[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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
remove_circle_outline
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
remove_circle_outline

Search Results (2,471)

Search Parameters:
Keywords = MD simulations

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
27 pages, 801 KiB  
Article
Maximizing Computation Rate for Sustainable Wireless-Powered MEC Network: An Efficient Dynamic Task Offloading Algorithm with User Assistance
by Huaiwen He, Feng Huang, Chenghao Zhou, Hong Shen and Yihong Yang
Mathematics 2024, 12(16), 2478; https://doi.org/10.3390/math12162478 (registering DOI) - 10 Aug 2024
Viewed by 260
Abstract
In the Internet of Things (IoT) era, Mobile Edge Computing (MEC) significantly enhances the efficiency of smart devices but is limited by battery life issues. Wireless Power Transfer (WPT) addresses this issue by providing a stable energy supply. However, effectively managing overall energy [...] Read more.
In the Internet of Things (IoT) era, Mobile Edge Computing (MEC) significantly enhances the efficiency of smart devices but is limited by battery life issues. Wireless Power Transfer (WPT) addresses this issue by providing a stable energy supply. However, effectively managing overall energy consumption remains a critical and under-addressed aspect for ensuring the network’s sustainable operation and growth. In this paper, we consider a WPT-MEC network with user cooperation to migrate the double near–far effect for the mobile node (MD) far from the base station. We formulate the problem of maximizing long-term computation rates under a power consumption constraint as a multi-stage stochastic optimization (MSSO) problem. This approach is tailored for a sustainable WPT-MEC network, considering the dynamic and varying MEC network environment, including randomness in task arrivals and fluctuating channels. We introduce a virtual queue to transform the time-average energy constraint into a queue stability problem. Using the Lyapunov optimization technique, we decouple the stochastic optimization problem into a deterministic problem for each time slot, which can be further transformed into a convex problem and solved efficiently. Our proposed algorithm works efficiently online without requiring further system information. Extensive simulation results demonstrate that our proposed algorithm outperforms baseline schemes, achieving approximately 4% enhancement while maintain the queues stability. Rigorous mathematical analysis and experimental results show that our algorithm achieves O(1/V),O(V) trade-off between computation rate and queue stability. Full article
(This article belongs to the Section Mathematics and Computer Science)
Show Figures

Figure 1

Figure 1
<p>System model of WPMEC network with user-assistance.</p>
Full article ">Figure 2
<p>An illustrative time division structure.</p>
Full article ">Figure 3
<p>Average task computation rate and average task queue length over time slots.</p>
Full article ">Figure 4
<p>Average task computation rates with different control parameter <span class="html-italic">V</span>.</p>
Full article ">Figure 5
<p>Task queue lengths with different control parameter <span class="html-italic">V</span>.</p>
Full article ">Figure 6
<p>Average task computation rate and task queue length with different energy constraint <math display="inline"><semantics> <mi>γ</mi> </semantics></math>.</p>
Full article ">Figure 7
<p>Convergence performance of energy consumption with different parameter <span class="html-italic">V</span>.</p>
Full article ">Figure 8
<p>Offloading power of FU and NU with different Bandwidth <span class="html-italic">W</span>.</p>
Full article ">Figure 9
<p>Average task computation rates in different schemes over time slots.</p>
Full article ">Figure 10
<p>Average computation rates in different schemes with different bandwidth <span class="html-italic">W</span>.</p>
Full article ">Figure 11
<p>Average computation rates in different schemes with different distances between FU and NU.</p>
Full article ">Figure 12
<p>Average computation rates in different schemes with different task arrival rates of FU.</p>
Full article ">
14 pages, 16469 KiB  
Article
Rational Design of Non-Covalent Imprinted Polymers Based on the Combination of Molecular Dynamics Simulation and Quantum Mechanics Calculations
by Xue Yu, Jiangyang Mo, Mengxia Yan, Jianhui Xin, Xuejun Cao, Jiawen Wu and Junfen Wan
Polymers 2024, 16(16), 2257; https://doi.org/10.3390/polym16162257 - 9 Aug 2024
Viewed by 289
Abstract
Molecular imprinting is a promising approach for developing polymeric materials as artificial receptors. However, only a few types of molecularly imprinted polymers (MIPs) are commercially available, and most research on MIPS is still in the experimental phase. The significant limitation has been [...] Read more.
Molecular imprinting is a promising approach for developing polymeric materials as artificial receptors. However, only a few types of molecularly imprinted polymers (MIPs) are commercially available, and most research on MIPS is still in the experimental phase. The significant limitation has been a challenge for screening imprinting systems, particularly for weak functional target molecules. Herein, a combined method of quantum mechanics (QM) computations and molecular dynamics (MD) simulations was employed to screen an appropriate 2,4-dichlorophenoxyacetic acid (2,4-D) imprinting system. QM calculations were performed using the Gaussian 09 software. MD simulations were conducted using the Gromacs2018.8 software suite. The QM computation results were consistent with those of the MD simulations. In the MD simulations, a realistic model of the ‘actual’ pre-polymerisation mixture was obtained by introducing numerous components in the simulations to thoroughly investigate all non-covalent interactions during imprinting. This study systematically examined MIP systems using computer simulations and established a theoretical prediction model for the affinity and selectivity of MIPs. The combined method of QM computations and MD simulations provides a robust foundation for the rational design of MIPs. Full article
(This article belongs to the Section Polymer Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic of the steps involved in a typical molecular dynamics simulation for a pre-polymerisation mixture illustrated for a system based on a functional monomer (AM) and 2,4-D in chloroform: (<b>A</b>) Components added to the design. (<b>B</b>) A 5 ns production phase and the pre-polymerisation mixture after equilibration. (<b>C</b>) Statistical analysis results of the different pre-polymerisation mixtures produced.</p>
Full article ">Figure 2
<p>Atoms studied using radial distribution functions (RDFs).</p>
Full article ">Figure 3
<p>Optimised geometry of complexes; the ab initio mechanical quantum computation was based on Density Function Theory<tt> (</tt>DFT) at the Becke 3-parameter-Lee-Yang-Parr (B3LYP) level with 6-311+G** basis set.</p>
Full article ">Figure 4
<p>MESP-mapped molecular vdW surfaces of the complexes.</p>
Full article ">Figure 5
<p>Configurations of the last frames of imprinting systems P1–P3 using DMSO as the solvent. (2,4-D molecules (red); 4-VP molecules (black); TFMAA molecules (green); AM molecules (cyan); DMSO molecules (tan)).</p>
Full article ">Figure 6
<p>RDFs displaying probabilities of obtaining the atomic densities of the functional monomers at various separation distances from the 2,4-D functional groups in the imprinting systems P1–P3.</p>
Full article ">Figure 7
<p>Conformations of the last frames of the imprinting systems P4–P8 with chloroform as the solvent. (DMAEMA molecules (black); 2,4-D molecules (red); 4-VP molecules (black); AM molecules (cyan); TFMAA molecules (green); MAA molecules (green); chloroform (grey)).</p>
Full article ">Figure 8
<p>RDFs displaying probabilities of obtaining the atomic densities of the functional monomers for various separation distances from the 2,4-D functional groups in the imprinting systems P4–P8.</p>
Full article ">Figure 9
<p>Number of H–H bonds produced between the functional monomers and 2,4-D in the imprinting systems P4–P8 with chloroform as the solvent.</p>
Full article ">Figure 10
<p>Conformations of the last frame of the imprinting systems P8–P10. (2,4-D molecules (red); AM molecules (cyan); chloroform molecules (grey)).</p>
Full article ">Figure 11
<p>Number of H–H bonds produced between functional monomers and 2,4-D in the imprinting systems P8–P10.</p>
Full article ">
19 pages, 9250 KiB  
Article
Multi-Agent Deep Reinforcement Learning Based Dynamic Task Offloading in a Device-to-Device Mobile-Edge Computing Network to Minimize Average Task Delay with Deadline Constraints
by Huaiwen He, Xiangdong Yang, Xin Mi, Hong Shen and Xuefeng Liao
Sensors 2024, 24(16), 5141; https://doi.org/10.3390/s24165141 - 8 Aug 2024
Viewed by 446
Abstract
Device-to-device (D2D) is a pivotal technology in the next generation of communication, allowing for direct task offloading between mobile devices (MDs) to improve the efficient utilization of idle resources. This paper proposes a novel algorithm for dynamic task offloading between the active MDs [...] Read more.
Device-to-device (D2D) is a pivotal technology in the next generation of communication, allowing for direct task offloading between mobile devices (MDs) to improve the efficient utilization of idle resources. This paper proposes a novel algorithm for dynamic task offloading between the active MDs and the idle MDs in a D2D–MEC (mobile edge computing) system by deploying multi-agent deep reinforcement learning (DRL) to minimize the long-term average delay of delay-sensitive tasks under deadline constraints. Our core innovation is a dynamic partitioning scheme for idle and active devices in the D2D–MEC system, accounting for stochastic task arrivals and multi-time-slot task execution, which has been insufficiently explored in the existing literature. We adopt a queue-based system to formulate a dynamic task offloading optimization problem. To address the challenges of large action space and the coupling of actions across time slots, we model the problem as a Markov decision process (MDP) and perform multi-agent DRL through multi-agent proximal policy optimization (MAPPO). We employ a centralized training with decentralized execution (CTDE) framework to enable each MD to make offloading decisions solely based on its local system state. Extensive simulations demonstrate the efficiency and fast convergence of our algorithm. In comparison to the existing sub-optimal results deploying single-agent DRL, our algorithm reduces the average task completion delay by 11.0% and the ratio of dropped tasks by 17.0%. Our proposed algorithm is particularly pertinent to sensor networks, where mobile devices equipped with sensors generate a substantial volume of data that requires timely processing to ensure quality of experience (QoE) and meet the service-level agreements (SLAs) of delay-sensitive applications. Full article
(This article belongs to the Section Communications)
Show Figures

