <b>1</b> using the equilibrium geometries obtained using different basis sets (listed along the abscissa) compared to the corresponding reference theoretical data. The red numbers indicate the sizes of the basis sets for the elements of the second period. In these calculations, only the lowest energy conformers were taken into account.</p>
Full article ">Figure 3
<p>Compounds used in the analysis based on the comparison with experimental data (set <b>2</b>).</p>
Full article ">Figure 4
<p>MAEs for the <sup>13</sup>C NMR chemical shifts calculated for test set <b>2</b> using equilibrium geometries obtained using different basis sets (listed along the abscissa) against the corresponding experimental data. The second bars show the altered statistical figures evaluated without taking into account the chemical shift of C<sub>1</sub> of DMAc and fluorobenzene. The red numbers indicate the sizes of the basis sets for the second period elements.</p>
Full article ">Figure 5
<p>Correlation plot for the <sup>13</sup>C NMR shielding constants of test set <b>2</b> calculated at the GIAO-CCSD(T)/pecS-2 level using equilibrium geometries obtained at the CCSD/pecG-2 level against the corresponding experimental data.</p>
Full article ">
14 pages, 4803 KiB  
Article
Nature of Charge Transfer Effects in Complexes of Dopamine Derivatives Adsorbed on Graphene-Type Nanostructures
by Alex-Adrian Farcaş and Attila Bende
Int. J. Mol. Sci. 2024, 25(19), 10522; https://doi.org/10.3390/ijms251910522 - 29 Sep 2024
Viewed by 572
Abstract
Continuing the investigation started for dopamine (DA) and dopamine-o-quinone (DoQ) (see, the light absorption and charge transfer properties of the dopamine zwitterion (called dopamine-semiquinone or DsQ) adsorbed on the graphene nanoparticle surface is investigated using the ground state and linear-response time-dependent density functional [...] Read more.
Continuing the investigation started for dopamine (DA) and dopamine-o-quinone (DoQ) (see, the light absorption and charge transfer properties of the dopamine zwitterion (called dopamine-semiquinone or DsQ) adsorbed on the graphene nanoparticle surface is investigated using the ground state and linear-response time-dependent density functional theories, considering the ωB97X-D3BJ/def2-TZVPP level of theory. In terms of the strength of molecular adsorption on the surface, the DsQ form has 50% higher binding energy than that found in our previous work for the DA or DoQ cases (−20.24 kcal/mol vs. −30.41 kcal/mol). The results obtained for electronically excited states and UV-Vis absorption spectra show that the photochemical behavior of DsQ is more similar to DA than that observed for DoQ. Of the three systems analyzed, the DsQ-based complex shows the most active charge transfer (CT) phenomenon, both in terms of the number of CT-like states and the amount of charge transferred. Of the first thirty electronically excited states computed for the DsQ case, eleven are purely of the CT type, and nine are mixed CT and localized (or Frenkel) excitations. By varying the adsorption distance between the molecule and the surface vertically, the amount of charge transfer obtained for DA decreases significantly as the distance increases: for DoQ it remains stable, for DsQ there are states for which little change is observed, and for others, there is a significant change. Furthermore, the mechanistic compilation of the electron orbital diagrams of the individual components cannot describe in detail the nature of the excitations inside the complex. Full article
(This article belongs to the Special Issue Photochemistry in Molecular Clusters)
Show Figures

Figure 1

Figure 1
<p>The 2D chemical structure of (<b>a</b>) dopamine (or DA), (<b>b</b>) dopamine-o-quinone (or DoQ) and (<b>c</b>) dopamine-semiquinone (or DsQ).</p>
Full article ">Figure 2
<p>The geometry configuration of the dopamine-semiquinone (or DsQ) adsorbed on the GrNP surface computed at ωB97X-D3BJ/def2-TZVPP/CPCM level of theory.</p>
Full article ">Figure 3
<p>Theoretical UV absorption spectra of the graphene–DsQ binary complex and the individual constituents computed at the TDDFT/ωB97X-D3BJ/def2-TZVPP/CPCM level of theory.</p>
Full article ">Figure 4
<p>The molecular orbital energy scheme (in eV) of the individual, GrNP, and DsQ components and of the mixed GrNP–DsQ binary complex (H = HOMO (or Highest Occupied Molecular Orbital), L = LUMO (or Lowest Unoccupied Molecular Orbital)) based on the fragment orbital contribution analysis.</p>
Full article ">Figure 5
<p>The transferred charge (between 0 and 1 values of the elementary charge) calculated for different plane distances relative to the equilibrium geometry (z<sub>0</sub> + Δz, Δz between −0.3 Å and +1.0 Å) for DA (<b>a</b>), DoQ (<b>b</b>) and DsQ (<b>c</b>), respectively, adsorbed on the GrNP surface, computed at the TDDFT/ωB97X-D3BJ/def2-TZVPP level of theory.</p>
Full article ">Figure 6
<p>Molecular orbital energy schemes (in eV) built based on the fragment orbital contribution analysis of the individual, GrNP (1st col.), DA, DoQ, and DsQ (last col.) components and of the mixed (<b>a</b>) GrNP–DA, (<b>b</b>) GrNP–DoQ, and (<b>c</b>) GrNP–DsQ binary complexes (H = HOMO, L = LUMO) computed for Δ<span class="html-italic">z</span> = −0.3 (2nd col.), 0.0 (3rd col.), and +1.0 (4th col.) relative stacking distances.</p>
Full article ">
9 pages, 1716 KiB  
Article
Adsorption and Catalytic Reduction of Nitrogen Oxides (NO, N2O) on Disulfide Cluster Complexes of Cobalt and Iron—A Density Functional Study
by Ellie L. Uzunova and Ivelina M. Georgieva
Materials 2024, 17(19), 4764; https://doi.org/10.3390/ma17194764 - 28 Sep 2024
Viewed by 554
Abstract
The reactivity of nitrogen oxide, NO, as a ligand in complexes with [Fe2-S2] and [Co2-S2] non-planar rhombic cores is examined by density functional theory (DFT). The cobalt-containing nitrosyl complexes are less stable than the iron complexes because the Co-S bonds in the [Co2-S2] [...] Read more.
The reactivity of nitrogen oxide, NO, as a ligand in complexes with [Fe2-S2] and [Co2-S2] non-planar rhombic cores is examined by density functional theory (DFT). The cobalt-containing nitrosyl complexes are less stable than the iron complexes because the Co-S bonds in the [Co2-S2] core are weakened upon NO coordination. Various positions of NO were examined, including its binding to sulfur centers. The release of NO molecules can be monitored photochemically. The ability of NO to form a (NO)2 dimer provides a favorable route of electrochemical reduction, as protonation significantly stabilizes the dimeric species over the monomers. The quasilinear dimer ONNO, with trans-orientation of oxygen atoms, gains higher stability under protonation and reduction via proton–electron transfer. The first two reduction steps lead to an N2O intermediate, whose reduction is more energy demanding: in the two latter reaction steps the highest energy barrier for Co2S2(CO)6 is 109 kJ mol−1, and for Fe2S2(CO)6, it is 133 kJ mol−1. Again, the presence of favorable light absorption bands allows for a photochemical route to overcome these energy barriers. All elementary steps are exothermic, and the final products are molecular nitrogen and water. Full article
Show Figures

