[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
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,092)

Search Parameters:
Keywords = DMSO

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 3300 KiB  
Article
Glycerol Carbonate Production via Transesterification: The Effect of Support Porosity and Catalyst Basicity
by Ricardo L. Souza Júnior, Larissa C. Eira, Chaline Detoni and Mariana M. V. M. Souza
Processes 2024, 12(10), 2256; https://doi.org/10.3390/pr12102256 - 16 Oct 2024
Viewed by 403
Abstract
Glycerol transesterification with diethyl carbonate (DEC) using catalysts with different porosities as support for CaO was performed, seeking the evaluation of how textural properties influence glycerol conversion and product selectivity. A total of 20% CaO was supported on ZSM-5, K-10, MCM-41, SiO2 [...] Read more.
Glycerol transesterification with diethyl carbonate (DEC) using catalysts with different porosities as support for CaO was performed, seeking the evaluation of how textural properties influence glycerol conversion and product selectivity. A total of 20% CaO was supported on ZSM-5, K-10, MCM-41, SiO2, and γ-Al2O3. Catalysts showed a well-dispersed active phase of CaO in all the supports and no changes in the support crystalline structure were noticed. Reactions were performed in dimethyl sulfoxide (DMSO), 10 wt.% of catalyst in relation to glycerol, at 130 °C, and 1:3 glycerol/DEC molar ratio. According to our results, the higher the pore volume and pore size, the higher the glycerol conversion. On the other hand, concerning selectivity, higher glycerol carbonate selectivities were reached when strong basic sites were present. A total of 86% glycerol conversion and 91% glycerol carbonate selectivity were found using 60% CaO supported on γ-Al2O3 after 5 h of reaction. Full article
(This article belongs to the Section Catalysis Enhanced Processes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Conversion (<b>A</b>), GlyC selectivity (<b>B</b>), and GlyC yield after 6 h of reaction (<b>C</b>) of oxides. Reactions were performed at 130 °C, 1:3 for gly/DEC, and 10 wt% of catalyst.</p>
Full article ">Figure 2
<p>Conversion (<b>A</b>) and GlyC selectivity (<b>B</b>) of supported catalysts with 20% CaO. Reactions were performed at 130 °C, 1:3 for gly/DEC, and 10 wt% of catalyst.</p>
Full article ">Figure 3
<p>Comparison of pore volume × GlyC yield. Reactions performed at 130 °C, 1:3 for gly/DEC with 10 wt% catalyst after 5 h of reaction.</p>
Full article ">Figure 4
<p>Glycerol conversion (<b>A</b>) and GlyC selectivity (<b>B</b>) of CaO supported on γ-Al<sub>2</sub>O<sub>3</sub>. Reaction performed at 130 °C, 1:3 for gly/DEC, and 10 wt% of catalyst.</p>
Full article ">Figure 5
<p>Comparison of basic sites × conversion × selectivity × yield. Reactions performed at 130 °C, 1:3 for gly/DEC with 10 wt% catalyst for 5 h of reaction.</p>
Full article ">Figure 6
<p>XRD patterns of MgO, CaO, and SrO.</p>
Full article ">Figure 7
<p>XRD patterns of support and supported catalysts (<b>A</b>) SiO<sub>2</sub>, (<b>B</b>) ZSM-5, (<b>C</b>) K-10, and (<b>D</b>) MCM-41.</p>
Full article ">Figure 8
<p>XRD patterns of support and catalyst supported on γ-Al<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 9
<p>CO<sub>2</sub>-TPD profiles of MgO, CaO, and SrO.</p>
Full article ">Figure 10
<p>CO<sub>2</sub>-TPD profiles of the supports and catalysts: (<b>A</b>) SiO<sub>2</sub>, (<b>B</b>) ZSM-5 (<b>C</b>) K-10, and (<b>D</b>) MCM-41.</p>
Full article ">Figure 11