Figure 1

Figure 1
<p>Architecture of D2D–MEC network.</p>
Full article ">Figure 2
<p>Architecture of multi-agent DRL algorithm.</p>
Full article ">Figure 3
<p>Performance evaluation under different batch sizes. (<b>a</b>) average delay; (<b>b</b>) ratio of dropped tasks.</p>
Full article ">Figure 4
<p>Performance evaluation under different learning rates. (<b>a</b>) average delay; (<b>b</b>) ratio of dropped tasks.</p>
Full article ">Figure 5
<p>Performance evaluation under different task generation probabilities. (<b>a</b>) average delay; (<b>b</b>) ratio of dropped tasks.</p>
Full article ">Figure 6
<p>Performance evaluation under different penalties. (<b>a</b>) average delay; (<b>b</b>) ratio of dropped tasks.</p>
Full article ">Figure 7
<p>Performance of different algorithms. (<b>a</b>) average delay; (<b>b</b>) ratio of dropped tasks.</p>
Full article ">Figure 8
<p>Performance of different algorithms across different edge CPU frequencies. (<b>a</b>) average delay; (<b>b</b>) ratio of dropped tasks.</p>
Full article ">Figure 9
<p>Performance of different algorithms across different task deadlines. (<b>a</b>) average delay; (<b>b</b>) ratio of dropped tasks.</p>
Full article ">
14 pages, 8541 KiB  
Article
Preparation, Thermal Behavior, and Conformational Stability of HMX/Cyclopentanone Cocrystallization
by Yuting Tao, Shaohua Jin, Tongbin Wang, Chongchong She, Kun Chen, Junfeng Wang and Lijie Li
Crystals 2024, 14(8), 711; https://doi.org/10.3390/cryst14080711 - 8 Aug 2024
Viewed by 287
Abstract
The cocrystallization of 1,3,5,7-tetranitro-1,3,5,7-tetrazolidine (HMX) with cyclopentanone was achieved via a controlled cooling method, followed by comprehensive characterization that confirmed the α-configuration of HMX within the cocrystal. The enthalpy of dissolution of HMX in cyclopentanone was assessed across a range of temperatures using [...] Read more.
The cocrystallization of 1,3,5,7-tetranitro-1,3,5,7-tetrazolidine (HMX) with cyclopentanone was achieved via a controlled cooling method, followed by comprehensive characterization that confirmed the α-configuration of HMX within the cocrystal. The enthalpy of dissolution of HMX in cyclopentanone was assessed across a range of temperatures using a C-80 Calvert microcalorimeter, revealing an endothermic dissolution process. Subsequently, the molar enthalpy of dissolution was determined, and kinetic equations describing the dissolution rate were derived for temperatures of 303.15, 308.15, 313.15, 318.15, and 323.15 K as follows: dα⁄dt = 10−2.46(1 − α)0.35, dα⁄dt = 10−2.19(1 − α)0.79, dα⁄dt = 10−1.76(1 − α)1.32, dα⁄dt = 10−1.86(1 − α)0.46, and dα⁄dt = 10−2.02(1 − α)0.70, respectively. Additionally, molecular dynamics (MD) simulations investigated the intermolecular interactions of the HMX/cyclopentanone cocrystallization process, demonstrating a transformation of HMX from β- to α-conformation within the cyclopentanone environment. Theoretical calculations performed at the ωB97XD/6-311G(d,p) level affirmed that α-HMX exhibited stronger binding affinity toward cyclopentanone compared to β-HMX, corroborating experimental findings. A comprehensive understanding of the dissolution behavior of HMX in cyclopentanone holds significant implications for crystal growth methodologies and cocrystallization processes. Such insights are pivotal for optimizing HMX dissolution processes and offer valuable perspectives for developing and designing advanced energetic materials. Full article
(This article belongs to the Special Issue Co-Crystals and Polymorphic Transition in Energetic Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Optical microscope images of (<b>a</b>) raw HMX and (<b>b</b>) HMX/cyclopentanone cocrystallization and (<b>c</b>,<b>d</b>) polarized microscope images of HMX/cyclopentanone cocrystallization.</p>
Full article ">Figure 2
<p>XRD patterns of pure β-HMX, α-HMX, raw-HMX, and HMX/cyclopentanone cocrystallization.</p>
Full article ">Figure 3
<p>FT−IR spectra of the raw materials and the cocrystallization.</p>
Full article ">Figure 4
<p>TG−DTG curves of HMX/cyclopentanone cocrystallization.</p>
Full article ">Figure 5
<p>DSC curves of pure β-HMX, α-HMX, and HMX/cyclopentanone cocrystallization.</p>
Full article ">Figure 6
<p>High-pressure DSC curves of pure β-HMX and HMX/cyclopentanone cocrystallization.</p>
Full article ">Figure 7
<p>The curve of heat−flow of HMX in cyclopentanone.</p>
Full article ">Figure 8
<p>The relationship of reaction rate constant (<span class="html-italic">k</span>) versus temperature (<span class="html-italic">T</span>) for the dissolution of HMX in cyclopentanone (the red line is the best−fit line).</p>
Full article ">Figure 9
<p>Screenshots of MD simulations of HMX in cyclopentanone at 313.15 K: (<b>a</b>) 0 ps; (<b>b</b>) 345 ps; (<b>c</b>) 666 ps; and (<b>d</b>) 670 ps. (HMX molecules are represented by a ball-and-stick, and cyclopentanone molecules are represented by a line.).</p>
Full article ">Figure 10
<p>α-HMX and cyclopentanone dimer structures optimized at the ωB97XD/6-311G(d,p) level ((<b>a</b>) α-HMX/α-HMX; (<b>b</b>) cyclopentanone/cyclopentanone; (<b>c</b>) α-HMX/cyclopentanone).</p>
Full article ">Figure 11
<p>(<b>a</b>) Final structures after equilibrated for by after molecular dynamics simulation; (<b>b</b>) Interaction of HMX with cyclopentanone (The dashed line indicates the distance between the O atom of the cyclopentanone carbonyl and the four methylene H atom of HMX in Å).</p>
Full article ">
16 pages, 3431 KiB  
Article
Radioiodinated Anastrozole and Epirubicin for HER2-Targeted Cancer Therapy: Molecular Docking and Dynamics Insights with Implications for Nuclear Imaging
by Mazen Abdulrahman Binmujlli
Processes 2024, 12(8), 1659; https://doi.org/10.3390/pr12081659 - 7 Aug 2024
Viewed by 329
Abstract
This study evaluates radioiodinated anastrozole ([125I]anastrozole) and epirubicin ([125I]epirubicin) for HER2-targeted cancer therapy, utilizing radiopharmaceutical therapy (RPT) for personalized treatment of HER2-positive cancers. Through molecular docking and dynamics simulations (200 ns), it investigates these compounds’ binding affinities and mechanisms [...] Read more.
This study evaluates radioiodinated anastrozole ([125I]anastrozole) and epirubicin ([125I]epirubicin) for HER2-targeted cancer therapy, utilizing radiopharmaceutical therapy (RPT) for personalized treatment of HER2-positive cancers. Through molecular docking and dynamics simulations (200 ns), it investigates these compounds’ binding affinities and mechanisms to the HER2 receptor compared to lapatinib, a known HER2 inhibitor. Molecular docking studies identified [125I]epirubicin with the highest ΔGbind (−10.92 kcal/mol) compared to lapatinib (−10.65 kcal/mol) and [125I]anastrozole (−9.65 kcal/mol). However, these differences were not statistically significant. Further molecular dynamics (MD) simulations are required to better understand the implications of these findings on the therapeutic potential of the compounds. MD simulations affirmed a stable interaction with the HER2 receptor, indicated by an average RMSD of 4.51 Å for [125I]epirubicin. RMSF analysis pointed to significant flexibility at key receptor regions, enhancing the inhibitory action against HER2. The [125I]epirubicin complex maintained an average of four H-bonds, indicating strong and stable interactions. The average Rg values for [125I]anastrozole and [125I]epirubicin complexes suggest a modest increase in structural flexibility without compromising protein compactness, reflecting their potential to induce necessary conformational changes in the HER2 receptor function. These analyses reveal enhanced flexibility and specific receptor region interactions, suggesting adaptability in binding, which could augment the inhibitory action against HER2. MM-PBSA calculations indicate the potential of these radioiodinated compounds as HER2 inhibitors. Notably, [125I]epirubicin exhibited a free binding energy of −65.81 ± 0.12 kJ/mol, which is comparable to lapatinib at −64.05 ± 0.11 kJ/mol and more favorable than [125I]anastrozole at −57.18 ± 0.12 kJ/mol. The results suggest electrostatic interactions as a major contributor to the binding affinity. The computational analysis underscores that [125I]anastrozole and [125I]epirubicin may have a promising role as HER2 inhibitors, especially [125I]epirubicin due to its high binding affinity and dynamic receptor interactions. These findings, supported by molecular docking scores and MM-PBSA binding energies, advocate for their potential superior inhibitory capability against the HER2 receptor. To validate these computational predictions and evaluate the therapeutic potential of these compounds for HER2-targeted cancer therapy, it is essential to conduct empirical validation through both in vitro and in vivo studies. Full article
Show Figures