Figure 1

Figure 1
<p>Mixed nitrosyl–carbonyl and pure nitrosyl complexes: (<b>a</b>) Co<sub>2</sub>S<sub>2</sub>(CO)<sub>5</sub>(NO); (<b>b</b>) Co<sub>2</sub>S<sub>2</sub>(CO)<sub>4</sub>(NO)<sub>2</sub>; (<b>c</b>) Co<sub>2</sub>S<sub>2</sub>(NO)<sub>4</sub>; and (<b>d</b>) Co<sub>2</sub>S<sub>2</sub>(CO)<sub>6</sub>(NO) with S-NO bond. Legend: cobalt cations are light-blue large balls, sulfur atoms are yellow, nitrogen is dark blue, oxygen is red, and carbon is gray. N-O bond lengths are marked red, and C-O bond lengths are black. The M-S bonds, M-N bonds, and S-N bonds are described in <a href="#materials-17-04764-t002" class="html-table">Table 2</a>. The corresponding Fe complexes are presented in <a href="#app1-materials-17-04764" class="html-app">Figure S3 in Supplementary Materials</a>.</p>
Full article ">Figure 2
<p>The reaction path of N<sub>2</sub>O<sub>2</sub> dissociation to N<sub>2</sub>O on Co<sub>2</sub>S<sub>2</sub>(CO)<sub>6</sub> and Fe<sub>2</sub>S<sub>2</sub>(CO)<sub>6</sub>. TS1 denotes the energy barrier of reaction step (1) and TS2is the energy barrier of elementary step (2). RC—reaction coordinate.</p>
Full article ">Figure 3
<p>The reaction path of N<sub>2</sub>O dissociation to N<sub>2</sub> and H<sub>2</sub>O on hexacarbonyl complexes with Co<sub>2</sub>S<sub>2</sub> and Fe<sub>2</sub>S<sub>2</sub> core. TS3 denotes the energy barrier to reaction step (3) and TS4 is the energy barrier to elementary step (4). RC—reaction coordinate.</p>
Full article ">Figure 4
<p>The binding of the OH<sup>∙</sup> group to Fe<sub>2</sub>S<sub>2</sub>(CO)<sub>6</sub>. (<b>a</b>) Midway position of the OH<sup>∙</sup> group bonded directly to the Fe centers; (<b>b</b>) OH<sup>∙</sup> group bonded to the sulfur atom. Configuration (<b>a</b>) is the global minimum, found as 30 kJ mol<sup>−1</sup> below configuration (<b>b</b>). Iron cations are large aqua-blue balls, sulfur atoms are yellow, nitrogen is dark blue, oxygen is red, and carbon is gray. The cobalt complex is presented in <a href="#app1-materials-17-04764" class="html-app">Figure S7 in Supplementary Materials</a>.</p>
Full article ">Figure 5
<p>Transition state structures along the reaction path of ONNO reduction for the [Co2-S2] complex. TS1 reveals the first reduction step of ONNO. TS2 corresponds to the reduction of ONNOH<sup>∙</sup>. TS3 refers to the reduction of N<sub>2</sub>O. Legend is the same as <a href="#materials-17-04764-f001" class="html-fig">Figure 1</a>. The transition states TS1, TS2, TS3, and TS4 (water molecule release) for the [Fe2-S2] complex are presented as <a href="#app1-materials-17-04764" class="html-app">Supplementary Materials in Figure S7</a>, and their coordinates are included.</p>
Full article ">
Back to TopTop