<p>CO<sub>2</sub>-TPD profiles of the catalysts supported on Al<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 12
<p>Recycle tests for the 60CaO/γ-Al<sub>2</sub>O<sub>3</sub> catalyst. Reactions performed at 130 °C, 1:3 for gly/DEC with 10 wt% catalyst.</p>
Full article ">Figure 13
<p>XRD patterns of the 60CaO/γ-Al<sub>2</sub>O<sub>3</sub> catalyst evaluated in recycle tests.</p>
Full article ">
24 pages, 9382 KiB  
Article
Polyacrylonitrile Ultrafiltration Membrane for Separation of Used Engine Oil
by Alexandra Nebesskaya, Anastasia Kanateva, Roman Borisov, Alexey Yushkin, Vladimir Volkov and Alexey Volkov
Polymers 2024, 16(20), 2910; https://doi.org/10.3390/polym16202910 (registering DOI) - 16 Oct 2024
Viewed by 271
Abstract
The separation of used engine oil (UEO) with an ultrafiltration (UF) membrane made of commercial copolymer of poly(acrylonitrile-co-methyl acrylate) (P(AN-co-MA)) has been investigated. The P(AN-co-MA) sample was characterized by using FTIR spectroscopy, 13C NMR spectroscopy, and XRD. The UF membrane with a [...] Read more.
The separation of used engine oil (UEO) with an ultrafiltration (UF) membrane made of commercial copolymer of poly(acrylonitrile-co-methyl acrylate) (P(AN-co-MA)) has been investigated. The P(AN-co-MA) sample was characterized by using FTIR spectroscopy, 13C NMR spectroscopy, and XRD. The UF membrane with a mean pore size of 23 nm was fabricated by using of non-solvent-induced phase separation method—the casting solution of 13 wt.% P(AN-co-MA) in dimethylsulfoxide (DMSO) was precipitated in the water bath. Before the experiment, the used engine oil was diluted with toluene, and the resulting UEO solution in toluene (100 g/L) was filtered through the UF membrane in the dead-end filtration mode. Special attention was given to the evaluation of membrane fouling; for instance, the permeability of UEO solution was dropped from its initial value of 2.90 L/(m2·h·bar) and then leveled off at 0.75 L/(m2·h·bar). However, the membrane cleaning (washing with toluene) allowed a recovery of 79% of the initial pure toluene flux (flux recovery ratio), indicating quite attractive membrane resistance toward irreversible fouling with engine oil components. The analysis of the feed, retentate, and permeate by various analytical methods showed that the filtration through the UF membrane made of P(AN-co-MA) provided the removal of major contaminants of used engine oil including polymerization products and metals (rejection—96.3%). Full article
(This article belongs to the Section Polymer Membranes and Films)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Membrane preparation scheme.</p>
Full article ">Figure 2
<p><sup>13</sup>C NMR spectroscopy spectra poly(acrylonitrile-co-methyl acrylate).</p>
Full article ">Figure 3
<p>FTIR spectrum of P(AN-co-MA).</p>
Full article ">Figure 4
<p>XRD spectrum of polymer: “crystalline” peaks—transparent, “amorphous”—shaded.</p>
Full article ">Figure 5
<p>Asymmetric ultrafiltration P(AN-co-MA) membrane images. (<b>a</b>) the original membrane, (<b>b</b>) demonstration of membrane flexibility.</p>
Full article ">Figure 6
<p>SEM images of the cross-section (<b>a</b>) and the surface (<b>b</b>) of the P(AN-co-MA) membrane.</p>
Full article ">Figure 7
<p>Time dependence of the UEO solution permeance through the P(AN-co-MA) membrane.</p>
Full article ">Figure 8
<p>The recovery rate of the membrane fouled during filtration of UEO solution in toluene (100 g/L).</p>
Full article ">Figure 9
<p>FTIR spectrum of the P(AN-co-MA) membrane surface before and after filtering.</p>