Figure 1

Figure 1
<p>Depiction of (<b>a</b>) radioiodinated anastrozole ([<sup>125</sup>I]anastrozole), (<b>b</b>) radioiodinated epirubicin ([<sup>125</sup>I]epirubicin), and (<b>c</b>) lapatinib.</p>
Full article ">Figure 2
<p>Three-dimensional and two-dimensional interactions of lapatinib (<b>a</b>,<b>b</b>), [<sup>125</sup>I]anastrozole (<b>c</b>,<b>d</b>), and [<sup>125</sup>I]epirubicin (<b>e</b>,<b>f</b>) within the active binding site of the human HER2 receptor (PDB ID: 3RCD). Distances are given in angstroms (Å). Discovery Studio visualizer was used to generate these models.</p>
Full article ">Figure 3
<p>Root-mean-square deviation (RMSD) analysis for molecular dynamics (MD) simulation trajectories over a span of 200 ns. (<b>a</b>) RMSD plots of the HER2 protein backbone, reflecting molecular variations after interaction with [<sup>125</sup>I]anastrozole, [<sup>125</sup>I]epirubicin, lapatinib, and the co-crystallized ligand TAK-285. (<b>b</b>) The RMSD plots also reveal structural modifications of [<sup>125</sup>I]anastrozole, [<sup>125</sup>I]epirubicin, lapatinib, and the co-crystallized ligand TAK-285 into the active binding site HER2.</p>
Full article ">Figure 4
<p>RMSF (root-mean-square fluctuation) plots illustrating the behavior of backbone atoms in HER2 over a 200 ns MD simulation across all systems. The RMSF data show the fluctuations of individual protein residues as they interact with the ligands throughout the simulation.</p>
Full article ">Figure 5
<p>Radius of gyration (Rg) plots for HER2 backbone atoms across all systems during a 200 ns MD simulation.</p>
Full article ">Figure 6
<p>Hydrogen bond profiles for interactions of (<b>a</b>) HER2–TAK-285, (<b>b</b>) HER2–lapatinib, (<b>c</b>) HER2–[<sup>125</sup>I]anastrozole, and (<b>d</b>) HER2–[<sup>125</sup>I]epirubicin were derived from MD simulations spanning 0–200 ns.</p>
Full article ">
28 pages, 5588 KiB  
Article
Pharmacophore-Assisted Covalent Docking Identifies a Potential Covalent Inhibitor for Drug-Resistant Genotype 3 Variants of Hepatitis C Viral NS3/4A Serine Protease
by Kanzal Iman, Muhammad Usman Mirza, Fazila Sadia, Matheus Froeyen, John F. Trant and Safee Ullah Chaudhary
Viruses 2024, 16(8), 1250; https://doi.org/10.3390/v16081250 - 3 Aug 2024
Viewed by 1189
Abstract
The emergence of drug-resistance-inducing mutations in Hepatitis C virus (HCV) coupled with genotypic heterogeneity has made targeting NS3/4A serine protease difficult. In this work, we investigated the mutagenic variations in the binding pocket of Genotype 3 (G3) HCV NS3/4A and evaluated ligands for [...] Read more.
The emergence of drug-resistance-inducing mutations in Hepatitis C virus (HCV) coupled with genotypic heterogeneity has made targeting NS3/4A serine protease difficult. In this work, we investigated the mutagenic variations in the binding pocket of Genotype 3 (G3) HCV NS3/4A and evaluated ligands for efficacious inhibition. We report mutations at 14 positions within the ligand-binding residues of HCV NS3/4A, including H57R and S139P within the catalytic triad. We then modelled each mutational variant for pharmacophore-based virtual screening (PBVS) followed by covalent docking towards identifying a potential covalent inhibitor, i.e., cpd-217. The binding stability of cpd-217 was then supported by molecular dynamic simulation followed by MM/GBSA binding free energy calculation. The free energy decomposition analysis indicated that the resistant mutants alter the HCV NS3/4A–ligand interaction, resulting in unbalanced energy distribution within the binding site, leading to drug resistance. Cpd-217 was identified as interacting with all NS3/4A G3 variants with significant covalent docking scores. In conclusion, cpd-217 emerges as a potential inhibitor of HCV NS3/4A G3 variants that warrants further in vitro and in vivo studies. This study provides a theoretical foundation for drug design and development targeting HCV G3 NS3/4A. Full article
(This article belongs to the Special Issue Recent Advances in Anti-HCV, Anti-HBV and Anti-flavivirus Agents)
Show Figures

Figure 1

Figure 1
<p>Framework for in silico analysis of HCV NS3/4A G3. Flowchart summarising the computational framework for modelling and targeting HCV NS3/4A G3 towards HCV treatment.</p>
Full article ">Figure 2
<p>Multiple sequence alignment of HCV NS3/4A Genotype 3 variants. Amino acid substitutions at 14 positions are identified with the arrow in blue. The respective mutations at specified positions are identified in blue. The conserved residues are highlighted in red. Note: Variant_1 to variant_14 as G3.v1 to G3.v14.</p>
Full article ">Figure 3
<p>Chemical structure of CHEMBL569970 (cpd-217; PubChem45485999). The lead indicated binding potential with all HCV NS3/4A variants.</p>
Full article ">Figure 4
<p>Surface representation of the binding pocket of HCV NS3/4A G3 variants docked with CHEMBL569970 (cps-217; PubChem45485999). Key interacting residues including Thr42, Leu/Phe43, Val55, Gly58, Asp81, Leu135, Lys136, Gly137, Ser138, Ser139, Phe154, Arg155, Ala156, Ala157, Met485, Phe486, Gly525, Gln526, Asp527, and Asp528 are labelled. All HCV NS3/4A G3 variants 1 to 14 are highlighted in (<b>A</b>–<b>N</b>), respectively.</p>
Full article ">Figure 5
<p>Ligand interaction diagrams of CHEMBL569970 (PubChem45485999) with residues inside of the binding pocket of HCV NS3/4A G3 variants. Residues making hydrogen bond (H-bond) interactions include Gln41, Thr42, His/Arg57, Asp81, Leu135, Gly137, Ser139, Arg155, Ala157, Asp487, and Gln526. Asp81 is involved in making the salt bridge in G3.v3 and 13. Lys136 (G3.v10) and H57 (G3.v4) are involved in hydrophobic interactions, including Pi–cation and Pi–Pi stacking, respectively. All HCV NS3/4A G3 variants 1 to 14 are highlighted in (<b>A</b>–<b>N</b>), respectively.</p>
Full article ">Figure 6
<p>Mean Square Deviation (RMSD) plots of the backbone atoms (CA, N, C) of the HCV NS3/4A G3 variants and heavy atoms of cpd-217 relative to the initial structure over a 100 ns molecular dynamics simulation. The RMSD values are plotted as a function of time (ns) for each variant of compound 217: (<b>A</b>) cpd217/G3.v1, (<b>B</b>) cpd217/G3.v2, (<b>C</b>) cpd217/G3.v3, (<b>D</b>) cpd217/G3.v4, (<b>E</b>) cpd217/G3.v5, (F) cpd217/G3.v6, (<b>G</b>) cpd217/G3.v7, (<b>H</b>) cpd217/G3.v8, (<b>I</b>) cpd217/G3.v9, (<b>J</b>) cpd217/G3.v10, (<b>K</b>) cpd217/G3.v11, (<b>L</b>) cpd217/G3.v12, (<b>M</b>) cpd217/G3.v13, (<b>N</b>) cpd217/G3.v14. Each plot illustrates the structural stability and conformational changes of the compound-protease complexes over time, with the RMSD measured in Ångströms (Å).</p>
Full article ">Figure 7
<p>Root mean square deviation of HCV NS3/4A G3 wildtype (WT) with all G3 variants. The MD-simulated HCV NS3/4A G3.WT (yellow) is superimposed on G3 variants (cornflower blue), G3.v1–v7 in (<b>A</b>) and G3.v8–v14 in (<b>B</b>), after 100 ns. The NS4A is coloured orange red and catalytic triad residues are represented by a ball and stick representation (yellow). The RMSF plots of HCV NS3/4A G3.WT (black line) with G3.v1–v7 in (<b>C</b>) and G3.v8–v14 in (<b>D</b>) are displayed. The residue numbers are along the x-axis and fluctuations in Å are along y-axis, while catalytic triad residues are highlighted in red.</p>
Full article ">Figure 8
<p>Per-residue decomposition analysis of the potential binding site residues of HCV NS3/4A G3 variants in the presence of cpd-217. The colour codes are represented for each variant, while the wildtype (PDB ID: 4A92) is coloured black. The mutated residues in corresponding variants are highlighted in red. The values are measured in kcal/mol.</p>
Full article ">
25 pages, 3629 KiB  
Article
Unraveling Protein-Metabolite Interactions in Precision Nutrition: A Case Study of Blueberry-Derived Metabolites Using Advanced Computational Methods
by Dipendra Bhandari, Kiran Kumar Adepu, Andriy Anishkin, Colin D. Kay, Erin E. Young, Kyle M. Baumbauer, Anuradha Ghosh and Sree V. Chintapalli
Metabolites 2024, 14(8), 430; https://doi.org/10.3390/metabo14080430 - 3 Aug 2024
Viewed by 473
Abstract
Metabolomics, the study of small-molecule metabolites within biological systems, has become a potent instrument for understanding cellular processes. Despite its profound insights into health, disease, and drug development, identifying the protein partners for metabolites, especially dietary phytochemicals, remains challenging. In the present study, [...] Read more.
Metabolomics, the study of small-molecule metabolites within biological systems, has become a potent instrument for understanding cellular processes. Despite its profound insights into health, disease, and drug development, identifying the protein partners for metabolites, especially dietary phytochemicals, remains challenging. In the present study, we introduced an innovative in silico, structure-based target prediction approach to efficiently predict protein targets for metabolites. We analyzed 27 blood serum metabolites from nutrition intervention studies’ blueberry-rich diets, known for their health benefits, yet with elusive mechanisms of action. Our findings reveal that blueberry-derived metabolites predominantly interact with Carbonic Anhydrase (CA) family proteins, which are crucial in acid-base regulation, respiration, fluid balance, bone metabolism, neurotransmission, and specific aspects of cellular metabolism. Molecular docking showed that these metabolites bind to a common pocket on CA proteins, with binding energies ranging from −5.0 kcal/mol to −9.0 kcal/mol. Further molecular dynamics (MD) simulations confirmed the stable binding of metabolites near the Zn binding site, consistent with known compound interactions. These results highlight the potential health benefits of blueberry metabolites through interaction with CA proteins. Full article
(This article belongs to the Section Bioinformatics and Data Analysis)
Show Figures