Full article ">Figure 10
<p>Photographs of (1) feed (UEO as received), (2) permeate, and (3) retentate after filtrations of UEO solutions in toluene (100 g/L): permeate and retentate after removal of toluene by distillation.</p>
Full article ">Figure 11
<p>Metal content in the UEO, permeate, and retentate. (<b>a</b>) Zn and Na, (<b>b</b>) Cu, Pb and Fe.</p>
Full article ">Figure 12
<p>Group hydrocarbon composition of UEO and permeate.</p>
Full article ">Figure 13
<p>Chromatograms of used engine oil, permeate, and retentate obtained by the fingerprint method during filtration through a P(AN-co-MA) membrane. Conditions: 50 °C (2 min), 4 °C/min, 300 °C (40 min); carrier gas—helium, column SP-Sil 5 CB; inlet column pressure: 312.8 kPa.</p>
Full article ">Figure 14
<p><sup>1</sup>H NMR spectroscopy spectra of oils (9.0−6.0 ppm): feed, permeate, and retentate.</p>
Full article ">Figure 15
<p><sup>1</sup>H NMR spectroscopy spectra of oils (2.0−0.0 ppm): feed, permeate, and retentate.</p>
Full article ">
12 pages, 4892 KiB  
Article
2-Pyridylmetallocenes, Part IX. Sulphur-Substituted 2-Pyridylferrocene: Synthesis and Reactivity towards Pt(II) and Hg(II)
by Stefan Weigand and Karlheinz Sünkel
Molecules 2024, 29(20), 4884; https://doi.org/10.3390/molecules29204884 (registering DOI) - 15 Oct 2024
Viewed by 230
Abstract
Thio-substituted 2-pyridylferrocenes [CpFe{C5H3(X)(C5H4N)}] (X = SOTol, 3; SMe, 5) were prepared from [CpFe(C5H4R)] (R = SOTol, 1; 2-C5H4N, 2) in moderate yields. The [...] Read more.
Thio-substituted 2-pyridylferrocenes [CpFe{C5H3(X)(C5H4N)}] (X = SOTol, 3; SMe, 5) were prepared from [CpFe(C5H4R)] (R = SOTol, 1; 2-C5H4N, 2) in moderate yields. The reactions of 3 and 5 with [PtCl2(DMSO)2] yielded the binuclear N, S chelated complexes [CpFe{C5H3(X)(C5H4N)}-(к-N,S)-PtCl2] (X= SOTol, 4, SMe, 6), while the reaction of 5 with Hg(OAc)2/LiCl led to cyclomercuration with generation of [CpFe{C5H2(SMe)(C5H4N)(HgCl)}], 7. The crystal structures of 6·CH2Cl2 and 7 were determined. The structure of 6 showed a weak intramolecular Fe…Pt interaction and several weak intermolecular interactions involving all Cl atoms. Weak intermolecular interactions between Hg and S atoms in the cyclomercurated 7 led to a tetrameric structure involving a Hg2S2 ring. Full article
Show Figures

Figure 1

Figure 1
<p>Molecular structure of compound <b>6</b>. Displacement ellipsoids at the 50% probability level.</p>
Full article ">Figure 2
<p>Molecular structure of compound <b>7</b>, top views of molecules A and B. Displacement ellipsoids are at the 50% probability level.</p>
Full article ">Figure 3
<p>Hydrogen Bonds in compound <b>6.</b> <span class="html-small-caps">mercury </span>standard colouring: Carbon: dark grey; hydrogen: light grey; chlorine: green; sulphur: yellow; iron: orange; nitrogen: blue. Red lines “hanging contacts”, blue lines: both molecules connected are completely in the picture.</p>
Full article ">Figure 4
<p>Hydrogen Bonds in compound <b>7.</b> <span class="html-small-caps">Mercury</span> colours as defined in <a href="#molecules-29-04884-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>The formation of four-membered Hg<sub>2</sub>S<sub>2</sub> rings in compound <b>7.</b> <span class="html-small-caps">Mercury</span> colours as defined in <a href="#molecules-29-04884-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 6
<p>Packing diagram (<span class="html-small-caps">mercury</span>) of compound <b>6</b>, viewed along <span class="html-italic">b.</span> <span class="html-small-caps">Mercury</span> colours as defined in <a href="#molecules-29-04884-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 7
<p>Packing diagram (<span class="html-small-caps">mercury</span>) of compound <b>7</b>, viewed along <span class="html-italic">a.</span> <span class="html-small-caps">Mercury</span> colours as defined in <a href="#molecules-29-04884-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Scheme 1
<p>Synthesis of <b>3</b>.</p>
Full article ">Scheme 2
<p>Synthesis of <b>5</b>.</p>
Full article ">Scheme 3
<p>Reaction of <b>3</b> with [PtCl<sub>2</sub>(DMSO)<sub>2</sub>] and NaOAc.</p>
Full article ">Scheme 4
<p>Reaction of <b>5</b> with [PtCl<sub>2</sub>(DMSO)<sub>2</sub>] and Hg(OAc)<sub>2</sub>/LiCl.</p>
Full article ">
17 pages, 2164 KiB  
Article
Implications of Lead (Pb)-Induced Transcriptomic and Phenotypic Alterations in the Aged Zebrafish (Danio rerio)
by Chia-Chen Wu, Danielle N. Meyer, Alex Haimbaugh and Tracie R. Baker
Toxics 2024, 12(10), 745; https://doi.org/10.3390/toxics12100745 - 14 Oct 2024
Viewed by 404
Abstract
Lead (Pb) is a well-known neurotoxin with established adverse effects on the neurological functions of children and younger adults, including motor, learning, and memory abilities. However, its potential impact on older adults has received less attention. Using the zebrafish model, our study aims [...] Read more.
Lead (Pb) is a well-known neurotoxin with established adverse effects on the neurological functions of children and younger adults, including motor, learning, and memory abilities. However, its potential impact on older adults has received less attention. Using the zebrafish model, our study aims to characterize the dose–response relationship between environmentally relevant Pb exposure levels and their effects on changes in behavior and transcriptomics during the geriatric periods. We exposed two-year-old zebrafish to waterborne lead acetate (1, 10, 100, 1000, or 10,000 µg/L) or a vehicle (DMSO) for 5 days. While lower concentrations (1–100 µg/L) reflect environmentally relevant Pb levels, higher concentrations (1000–10,000 µg/L) were included to assess acute toxicity under extreme exposure scenarios. We conducted adult behavior assessment to evaluate the locomotor activity following exposure. The same individual fish were subsequently sacrificed for brain dissection after a day of recovery in the aquatic system. RNA extraction and sequencing were then performed to evaluate the Pb-induced transcriptomic changes. Higher (1000–10,000 ug/L) Pb levels induced hyperactive locomotor patterns in aged zebrafish, while lower (10–100 ug/L) Pb levels resulted in the lowest locomotor activity compared to the control group. Exposure to 100 µg/L led to the highest number of differentially expressed genes (DEGs), while 10,000 µg/L induced larger fold changes in both directions. The neurological pathways impacted by Pb exposure include functions related to neurotransmission, such as cytoskeletal regulation and synaptogenesis, and oxidative stress response, such as mitochondrial dysfunction and downregulation of heat shock protein genes. These findings emphasize a U-shape dose–response relationship with Pb concentrations in locomotor activity and transcriptomic changes in the aging brain. Full article
Show Figures

Figure 1

Figure 1