Figure 1

Figure 1
<p>Computational workflow to determine the target protein from metabolites obtained from metabolomics. The metabolites undergo an intricate computational workflow pipeline aimed at identifying protein targets and elucidating their functional pathways. The identified target proteins and associated metabolites are then subjected to in silico methods such as molecular dynamic simulations and require further experimental validation.</p>
Full article ">Figure 2
<p>Clustering diagram of structurally similar compounds from diet-derived metabolites in human subjects following a blueberry diet intervention. (<b>A</b>) The plot depicts the relationship between cluster size and the number of clusters. Cluster-0 comprises 15 metabolite compounds, while Cluster-1 encompasses 12 metabolite compounds. (<b>B</b>) The plot illustrates the Silhouette score versus the number of clusters (K), with a Silhouette score of 0.9989 ± 0.0028 achieved for K = 2. A Silhouette score approaching 1 indicates effective clustering.</p>
Full article ">Figure 3
<p>The metabolic pathway interaction network of Cluster-0 metabolite (4-hydroxybenzoic acid). (<b>A</b>) The KEGG pathway for 4-hydroxybenzoic acid involvement in oxidative phosphorylation is distinctly highlighted in a pink color. (<b>B</b>) In the STITCH interaction network for 4-hydroxybenzoic acid, Coenzyme Q2 (COQ2) and Coenzyme Q6 (COQ6), along with CA II, emerge as the nearest and most robust interaction partners, depicted by prominent green lines.</p>
Full article ">Figure 4
<p>Molecular docking of four different metabolites with CA II protein. The docked state of adrenaline (in red), indisulam (in purple), histamine (in yellow), and Zn (depicted as an orange sphere) with CA II is illustrated. Adrenaline, indisulam, and histamine, recognized as experimentally proven compounds binding to CA II, are positioned within the binding site near the Zn (yellow sphere) in CA II, surrounded by histidine residues (in silver).</p>
Full article ">Figure 5
<p>AutoDock Vina predicted binding sites for selected Cluster-0 metabolites. The docking sites for metabolites are depicted as follows: (<b>A</b>) Hippuric acid (HA) is represented in green. (<b>B</b>) 3-hydroxyhippuric acid (HHPA) is highlighted in pink. (<b>C</b>) 3,4-dihydroxycinnamic acid is shown in yellow. These metabolites are positioned near the Zn (depicted as a silver-colored sphere and highlighted within the dark circle) binding cavity of CA II, surrounded by histidine residues (HIS64, HIS94, HIS96, and PHE131). Notably, these metabolites share the identical binding site with experimentally bound compounds.</p>
Full article ">Figure 6
<p>Predicted binding site of a single metabolite with all the CA family proteins. The docked state of 3-hydroxyhippuric acid, represented in green and depicted in licorice form, is illustrated near the Zn (depicted as a silver-colored sphere) binding site across various CA family proteins. The binding site of 3-hydroxyhippuric acid is consistently highlighted within the red circle for clarity. (<b>A</b>) CA II, (<b>B</b>) CA III, (<b>C</b>) CA IV, (<b>D</b>) CA V, (<b>E</b>) CA VI, (<b>F</b>) CA VII, (<b>G</b>) CA IX, (<b>H</b>) CA XII, (<b>I</b>) CA XIII, (<b>J</b>) CA XIV. Notably, 3-hydroxyhippuric acid occupies a similar binding site across all CA family proteins.</p>
Full article ">Figure 7
<p>Root Mean Squared Deviation (RMSD) of selected metabolites’ complex with the CA family proteins. RMSD plot indicates a stable interaction for both the protein–ligand complexes (<b>A</b>) CA I and 3, 4-Dihydroxycinnamic acid (DHCA), (<b>B</b>) CA II and 3-hydroxyhippuric acid (HHPA).</p>
Full article ">Figure 8
<p>Interaction details of selected metabolites with CA family proteins after molecular dynamics simulations. (<b>A</b>) CA II and 3-hydroxyhippuric acid (HHPA), (<b>B</b>) CA VII and 3, 4-Dihydroxycinnamic acid (DHCA). Hydrogen bond is represented by dotted lines expressed in Å. HHPA and DHCA are highlighted in a dark circle.</p>
Full article ">
24 pages, 12834 KiB  
Article
Prevention of Blood Incompatibility Related Hemagglutination: Blocking of Antigen A on Red Blood Cells Using In Silico Designed Recombinant Anti-A scFv
by Saleha Hafeez and Najam Us Sahar Sadaf Zaidi
Antibodies 2024, 13(3), 64; https://doi.org/10.3390/antib13030064 - 1 Aug 2024
Viewed by 701
Abstract
Critical blood shortages plague healthcare systems, particularly in lower-income and middle-income countries. This affects patients requiring regular transfusions and creates challenges during emergencies where universal blood is vital. To address these shortages and support blood banks during emergencies, this study reports a method [...] Read more.
Critical blood shortages plague healthcare systems, particularly in lower-income and middle-income countries. This affects patients requiring regular transfusions and creates challenges during emergencies where universal blood is vital. To address these shortages and support blood banks during emergencies, this study reports a method for increasing the compatibility of blood group A red blood cells (RBCs) by blocking surface antigen-A using anti-A single chain fragment variable (scFv). To enhance stability, the scFv was first modified with the addition of interdomain disulfide bonds. The most effective location for this modification was found to be H44-L232 of mutant-1a scFv. ScFv was then produced from E.coli BL21(DE3) and purified using a three-step process. Purified scFvs were then used to block maximum number of antigens-A on RBCs, and it was found that only monomers were functional, while dimers formed through incorrect domain-swapping were non-functional. These antigen-blocked RBCs displayed no clumping in hemagglutination testing with incompatible blood plasma. The dissociation constant KD was found to be 0.724 μM. Antigen-blocked RBCs have the potential to be given to other blood groups during emergencies. This innovative approach could significantly increase the pool of usable blood, potentially saving countless lives. Full article
(This article belongs to the Section Antibody Discovery and Engineering)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the design of mutant–1a anti–A scFv gene.</p>
Full article ">Figure 2
<p>(<b>A</b>) I-TASSER models of mutants-1a (interface 1) and 1b (interface 2) scFvs. Information about the models is provided in <a href="#app2-antibodies-13-00064" class="html-app">Appendix A</a> <a href="#antibodies-13-00064-f0A5" class="html-fig">Figure A5</a> (<b>B</b>) RMSF plot of amino acids. Double arrow represents the interacting amino acid of GS linker, black, red, and blue colors represent native, mutant-1a and mutant-1b scFvs respectively. (<b>C</b>) Distance between Cα-atoms of amino acids TRP-47 and THR-229 at interface 1 and (<b>D</b>) distance between Cα-atoms of amino acids ALA-106 and LEU-178 at interface 2.</p>
Full article ">Figure 3
<p>Snapshots of MD trajectories. Green, red, and blue colors represent amino acids of native, mutant-1a and mutant-1b scFvs respectively. Yellow color represents the binding center and red arrows represent the bond formed between amino acids. (<b>A</b>) Bond formed between TRP-228 of CDR-3 of VL and SER-131 of GS linker in native scFv. (<b>B</b>) Bond formed between PRO-227 (Grey color) of CDR-3 of VL and SER-131 of GS linker in mutant-1b scFv. (<b>C</b>) Snapshot showing the initial positions of interacting amino acids of CDRs at 0 ns. (<b>D</b>) Snapshot showing the final positions of interacting amino acids of CDRs at 200 ns.</p>
Full article ">Figure 4
<p>(<b>A</b>) RMSD vs. time plot of each scFv obtained along 200 ns simulation. Number of hydrogen bonds formed between antigen-A and scFv. (<b>B</b>) Native scFv, (<b>C</b>) Mutant-1a scFv and (<b>D</b>) Mutant-1b scFv. High resolution images are shown in <a href="#app2-antibodies-13-00064" class="html-app">Appendix A</a> <a href="#antibodies-13-00064-f0A6" class="html-fig">Figure A6</a>A–D.</p>
Full article ">Figure 5
<p>(<b>A</b>) Purification of mutant-1a anti-A scFv through nickel affinity chromatography. Coomassie R-250 stained 15% SDS PAGE gel showing purification of recombinant protein before (lanes 2–6) and after elution (lanes 7–9). EfA1 shows 4 bands of ~25 kDa, 50 kDa, 75 kDa and +100 kDa. (<b>B</b>) Separation of recombinant multimers on the basis of molecular weight through size exclusion chromatography. (<b>C</b>) Separation of monomers and dimers through centrifugal filters (30 kDa MWCO). Lane 3 shows the purified product (monomers and dimers) of size exclusion chromatography. Lanes 4 and 5 shows the recombinant protein recovered as filtrate (F) and concentrate (C) respectively.</p>
Full article ">Figure 6
<p>Graphs (<b>A</b>,<b>B</b>) show the gradual decrease in the concentrations of mutant-1a anti-A scFv in the supernatant containing A+ and AB+ RBCs. Each aliquot represents a portion of a larger sample with the same protein concentration for analysis. Each aliquot contains 0.73 mg/L and 0.39 mg/L of SEC purified monomeric and dimeric anti-A scFv respectively.</p>
Full article ">Figure 7
<p>(<b>A</b>) Coomassie R-250 stained 15% SDS PAGE gel shows decrease in concentration of scFv monomer (25 kDa) from 0 to 30 min (lanes 2–7 in the case of A+ and lanes 9–14 in the case of AB+). (<b>B</b>) Western blot showing all histidine tagged proteins: Ni-NTA purified proteins (lane 2), size exclusion chromatography separated proteins (lane 3), scFvs attached on A+ and AB+ RBCs (lanes 4 and 6) and scFvs in supernatant of both blood groups respectively (lanes 5 and 7).</p>
Full article ">Figure 8
<p>Blood incompatibility related hemagglutination reactions. Wells 1 and 7 show controls (A+ RBCs in anti-A IgM antibodies and B+ RBCs coated with anti-A scFv in anti-B IgM antibodies). Wells 2, 3 and 8 show antigen-A blocked A+ RBCs in anti-A IgM, anti-Rh (IgG and IgM), and B+ blood plasma containing anti-A antibodies respectively. Wells 4, 5, 10 and 11 show antigen-A blocked AB+ RBCs in anti-A IgM, anti-Rh (IgG and IgM), B+ blood plasma containing anti-A antibodies and anti-B IgM antibodies respectively. Wells 6, 9 and 12 are empty wells.</p>
Full article ">Figure 9
<p>(<b>A</b>) Coomassie R-250 stained 15% SDS PAGE gel shows purification of functional anti-A scFvs attached on RBCs by Ni-NTA chromatography. Lane 3 shows SEC purified product and lane 4 (fraction F1) shows final eluted fraction from Ni-NTA chromatography. (<b>B</b>) Absorbance vs. Log concentration (μM) graph generated by GraphPad Prism shows dissociation constant K<sub>D</sub> of monomeric functional anti-A scFv determined with a nonlinear regression model. Naked RBCs were used as negative control. Dilutions used ranged from 1 nM to 500 μM.</p>
Full article ">Figure A1
<p>Concept of blocking of blood group antigens using scFv.</p>
Full article ">Figure A2
<p>Structure of single chain fragment variable (scFv) showing two interfaces between heavy and light chain [<a href="#B14-antibodies-13-00064" class="html-bibr">14</a>].</p>
Full article ">Figure A3
<p>Structure of single chain fragment variable (scFv) showing ideal site of disulfide bond [<a href="#B14-antibodies-13-00064" class="html-bibr">14</a>].</p>
Full article ">Figure A4
<p>Formula for measuring the concentrations of each multimer (scFv)x in a solution.</p>
Full article ">Figure A5
<p>I-TASSER models of Anti-A scFvs. (<b>A</b>) Native anti-A scFv, (<b>B</b>) Mutant-1a anti-A scFv and (<b>C</b>) Mutant-1b anti-A scFv.</p>
Full article ">Figure A6