<p>Velocity distribution of 2-year-old <span class="html-italic">Danio rerio</span> after 5 days of Pb exposure. Colors represent exposure concentrations: 0 (light blue), 1 (dark blue), 10 (light green), 100 (dark green), 1000 (pink), and 10,000 (red). The subplot displays a box plot indicating the median, quartiles, and outliers of velocity for each exposure. Pairwise Dunn–Bonferroni post hoc test pair test compares exposure groups with control (** <span class="html-italic">p</span>-value &lt; 0.01, *** <span class="html-italic">p</span>-value &lt; 0.001).</p>
Full article ">Figure 2
<p>Upset plot of the frequency distrubition of differentially expressed genes (DEGs) and their intersection in 2-year-old <span class="html-italic">Danio rerio</span> after 5 days of Pb exposure. The upset plot was generated using the package UpSetR in R software (version 4.2.3).</p>
Full article ">Figure 3
<p>Volcano plots of differentially expressed genes (DEGs) in 2-year-old <span class="html-italic">Danio rerio</span> after 5 days of Pb exposure. Grey lines indicate a threshold of <span class="html-italic">p</span>-value &lt; 0.05 and an absolute log2 fold change at 0.75. Colors represent exposure concentrations: 1 (dark blue), 10 (light green), 100 (dark green), 1000 (pink), and 10,000 (red). Individual volcano plots along with the number of up- and downregulated DEGs and their corresponding log2 fold changes for each treatment can be found in <a href="#app1-toxics-12-00745" class="html-app">Figure S2</a>.</p>
Full article ">Figure 4
<p>Ingenuity Pathways Analysis for transcriptomic changes implicated in neurological diseases and canonical pathways after 5-day Pb exposure in 2-year-old <span class="html-italic">Danio rerio</span>. Dot symbols indicate the number of differentially expressed genes (DEGs) associated with each implication, with color representing the <span class="html-italic">p</span>-value.</p>
Full article ">Figure 5
<p>Highly differentially expressed genes (DEGs) relevant to aging-related functions in 2-year-old <span class="html-italic">Danio rerio</span> after 5-day Pb exposure, including non-monotonic transcriptomic responses. Colors represent the log2 fold changes (log2 FC) of each DEG.</p>
Full article ">
21 pages, 6481 KiB  
Article
In Situ Formation of Acidic Comonomer during Thermal Treatment of Copolymers of Acrylonitrile and Its Influence on the Cyclization Reaction
by Roman V. Toms, Daniil A. Ismaylov, Alexander Yu. Gervald, Nickolay I. Prokopov, Anna V. Plutalova and Elena V. Chernikova
Polymers 2024, 16(19), 2833; https://doi.org/10.3390/polym16192833 - 7 Oct 2024
Viewed by 577
Abstract
Binary and ternary copolymers of acrylonitrile (AN), tert-butyl acrylate (TBA), and n-butyl acrylate (BA) are synthesized through conventional radical polymerization in DMSO in the presence of 2-mercaptoethanol. The thermal behavior of binary and ternary copolymers is studied under argon atmosphere and [...] Read more.
Binary and ternary copolymers of acrylonitrile (AN), tert-butyl acrylate (TBA), and n-butyl acrylate (BA) are synthesized through conventional radical polymerization in DMSO in the presence of 2-mercaptoethanol. The thermal behavior of binary and ternary copolymers is studied under argon atmosphere and in air. It is demonstrated that the copolymers of AN contain 1–10 mol.% of TBA split isobutylene upon heating above 160 °C, resulting in the formation of the units of acrylic acid in the chain. The carboxylic groups formed in situ are responsible for the ionic mechanism of cyclization, which starts at lower temperatures compared with pure polyacrylonitrile (PAN) or AN copolymer with BA. The activation energy of cyclization through ionic and radical mechanisms depends on copolymer composition. For the ionic mechanism, the activation energy lies in the range ca. 100–130 kJ/mole, while for the radical mechanism, it lies in the range ca. 150–190 kJ/mole. The increase in the TBA molar part in the copolymer is followed by faster consumption of nitrile groups and the evolution of a ladder structure in both binary and ternary copolymers. Thus, the incorporation of a certain amount of TBA in PAN or its copolymer with BA allows tuning the temperature range of cyclization. This feature seems attractive for applications in the production of melt-spun PAN by choosing the appropriate copolymer composition and heating mode. Full article
(This article belongs to the Special Issue Advanced Polymer Materials: Synthesis, Structure, and Properties)
Show Figures

Figure 1

Figure 1
<p>The chemical structures of binary AN–TBA (<b>a</b>) and ternary copolymers AN–TBA–BA (<b>b</b>).</p>
Full article ">Figure 2
<p>Dependences of the monomer conversion on the polymerization time (<b>a</b>) and of the TBA average molar part in the copolymer on the monomer conversion (<b>b</b>) for copolymerization of AN and TBA at various monomer feeds (indicated in the Figures) and 0.05 mol.% of ME.</p>
Full article ">Figure 3
<p>TGA curves of the AN copolymers with various content of TBA; <span class="html-italic">f</span><sub>TBA</sub> = 0 (1), 1.0 (2), 2.5 (3), 5.0 (4), and 10.0 mol.% (5), and polyacrylic acid (6) recorded under an argon atmosphere at a heating rate of 10 K/min.</p>
Full article ">Figure 4
<p>FT-IR spectra of AN–TBA copolymers synthesized from different monomer feeds before and after thermal treatment of their 5 wt % solutions in DMSO: <span class="html-italic">f</span><sub>TBA</sub> = 5.0 (<b>a</b>), 7.5 (<b>b</b>), and 10.0 mol.% (<b>c</b>). (<b>a</b>) Conditions: room temperature (1), 3 h at 140 °C (2), 24 h at 140 °C (3), 2 h at 160 °C (4), 5 h at 160 °C (5), 24 h at 160 °C (6), 5 h at 180 °C (7), and 24 h at 180 °C (8); (<b>b</b>), (<b>c</b>) Conditions: room temperature (1), 3 h at 140 °C (2), 24 h at 140 °C (3), 2 h at 160 °C (4), 5 h at 160 °C (5), 24 h at 160 °C (6), and 5 h at 180 °C (7).</p>
Full article ">Figure 5
<p>FTIR spectra of AN–TBA copolymers synthesized from different monomer feeds before and after thermal treatment of their films at 180 °C: <span class="html-italic">f</span><sub>TBA</sub> = 5.0 (<b>a</b>), 7.5 (<b>b</b>), and 10.0 mol.% (<b>c</b>). Time: (<b>a</b>) 0 (1), 5 (2), 20 (3), and 40 min (4); (<b>b</b>), (<b>c</b>) 0 (1), 5 (2), 10 (3), 20 (4), 30 (5), and 40 min (6).</p>
Full article ">Figure 6
<p>DSC thermograms of copolymers of AN–TBA synthesized in the presence of 0.05 (<b>a</b>) and 0.3 mol.% ME (<b>b</b>) and containing different amounts of TBA; (<b>a</b>) <span class="html-italic">f</span><sub>TBA</sub> = 0 (1), 1.0 (2), 2.5 (3), 5.0 (4), 7.5 (5), and 10 mol.% (6); (<b>b</b>) <span class="html-italic">f</span><sub>TBA</sub> = 2.5 (1) and 5 mol.% (2). Heating rate 10 °C/min; recorded in argon.</p>
Full article ">Figure 7
<p>DSC thermograms of copolymers of AN–TBA synthesized from the monomer feed containing 5.0 (<b>a</b>) and 7.5 mol.% TBA and 0.05 mol.% ME (<b>b</b>). Heating rate 2.5 (1), 5 (2), 10 (3), and 20 °C/min (4); recorded in argon.</p>
Full article ">Figure 8
<p>FTIR spectra of AN–TBA copolymers synthesized from the mixture containing 10 mol.% TBA and heated at 200 (<b>a</b>) and 225 °C (<b>b</b>) in argon for the required period, specified in the Figures.</p>
Full article ">Figure 9