<p>(<b>A</b>) Refined image of Root Mean Square Deviation (RMSD) vs. Time plot graph. (<b>B</b>) Refined image of number of Hydrogen bonds formed between antigen and native anti-A scFv. (<b>C</b>) Refined image of number of Hydrogen bonds formed between antigen and mutant-1a anti-A scFv. (<b>D</b>) Refined image of number of Hydrogen bonds formed between antigen and mutant-1b anti-A scFv.</p>
Full article ">Figure A6 Cont.
<p>(<b>A</b>) Refined image of Root Mean Square Deviation (RMSD) vs. Time plot graph. (<b>B</b>) Refined image of number of Hydrogen bonds formed between antigen and native anti-A scFv. (<b>C</b>) Refined image of number of Hydrogen bonds formed between antigen and mutant-1a anti-A scFv. (<b>D</b>) Refined image of number of Hydrogen bonds formed between antigen and mutant-1b anti-A scFv.</p>
Full article ">Figure A7
<p>(<b>A</b>) Graph showing purification of recombinant protein through Ni-NTA chromatography. (<b>B</b>) Graph showing separation of scFv on the basis of sizes through size exclusion chromatography.</p>
Full article ">Figure A8
<p>Graph shows the absence of depletion of anti-A scFv in the supernatant after incubation with B+ RBCs.</p>
Full article ">Figure A9
<p>(<b>A</b>) Dimers due to flexible GS linker can adopt various shapes. Flexibility allowed dimers to pass through (<b>B</b>) Sephadex G-100 agarose bead and (<b>C</b>) 30 kDa MWCO membrane.</p>
Full article ">Figure A10
<p>I-TASSER generated models of dimers made from monomers of anti-A scFvs. Red color in models represents the complementary determining regions CDRs. Red and blue lines represent GS linker and disulfide bonds respectively. VH and VL are heavy and light chains respectively. (<b>A</b>) With interdomain disulfide bonds. (<b>B</b>) Without interdomain disulfide bonds.</p>
Full article ">
19 pages, 5381 KiB  
Article
The Inferential Binding Sites of GCGR for Small Molecules Using Protein Dynamic Conformations and Crystal Structures
by Mengru Wang, Xulei Fu, Limin Du, Fan Shi, Zichong Huang and Linlin Yang
Int. J. Mol. Sci. 2024, 25(15), 8389; https://doi.org/10.3390/ijms25158389 - 1 Aug 2024
Viewed by 287
Abstract
Glucagon receptor (GCGR) is a class B1 G-protein-coupled receptor that plays a crucial role in maintaining human blood glucose homeostasis and is a significant target for the treatment of type 2 diabetes mellitus (T2DM). Currently, six small molecules (Bay 27-9955, MK-0893, MK-3577, LY2409021, [...] Read more.
Glucagon receptor (GCGR) is a class B1 G-protein-coupled receptor that plays a crucial role in maintaining human blood glucose homeostasis and is a significant target for the treatment of type 2 diabetes mellitus (T2DM). Currently, six small molecules (Bay 27-9955, MK-0893, MK-3577, LY2409021, PF-06291874, and LGD-6972) have been tested or are undergoing clinical trials, but only the binding site of MK-0893 has been resolved. To predict binding sites for other small molecules, we utilized both the crystal structure of the GCGR and MK-0893 complex and dynamic conformations. We docked five small molecules and selected the best conformation based on binding mode, docking score, and binding free energy. We performed MD simulations to verify the binding mode of the selected small molecules. Moreover, when selecting conformations, results of competitive binding were referred to. MD simulation indicated that Bay 27-9955 exhibits moderate binding stability in Pocket 3. MK-3577, LY2409021, and PF-06291874 exhibited highly stable binding to Pocket 2, consistent with experimental results. However, LY2409021 may also bind to Pocket 5. Additionally, LGD-6972 exhibited relatively stable binding in Pocket 5. We also conducted structural modifications of LGD-6972 based on the results of MD simulations and predicted its analogues’ bioavailability, providing a reference for the study of GCGR small molecules. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Potential binding pockets and known small molecules of GCGR. (<b>A</b>) Orthosteric pocket obtained from GCGR apo state MD simulations. (<b>B</b>) Binding site of the small molecule MK-0893. (<b>C</b>) MDpocket predicted four allosteric pockets from GCGR/glucagon trajectories. The endogenous ligand glucagon is represented by a green cartoon. Each allosteric site is shown by a different colored surface. (<b>D</b>) Compounds with GCGR antagonistic activity that have been reported to be in or have been discontinued in clinical trials. The same groups are represented by pink ellipses. The compound currently undergoing clinical trials is represented by red dashed boxes.</p>
Full article ">Figure 2
<p>MK-0893 was used to certify the accuracy of the molecular docking and MD simulation methods. (<b>A</b>,<b>B</b>) The comparison of the representative conformation of the MK-0893 docking (light blue) and MD simulations (purple) with the crystal structure (yellow). Hydrogen bonds are indicated by dashed red lines. (<b>C</b>,<b>D</b>) RMSD frequency distribution and hydrogen bond frequency profile of small molecule MK-0893 during the last 20 ns of MD simulation (n = 3). Data are shown as mean ± SEM.</p>
Full article ">Figure 3
<p>The binding stability of Bay 27-9955 at Pocket 3 revealed in MD simulations. (<b>A</b>,<b>B</b>) Two representative conformations of Bay 27-9955 (cyan) in MD simulations. (<b>C</b>) Frequency curve of Bay 27-9955 RMSD values. (<b>D</b>,<b>E</b>) The binding free energy of key residues around the pocket as well as the energy decomposition of individual residues. The three trajectories of MD simulation are denoted by ‘x’, ‘*’, and ‘o’, respectively. ΔE<sub>vdw</sub>: van der Waals energy. ΔE<sub>ele</sub>: Electrostatic energy. ΔG<sub>polar</sub>: Polar solvation energy. ΔG<sub>bind</sub>: Total binding free energy. All values are in kcal/mol. All data in this figure are analyzed from the last 20 ns of the MD trajectory.</p>
Full article ">Figure 4
<p>The binding stability of MK-3577 at Pocket 2 revealed in MD simulations. (<b>A</b>) The representative conformation of MK-3577 (magenta) in MD simulations. (<b>B</b>) Hydrogen bond interactions between MK-3577 and residues over time. (<b>C</b>) Frequency curve of MK-3577 RMSD values. (<b>D</b>) The binding free energy of key residues around the pocket. The three trajectories of MD simulation are denoted by ‘x’, ‘*’, and ‘o’, respectively. All values are in kcal/mol. All data in this figure are analyzed from the last 20 ns of the MD trajectory.</p>
Full article ">Figure 5
<p>The binding stability of LY2409021 at Pocket 2 revealed in MD simulations. (<b>A</b>,<b>B</b>) Two representative conformations of LY2409021 (deep teal) in MD simulations. (<b>C</b>) Frequency curve of LY2409021 RMSD values. (<b>D</b>) Hydrogen bond interactions between LY2409021 and residues over time. (<b>E</b>,<b>F</b>) The binding free energy of key residues around the pocket as well as the energy decomposition of individual residues. The three trajectories of MD simulation are denoted by ‘x’, ‘*’, and ‘o’, respectively. ΔE<sub>vdw</sub>: van der Waals energy. ΔE<sub>ele</sub>: Electrostatic energy. ΔG<sub>polar</sub>: Polar solvation energy. ΔG<sub>bind</sub>: Total binding free energy. All values are in kcal/mol. All data in this figure are analyzed from the last 20 ns of the MD trajectory.</p>
Full article ">Figure 6
<p>The binding stability of LY2409021 at Pocket 5 revealed in MD simulations. (<b>A</b>,<b>B</b>) Two representative conformations of LY2409021 (deepteal) in MD simulations. (<b>C</b>) Frequency curve of LY2409021 RMSD values. (<b>D</b>) Hydrogen bond interactions between LY2409021 and residues over time. (<b>E</b>) The binding free energy of key residues around the pocket. The three trajectories of MD simulation are denoted by ‘x’, ‘*’, and ‘o’, respectively. All values are in kcal/mol. All data in this figure are analyzed from the last 20 ns of the MD trajectory.</p>
Full article ">Figure 7
<p>The binding stability of PF-06291874 at Pocket 5 revealed in MD simulations. (<b>A</b>,<b>B</b>) Two representative conformations of PF-06291874 (orange) in MD simulations. (<b>C</b>) Frequency curve of PF-06291874 RMSD values. (<b>D</b>) Hydrogen bond interactions between PF-06291874 and residues over time. (<b>E</b>) The binding free energy of key residues around the pocket. The three trajectories of MD simulation are denoted by ‘x’, ‘*’, and ‘o’, respectively. All values are in kcal/mol. All data in this figure are analyzed from the last 20 ns of the MD trajectory.</p>
Full article ">Figure 8
<p>The binding stability of LGD-6972 at Pocket 2 revealed in MD simulations. (<b>A</b>,<b>B</b>) Two representative conformations of LGD-6972 (marine) in MD simulations. (<b>C</b>) Frequency curve of LGD-6972 RMSD values. (<b>D</b>) Hydrogen bond interactions between LGD-6972 and residues over time. (<b>E</b>,<b>F</b>) The binding free energy of key residues around the pocket as well as the energy decomposition of individual residues. The three trajectories of MD simulation are denoted by ‘x’, ‘*’, and ‘o’, respectively. ΔE<sub>vdw</sub>: van der Waals energy. ΔE<sub>ele</sub>: Electrostatic energy. ΔG<sub>polar</sub>: Polar solvation energy. ΔG<sub>bind</sub>: Total binding free energy. All values are in kcal/mol. All data in this figure are analyzed from the last 20 ns of the MD trajectory.</p>
Full article ">Figure 9
<p>Modified binding modes and 2D structures based on the binding modes of LGD-1 (<b>A</b>) and LGD-2/3 (<b>B</b>). The modified groups are indicated by blue ellipses.</p>
Full article ">
15 pages, 5169 KiB  
Article
Aluminium Nitride Surface Characterization by Grinding with Laser–Ultrasonic Coupling
by He Zhang, Cong Sun, Yuan Hong, Yansheng Deng and Liang Ma
Materials 2024, 17(15), 3772; https://doi.org/10.3390/ma17153772 - 1 Aug 2024
Viewed by 299
Abstract
Aluminium nitride (AlN) materials are widely used in heat-dissipation substrates and electronic device packages. However, the application of aluminium nitride ceramics is hindered by the obvious anisotropy and high brittleness of its crystals, leading to poor material surface integrity and high grinding force. [...] Read more.
Aluminium nitride (AlN) materials are widely used in heat-dissipation substrates and electronic device packages. However, the application of aluminium nitride ceramics is hindered by the obvious anisotropy and high brittleness of its crystals, leading to poor material surface integrity and high grinding force. With the rapid development of microelectronics, the requirements for the material’s dimensional accuracy, machining efficiency, and surface accuracy are increasing. Therefore, a new machining process is proposed, combining laser and ultrasonic vibration with grinding. The laser–ultrasonic-assisted grinding (LUAG) of aluminium nitride is simulated by molecular dynamics (MD). Meanwhile, the effects of different processing techniques on grinding force, stress distribution, matrix damage mechanism, and subsurface damage depth are systematically investigated and verified by experiments. The results show that laser–ultrasonic-assisted grinding produces 50% lower grinding forces compared to traditional grinding (TG). The microhardness of AlN can reach more than 1200 HV, and the coefficient of friction and wear is reduced by 42.6%. The dislocation lines of the AlN substrate under this process are short but interlaced, making the material prone to phase transformation. Moreover, the subsurface damage depth is low, realising the substrate’s material hardening and wear resistance. These studies not only enhance the comprehension of material build-up and stress damage under the synergistic impact of laser, ultrasonic, and abrasive processing but also indicate that the proposed method can facilitate and realise high-performance machining of aluminium nitride substrate surfaces. Full article
Show Figures