<p>The dependences of <span class="html-italic">φ</span><sub>CN</sub> (<b>a</b>,<b>b</b>) and Es (<b>c</b>,<b>d</b>) on the time of thermal treatment of polymer films at 200 (<b>a</b>,<b>c</b>) and 225 °C (<b>b</b>,<b>d</b>): PAN (1), copolymers AN–TBA synthesized from the monomer mixture with TBA content 1.0 (2), 2.5 (3), 5.0 (4), and 10.0 mol.% (5).</p>
Full article ">Figure 10
<p>DSC thermograms of AN–TBA copolymers synthesized in the presence of 0.05 mol.% ME and different amounts of TBA; <span class="html-italic">f</span><sub>TBA</sub> = 0 (1), 1.0 (2), 2.5 (3), 5.0 (4), 7.5 (5), and 10.0 mol.% (6). Heating rate 10 °C/min; recorded in air.</p>
Full article ">Figure 11
<p>FTIR spectra of AN–TBA copolymers synthesized from the mixture containing 10 mol.% TBA and heated at 200 (<b>a</b>) and 225 °C (<b>b</b>) in air for the required period, specified in the Figures.</p>
Full article ">Figure 12
<p>The dependences of <span class="html-italic">φ</span><sub>CN</sub> (<b>a</b>,<b>b</b>) and E<sub>s</sub> (<b>c</b>,<b>d</b>) on the time of thermal treatment at 200 (<b>a</b>,<b>c</b>) and 225 °C (<b>b</b>,<b>d</b>) in air for AN–TBA copolymers synthesized from the mixture containing 10.0 mol.% TBA. (<b>a</b>,<b>c</b>) <span class="html-italic">f</span><sub>TBA</sub> = 1 (1), 2.5 (2), 5 (3), and 10 mol.% (4); (<b>b</b>,<b>d</b>) <span class="html-italic">f</span><sub>TBA</sub> = 0 (1), 1 (2), 2.5 (3), 5 (4), and 10 mol.% (5).</p>
Full article ">Figure 13
<p>DSC thermograms of (<b>a</b>) copolymers of AN/BA containing different amounts of BA, f<sub>BA</sub> = 5 (1), 10 (2), 15 (3), and 20 mol.% (4); (<b>b</b>) terpolymers of AN/BA/TBA containing 15 mol.% of BA and different amounts of TBA, f<sub>TBA</sub> = 0 (1), 2.5 (2), 5.0 (3), and 10.0 mol.% (4). Heating rate 10 °C/min; recorded in argon.</p>
Full article ">Figure 14
<p>TGA curves of the AN copolymers containing 15 mol.% BA and various content of TBA, f<sub>TBA</sub> = 0 (1), 1.0 (2), and 2.5 mol.% (3), recorded under an argon atmosphere at a heating rate of 10 K/min.</p>
Full article ">Figure 15
<p>FTIR spectra of AN–BA–TBA terpolymers synthesized from the mixture containing 15 mol.% BA and 5 mol.% TBA and heated at 200 (<b>a</b>) and 225 °C (<b>b</b>) in argon for the required period, specified in the Figures.</p>
Full article ">Figure 16
<p>The dependences of <span class="html-italic">φ</span><sub>CN</sub> (<b>a</b>,<b>b</b>) and E<sub>s</sub> (<b>c</b>,<b>d</b>) on the time of thermal treatment at 200 (<b>a</b>,<b>c</b>) and 225 °C (<b>b</b>,<b>d</b>) in argon for PAN (1) and AN–BA–TBA copolymers synthesized from the mixture containing 15 mol.% BA and different amounts of TBA: <span class="html-italic">f</span><sub>TBA</sub> = 0 (2), 1.0 (3), 2.5 (4), and 5.0 mol.% (5).</p>
Full article ">Figure 17
<p>FTIR spectra of AN–BA–TBA terpolymers synthesized from the mixture containing 15 mol.% BA and 2.5 (<b>a</b>) and 5.0 mol.% TBA (<b>b</b>) and heated at 225 °C in air for the required period, specified in the Figures.</p>
Full article ">Figure 18
<p>The dependences of <span class="html-italic">φ</span><sub>CN</sub> (<b>a</b>,<b>b</b>) and E<sub>s</sub> (<b>c</b>,<b>d</b>) on the time of thermal treatment at 200 (<b>a</b>,<b>c</b>) and 225 °C (<b>b</b>,<b>d</b>) in air for AN polymers: (<b>a</b>,<b>c</b>) PAN (1) and AN–BA–TBA copolymers synthesized from the mixture containing 15 mol.% BA and different amounts of TBA: <span class="html-italic">f</span><sub>TBA</sub> = 0 (2), 1.0 (3), 2.5 (4), and 5.0 mol.% (5); (<b>b</b>,<b>d</b>). AN–BA–TBA copolymers synthesized from the mixture containing 15 mol.% BA and different amounts of TBA: <span class="html-italic">f</span><sub>TBA</sub> = 0 (1), 1.0 (3), and 2.5 mol.% (4), and AN–TBA copolymer with 1 mol.% TBA (2).</p>