Figure 1

Figure 1
<p>Three-dimensional MD simulation model.</p>
Full article ">Figure 2
<p>The RDF of Al–N, Al–Al, and N–N bonds of AlN via the four processing technologies.</p>
Full article ">Figure 3
<p>The grinding force and the amount of growth via the four processing technologies. (<b>a</b>) The grinding force (<b>b</b>) The grinding force.</p>
Full article ">Figure 4
<p>The atom flow field via the four processing technologies. (<b>a</b>) TG (<b>b</b>) LAG (<b>c</b>) UVAG (<b>d</b>) LUAG.</p>
Full article ">Figure 5
<p>The von Mises shear stress via the four processing technologies. (<b>a1</b>–<b>a2</b>) TG (<b>b1</b>–<b>b2</b>) LAG (<b>c1</b>–<b>c2</b>) UVAG (<b>d1</b>–<b>d2</b>) LUAG.</p>
Full article ">Figure 6
<p>The subsurface damage depth via the four processing technologies. (<b>a</b>) TG (<b>b</b>) LAG (<b>c</b>) UVAG (<b>d</b>) LUAG.</p>
Full article ">Figure 7
<p>The surface mechanical properties via the four processing technologies. (<b>a</b>) TG (<b>b</b>) LAG (<b>c</b>) UVAG (<b>d</b>) LUAG.</p>
Full article ">Figure 8
<p>LUAG experimental platform.</p>
Full article ">Figure 9
<p>Microhardness and surface roughness of the machining area.</p>
Full article ">Figure 10
<p>Measurement results of dynamic grinding force.</p>
Full article ">Figure 11
<p>Statistics of friction and wear test results.</p>
Full article ">Figure 12
<p>X-ray diffraction (XRD) spectrum of AlN ceramics. (<b>a</b>) Raw surface (<b>b</b>) TG surface (<b>c</b>) LAG surface (<b>d</b>) UVAG surface (<b>e</b>) LUAG surface.</p>
Full article ">
17 pages, 3847 KiB  
Article
Molecular Dynamics Simulations of Effects of Geometric Parameters and Temperature on Mechanical Properties of Single-Walled Carbon Nanotubes
by Lida Najmi and Zhong Hu
J. Compos. Sci. 2024, 8(8), 293; https://doi.org/10.3390/jcs8080293 - 30 Jul 2024
Viewed by 410
Abstract
Carbon nanotubes (CNTs) are considered an advanced form of carbon. They have superior characteristics in terms of mechanical and thermal properties compared to other available fibers and can be used in various applications, such as supercapacitors, sensors, and artificial muscles. The properties of [...] Read more.
Carbon nanotubes (CNTs) are considered an advanced form of carbon. They have superior characteristics in terms of mechanical and thermal properties compared to other available fibers and can be used in various applications, such as supercapacitors, sensors, and artificial muscles. The properties of single-walled carbon nanotubes (SWNTs) are significantly affected by geometric parameters such as chirality and aspect ratio, and testing conditions such as temperature and strain rate. In this study, the effects of geometric parameters and temperature on the mechanical properties of SWNTs were studied by molecular dynamics (MD) simulations using the Large-scaled Atomic/Molecular Massively Parallel Simulator (LAMMPS). Based on the second-generation reactive empirical bond order (REBO) potential, SWNTs of different diameters were tested in tension and compression under different strain rates and temperatures to understand their effects on the mechanical behavior of SWNTs. It was observed that the Young’s modulus and the tensile strength decreases with increasing SWNT tube diameter. As the chiral angle increases, the tensile strength increases, while the Young’s modulus decreases. The simulations were repeated at different temperatures of 300 K, 900 K, 1500 K, 2100 K and different strain rates of 1 × 10−3/ps, 0.75 × 10−3/ps, 0.5 × 10−3/ps, and 0.25 × 10−3/ps to investigate the effects of temperature and strain rate, respectively. The results show that the ultimate tensile strength of SWNTs increases with increasing strain rate. It is also seen that when SWNTs were stretched at higher temperatures, they failed at lower stresses and strains. The compressive behavior results indicate that SWNTs tend to buckle under lower stresses and strains than those under tensile stress. The simulation results were validated by and consistent with previous studies. The presented approach can be applied to investigate the properties of other advanced materials. Full article
(This article belongs to the Special Issue Theoretical and Computational Investigation on Composite Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic model and boundary condition setup for tensile or compressive testing.</p>
Full article ">Figure 2
<p>Tensile behavior and failure modes of selected SWNTs with a diameter of approximately 0.75 nm. (<b>A</b>): initial SWNT configuration before displacement; (<b>B</b>): SWNT configuration when necking occurs; (<b>C</b>): SWNT configuration after failure.</p>
Full article ">Figure 3
<p>The stress–strain behavior of zigzag SWNTs under tensile with a diameter of ~0.75 nm, a nanotube length of 5 nm, and strain rate of 0.001/ps at different temperatures.</p>
Full article ">Figure 4
<p>Tensile strength of armchair SWNTs with a nanotube length of 5 nm at different diameters and temperatures.</p>
Full article ">Figure 5
<p>Tensile strength of zigzag SWNTs with a nanotube length of 5 nm at different diameters and temperatures.</p>
Full article ">Figure 6
<p>Young’s modulus of armchair SWNTs with a nanotube length of 5 nm at different diameters and temperatures.</p>
Full article ">Figure 7
<p>Young’s modulus of zigzag SWNTs with a nanotube length of 5 nm at different diameters and temperatures.</p>
Full article ">Figure 8
<p>Tensile strength of SWNTs with a tube length of 5 nm and a diameter of 0.75 nm at different chirality and temperatures.</p>
Full article ">Figure 9
<p>Young’s modulus of SWNTs with a tube length of 5 nm and a diameter of 0.75 nm at different chirality and temperatures.</p>
Full article ">Figure 10
<p>Tensile strength of zigzag SWNTs with a tube length of 5 nm and a diameter of 0.75 nm at different strain rates and temperatures.</p>
Full article ">Figure 11
<p>Compressive failure modes of zigzag SWNTs with a diameter of approximately 0.75 nm and a length of 5 nm. (<b>A</b>): initial configuration; (<b>B</b>): configuration when buckling occurs; (<b>C</b>): configuration after failure.</p>
Full article ">Figure 12
<p>Compressive stress–strain behavior of z SWNTs with a diameter of 0.75 nm and a nanotube length of 5 nm at different temperatures.</p>
Full article ">
13 pages, 5074 KiB  
Article
Docking, MD Simulations, and DFT Calculations: Assessing W254’s Function and Sartan Binding in Furin
by Nikitas Georgiou, Thomas Mavromoustakos and Demeter Tzeli
Curr. Issues Mol. Biol. 2024, 46(8), 8226-8238; https://doi.org/10.3390/cimb46080486 - 30 Jul 2024
Viewed by 255
Abstract
Furins are serine endoproteases that are involved in many biological processes, where they play important roles in normal metabolism, in the activation of various pathogens, while they are a target for therapeutic intervention. Dichlorophenyl-pyridine “BOS” compounds are well known drugs that are used [...] Read more.
Furins are serine endoproteases that are involved in many biological processes, where they play important roles in normal metabolism, in the activation of various pathogens, while they are a target for therapeutic intervention. Dichlorophenyl-pyridine “BOS” compounds are well known drugs that are used as inhibitors of human furin by an induced-fit mechanism, in which tryptophan W254 in the furin catalytic cleft acts as a molecular transition energy gate. The binding of “BOS” drug into the active center of furin has been computationally studied using the density functional theory (DFT) and ONIOM multiscaling methodologies. The binding enthalpies of the W254 with the furin-BOS is −32.8 kcal/mol (“open”) and −18.8 kcal/mol (“closed”), while the calculated torsion barrier was found at 30 kcal/mol. It is significantly smaller than the value of previous MD calculations due to the relaxation of the environment, i.e., nearby groups of the W254, leading to the reduction of the energy demands. The significant lower barrier explains the experimental finding that the dihedral barrier of W254 is overcome. Furthermore, sartans were studied to evaluate their potential as furin inhibitors. Sartans are AT1 antagonists, and they effectively inhibit the hypertensive effects induced by the peptide hormone Angiotensin II. Here, they have been docked into the cavity to evaluate their effect on the BOS ligand via docking and molecular dynamics simulations. A consistent binding of sartans within the cavity during the simulation was found, suggesting that they could act as furin inhibitors. Finally, sartans interact with the same amino acids as W254, leading to a competitive binding that may influence the pharmacological efficacy and potential drug interactions of sartans. Full article
(This article belongs to the Special Issue Synthesis and Theoretical Study of Bioactive Molecules)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of (<b>a</b>) sartans, (<b>b</b>) furin enzyme, (<b>c</b>) BOS-318 ligand, and (<b>d</b>) tryptophan amino acid. (C atoms: grey balls, H: white, O: red, N: blue, F: cyan, Cl: green).</p>
Full article ">Figure 2
<p>Optimized conformations of furin–BOS complex with “open” and “closed” conformation of W254 amino acids via ONIOM(M06-2X/6-311+G(d,p):PM6) method.</p>
Full article ">Figure 3
<p>Interactions of losartan with 7LCU in (<b>a</b>) 2D and (<b>b</b>) 3D presentation.</p>
Full article ">Figure 4
<p>Interactions of Benzimidazole bis-N,N′-biphenyltetrazole with 7LCU in (<b>a</b>) 2D and (<b>b</b>) 3D presentation.</p>
Full article ">Figure 5
<p>Interactions of BV6 with 7LCU in (<b>a</b>) 2D and (<b>b</b>) 3D presentation.</p>
Full article ">Figure 6
<p>Interactions of candesartan with 7LCU in (<b>a</b>) 2D and (<b>b</b>) 3D presentation.</p>
Full article ">Figure 7
<p>Interactions of losartan carboxylic acid with 7LCU in (<b>a</b>) 2D and (<b>b</b>) 3D presentation.</p>
Full article ">Figure 8
<p>Interactions of nirmitrevil with 7LCU in (<b>a</b>) 2D and (<b>b</b>) 3D presentation.</p>
Full article ">Figure 9
<p>Interactions of telmisartan with 7LCU in (<b>a</b>) 2D and (<b>b</b>) 3D presentation.</p>
Full article ">Figure 10
<p>Interactions of benzimidazole-N-biphenyltetrazole with 7LCU in (<b>a</b>) 2D and (<b>b</b>) 3D presentation.</p>
Full article ">Figure 11
<p>Protein’s Ca atoms display a RMSD value &lt;8.0 Å. Blue line for the protein and red line for the ligand.</p>
Full article ">
20 pages, 7443 KiB  
Article
Interactions between Inhibitors and 5-Lipoxygenase: Insights from Gaussian Accelerated Molecular Dynamics and Markov State Models
by Yuyang Liu, Kaiyu Wang, Fuyan Cao, Nan Gao and Wannan Li
Int. J. Mol. Sci. 2024, 25(15), 8295; https://doi.org/10.3390/ijms25158295 - 30 Jul 2024
Viewed by 343
Abstract
Inflammation is a protective stress response triggered by external stimuli, with 5-lipoxygenase (5LOX) playing a pivotal role as a potent mediator of the leukotriene (Lts) inflammatory pathway. Nordihydroguaiaretic acid (NDGA) functions as a natural orthosteric inhibitor of 5LOX, while 3-acetyl-11-keto-β-boswellic acid (AKBA) acts [...] Read more.
Inflammation is a protective stress response triggered by external stimuli, with 5-lipoxygenase (5LOX) playing a pivotal role as a potent mediator of the leukotriene (Lts) inflammatory pathway. Nordihydroguaiaretic acid (NDGA) functions as a natural orthosteric inhibitor of 5LOX, while 3-acetyl-11-keto-β-boswellic acid (AKBA) acts as a natural allosteric inhibitor targeting 5LOX. However, the precise mechanisms of inhibition have remained unclear. In this study, Gaussian accelerated molecular dynamics (GaMD) simulation was employed to elucidate the inhibitory mechanisms of NDGA and AKBA on 5LOX. It was found that the orthosteric inhibitor NDGA was tightly bound in the protein’s active pocket, occupying the active site and inhibiting the catalytic activity of the 5LOX enzyme through competitive inhibition. The binding of the allosteric inhibitor AKBA induced significant changes at the distal active site, leading to a conformational shift of residues 168–173 from a loop to an α-helix and significant negative correlated motions between residues 285–290 and 375–400, reducing the distance between these segments. In the simulation, the volume of the active cavity in the stable conformation of the protein was reduced, hindering the substrate’s entry into the active cavity and, thereby, inhibiting protein activity through allosteric effects. Ultimately, Markov state models (MSM) were used to identify and classify the metastable states of proteins, revealing the transition times between different conformational states. In summary, this study provides theoretical insights into the inhibition mechanisms of 5LOX by AKBA and NDGA, offering new perspectives for the development of novel inhibitors specifically targeting 5LOX, with potential implications for anti-inflammatory drug development. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Sequence and the secondary structure diagram of 5LOX. (<b>B</b>) The active residues around the NDAG inhibitor binding to 5LOX; (<b>C</b>) the active residues around the AKBA inhibitor binding to the 5LOX–AKBA complex.</p>
Full article ">Figure 2
<p>The binding sites of the substrate arachidonic acid (AA) in 5LOX proteins. (The substrate AA is red, FE<sup>2+</sup> is orange and the active residues are yellow).</p>
Full article ">Figure 3
<p>Analysis of structural stability. (<b>A</b>) The temporal evolution of the RMSD from their initial structure of the Apo, 5LOX–AA, 5LOX–NDGA, and 5LOX–AKBA systems. (<b>B</b>) Distribution of RMSD values in the four systems. (<b>C</b>) The temporal evolution of the R<sub>g</sub> from their initial structure of the Apo, 5LOX–AA, 5LOX–NDGA, and 5LOX–AKBA systems. (<b>D</b>) Distribution of R<sub>g</sub> values in the four systems. (<b>E</b>) The temporal evolution of the SASA from their initial structure of the Apo, 5LOX–AA, 5LOX–NDGA, and 5LOX–AKBA systems. (<b>F</b>) Distribution of SASA values in the four systems. The median (the horizontal line in the center), the mean (the black dot), and the interquartile range (the upper and lower edges of the box) are shown in the box plot for each set of data.</p>
Full article ">Figure 4
<p>(<b>A</b>) RMSF diagrams for the four systems. (<b>B</b>) Structure of residues 170–175 in red. (<b>C</b>) Structure of residues 176–195 in green. (<b>D</b>) Structure of residues 280–300 in blue. (<b>E</b>) Structure of residues 406–425 in pink.</p>
Full article ">Figure 5
<p>The probability of secondary structure changes in residues 150–200 for (<b>A</b>) Apo, (<b>B</b>) 5LOX–AA, (<b>C</b>) 5LOX–NDGA, and (<b>D</b>) 5LOX–AKBA. (<b>E</b>) Comparison of the structures of residues 168–173 and residues 285–290 before (yellow) and after (red) the binding of the allosteric inhibitor AKBA.</p>
Full article ">Figure 6
<p>The dynamical cross-correlation matrix diagrams of (<b>A</b>) Apo, (<b>B</b>) 5LOX–AA, (<b>C</b>) 5LOX–NDGA, and (<b>D</b>) 5LOX–AKBA systems. (The large red box contains residues 168–173 and 375–425, and the small red box contains 285–290 and 375–400).</p>
Full article ">Figure 7
<p>The free-energy landscape for the following four systems: (<b>A</b>) Apo, (<b>B</b>) 5LOX–AA, (<b>C</b>) 5LOX–NDGA, and (<b>D</b>) 5LOX–AKBA. The active cavity structures of representative conformations in the low-energy region are indicated by different colors.</p>
Full article ">Figure 8
<p>(<b>A</b>) Energy landscape colored by secondary structure percentage of 5LOX–NDGA system. (<b>B</b>) Energy landscape colored by secondary structure percentage of 5LOX–AKBA system. (<b>C</b>) Markov state models and mean first passage time of 5LOX–NDGA. (<b>D</b>) Markov state models and mean first passage time of 5LOX–AKBA. (Residues 155–195 are blue and residues 280–295 are red).</p>
Full article ">
18 pages, 3728 KiB  
Article
Very Strong Hydrogen Bond in Nitrophthalic Cocrystals
by Kinga Jóźwiak, Aneta Jezierska, Jarosław J. Panek, Andrzej Kochel, Barbara Łydżba-Kopczyńska and Aleksander Filarowski
Molecules 2024, 29(15), 3565; https://doi.org/10.3390/molecules29153565 - 29 Jul 2024
Viewed by 518
Abstract
This work presents the studies of a very strong hydrogen bond (VSHB) in biologically active phthalic acids. Research on VSHB comes topical due to its participation in many biological processes. The studies cover the modelling of intermolecular interactions and phthalic acids with 2,4,6-collidine [...] Read more.
This work presents the studies of a very strong hydrogen bond (VSHB) in biologically active phthalic acids. Research on VSHB comes topical due to its participation in many biological processes. The studies cover the modelling of intermolecular interactions and phthalic acids with 2,4,6-collidine and N,N-dimethyl-4-pyridinamine complexes with aim to obtain a VSHB. The four synthesized complexes were studied by experimental X-ray, IR, and Raman methods, as well as theoretical Car–Parrinello Molecular Dynamics (CP-MD) and Density Functional Theory (DFT) simulations. By variation of the steric repulsion and basicity of the complex’ components, a very short intramolecular hydrogen bond was achieved. The potential energy curves calculated by the DFT method were characterized by a low barrier (0.7 and 0.9 kcal/mol) on proton transfer in the OHN intermolecular hydrogen bond for 3-nitrophthalic acid with either 2,4,6-collidine or N,N-dimethyl-4-pyridinamine cocrystals. Moreover, the CP-MD simulations exposed very strong bridging proton dynamics in the intermolecular hydrogen bonds. The accomplished crystallographic and spectroscopic studies indicate that the OHO intramolecular hydrogen bond in 4-nitrophthalic cocrystals is VSHB. The influence of a strong steric effect on the geometry of the studied cocrystals and the stretching vibration bands of the carboxyl and carboxylate groups was elaborated. Full article
(This article belongs to the Special Issue Molecular Modeling: Advancements and Applications, 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of 3-nitrophthalic acid with 2,4,6-collidine (<b>3NFA-2C</b>), 4-nitrophthalic acid with 2,4,6-collidine (<b>4NFA-C</b>), 4-nitrophthalic acid with N,N-dimethyl-4-pyridinamine (<b>4NFA-DMAP</b>), and 3-nitrophthalic acid–N,N-dimethyl-4-pyridinamine dihydrate (<b>3NFA-2W-2DMAP</b>) complexes.</p>
Full article ">Figure 2
<p>Crystal structures of the <b>3NFA-2C</b>, <b>4NFA-C</b>, <b>4NFA-DMAP</b>, and <b>3NFA-2W-2DMAP</b> cocrystals. Hydrogen bonds are denoted with dashed lines. Displacement ellipsoids are plotted at 50% probability level.</p>
Full article ">Figure 3
<p>Experimental ATR and Raman spectra of <b>3NFA</b>, <b>4NFA</b>, <b>3NFA-2C</b>, <b>4NFA-C</b>, <b>3NFA-2W-2DMAP</b>, and <b>4NFA-DMAP</b>.</p>
Full article ">Figure 4
<p>The time evolution of donor–proton (green, d(DH) in Å), proton–acceptor (blue, d(AH) in Å), and donor–acceptor (red, d(DA) in Å) distances simulated by the CP-MD method in the solid state (T = 300 K) for the <b>3NFA-2C</b>, <b>4NFA-C</b>, <b>3NFA-2W-2DMAP</b>, and <b>4NFA-DMAP</b> complexes.</p>
Full article ">Figure 5
<p>The experimental ATR spectra and atomic velocity power spectra for the bridging protons calculated by the CP-MD method of the <b>3NFA-2C</b>, <b>4NFA-C</b>, <b>3NFA-2W-2DMAP</b>, and <b>4NFA-DMAP</b> complexes. The <b>3NFA</b> and <b>4NFA</b> complexes are presented on the left and right panels, respectively.</p>
Full article ">Figure 6
<p>Fragments of experimental ATR spectra and X-ray structures of 3-nitrophthalic acid (<b>3-NFA</b>) [<a href="#B56-molecules-29-03565" class="html-bibr">56</a>] and 4-nitrophthalic acid (<b>4-NFA</b>) as well as 3-nitrophthalic acid–2,4,6-collidine (<b>3NFA-2C</b>), 4-nitrophthalic acid–2,4,6-collidine (<b>4NFA-C</b>), 3-nitrophthalic acid–N,N-dimethyl-4-pyridinamine dihydrate (<b>3NFA-2W-2DMAP</b>), and 4-nitrophthalic acid–N,N-dimethyl-4-pyridinamine (<b>4NFA-DMAP</b>) complexes.</p>
Full article ">Figure 7
<p>Calculated (B3LYP-D3/6-311+G(d,p)) potential energy curves for gradual elongation of one proton within the inter/intramolecular hydrogen bonds in the <b>3NFA-2C</b>, <b>4NFA-C</b>, <b>3NFA-2DMAP</b>, and <b>4NFA-DMAP</b> complexes. The black and white arrows indicate the N-H⋯O and O-H⋯O hydrogen bonds, respectively.</p>
Full article ">Figure 8
<p>Two-dimensional histograms for the hydrogen atom position in the respective hydrogen bonds obtained by CP-MD simulations at 300 K. Y axes denote the donor–acceptor distances, X axes are the donor–proton distances. Isocontours are drawn at 1 (blue), 5 (green), and 20 (red) Å<sup>−2</sup> probability density values (<b>upper</b> panels). Radial distribution function (RDF) of the studied hydrogen bonds (<b>lower</b> panels).</p>
Full article ">
15 pages, 8468 KiB  
Article
Groundwater Discharge Limits of Mountain Tunnels Based on the Normal Growth of Typical Herbaceous Plants
by Yuanfu Zhou, Xuefu Zhang, Yuanpeng Liu and Yuanguang Yang
Appl. Sci. 2024, 14(15), 6561; https://doi.org/10.3390/app14156561 - 26 Jul 2024
Viewed by 439
Abstract
The construction of mountain tunnels can lead to groundwater loss and severely impact plant growth. In order to study the limited discharge of groundwater in mountain tunnels for the normal growth of typical herbaceous plants, a tunnel in the alpine meadow area of [...] Read more.
The construction of mountain tunnels can lead to groundwater loss and severely impact plant growth. In order to study the limited discharge of groundwater in mountain tunnels for the normal growth of typical herbaceous plants, a tunnel in the alpine meadow area of Qinghai Province was taken as the research objective. Based on transplant experiments, numerical simulations, and the empirical calculation of tunnel discharge limits, the minimum water level required for the normal growth of herbaceous plants, groundwater changes, and grouting parameters during tunnel construction, as well as limited discharge values of groundwater based on the normal growth requirements of plants, were studied. The results indicate that when the groundwater level declined by 0.6–0.8 m, herbaceous plants were able to normally grow. Generally, tunnel excavation lowered the groundwater level so that the normal growth of herbaceous plants was significantly affected. The reasonable grouting parameters were obtained by numerical simulation. They were able to ensure that the groundwater level decline was less than 0.8 m and ultimately recovered to over 90% of the pre-construction level. The herbaceous plants in Qinghai’s alpine grasslands were able to normally grow when the groundwater discharge limit was 0.2~4.0 m3/(m·d). This research offers guidance and support for managing groundwater discharge during tunnel construction in ecologically fragile areas, such as the Three Rivers Source in Qinghai. Full article
(This article belongs to the Special Issue Tunnel and Underground Engineering: Recent Advances and Challenges)
Show Figures

Figure 1

Figure 1
<p>Box model of experiment.</p>
Full article ">Figure 2
<p>The growth height of <span class="html-italic">Artemisia annua</span>, changing with time.</p>
Full article ">Figure 3
<p>Transplantation status of <span class="html-italic">Artemisia annua</span> 1 and 6. (<b>a</b>) <span class="html-italic">Artemisia annua</span> 1. (<b>b</b>) <span class="html-italic">Artemisia annua</span> 6.</p>
Full article ">Figure 4
<p>The growth height of <span class="html-italic">Stipa</span>, changing with time.</p>
Full article ">Figure 5
<p>Transplantation status of <span class="html-italic">Stipa</span> 1 and <span class="html-italic">Stipa</span> 7. (<b>a</b>) <span class="html-italic">Stipa</span> 1. (<b>b</b>) <span class="html-italic">Stipa</span> 7.</p>
Full article ">Figure 6
<p>Calculation model diagram.</p>
Full article ">Figure 7
<p>Tunnel drainage volume at each construction stage.</p>
Full article ">Figure 8
<p>Water level changes at three measuring points near the surface in each construction stage.</p>
Full article ">Figure 9
<p>Changes in water level in different construction and discharge stages. (<b>a</b>) Full drainage. (<b>b</b>) Limited drainage. (<b>c</b>) Full sealing.</p>
Full article ">Figure 10
<p>Water level at different grouting ranges at various measuring points during the construction process. (<b>a</b>) The lowest water level. (<b>b</b>) The final recovery water level.</p>
Full article ">Figure 11
<p>Water level at different measuring points with different permeability coefficients during the construction process. (<b>a</b>) The lowest water level. (<b>b</b>) The final recovery water level.</p>
Full article ">Figure 12
<p>The influence of different factors on the limited discharge value of groundwater. (<b>a</b>) Annual rainfall and water head height. (<b>b</b>) Precipitation infiltration coefficient and water head height. (<b>c</b>) Formation permeability coefficient and water head height.</p>
Full article ">
Back to TopTop