Full article ">Scheme 1
<p>Scheme of the chain transfer reaction during cyclization of the AN–alkyl acrylate copolymers.</p>
Full article ">Scheme 2
<p>Scheme of cyclization reaction of nitrile groups.</p>
Full article ">Scheme 3
<p>Proposed mechanism of poly(<span class="html-italic">tert</span>-butyl acrylate) pyrolysis [<a href="#B47-polymers-16-02833" class="html-bibr">47</a>].</p>
Full article ">Scheme 4
<p>Proposed mechanism of autocatalytic pyrolysis of poly(<span class="html-italic">tert</span>-butyl acrylate) [<a href="#B47-polymers-16-02833" class="html-bibr">47</a>].</p>
Full article ">Scheme 5
<p>Ionic mechanism of cyclization initiated by carboxylic groups.</p>
Full article ">
22 pages, 5377 KiB  
Article
Effect of Volume Fraction of Carbon Nanotubes on Structure Formation in Polyacrylonitrile Nascent Fibers: Mesoscale Simulations
by Pavel Komarov, Maxim Malyshev, Pavel Baburkin and Daria Guseva
ChemEngineering 2024, 8(5), 97; https://doi.org/10.3390/chemengineering8050097 - 26 Sep 2024
Viewed by 618
Abstract
We present a mesoscale model and the simulation results of a system composed of polyacrylonitrile (PAN), carbon nanotubes (CNTs), and a mixed solvent of dimethylsulfoxide (DMSO) and water. The model describes a fragment of a nascent PAN/CNT composite fiber during coagulation. This process [...] Read more.
We present a mesoscale model and the simulation results of a system composed of polyacrylonitrile (PAN), carbon nanotubes (CNTs), and a mixed solvent of dimethylsulfoxide (DMSO) and water. The model describes a fragment of a nascent PAN/CNT composite fiber during coagulation. This process represents one of the stages in the production of PAN composite fibers, which are considered as precursors for carbon fibers with improved properties. All calculations are based on dynamic density functional theory. The results obtained show that the greatest structural heterogeneity of the system is observed when water dominates in the composition of the mixed solvent, which is identified with the conditions of a non-solvent coagulation bath. The model also predicts that the introduction of CNTs can lead to an increase in structural heterogeneity in the polymer matrix with increasing water content in the system. In addition, it is shown that the presence of a surface modifier on the CNT surface, which increases the affinity of the filler to the polymer, can sufficiently reduce the inhomogeneity of the nascent fiber structure. Full article
(This article belongs to the Special Issue Engineering of Carbon-Based Nano/Micromaterials)
Show Figures

Figure 1

Figure 1
<p>Sketch of the dry-jet-wet spinning process and the interpretation of the internal states in the simulation cell. It can be assumed that fragments of a nascent fiber are located at different distances from the surface. Thus, the farther the simulation cell is from the surface, the greater the volume fraction of dimethyl sulfoxide (DMSO) it contains. As an alternative, it can be assumed that the simulation cell is located near the fiber surface. Therefore, the change in the DMSO/water ratio can be considered as (I) a quasi-stationary state of the system at different stages of the coagulation process or (II) the effect of the composition of the coagulation bath on the fiber structure at a fixed distance from the center. The cubic box shows an example of the polymer density distribution obtained in the simulations (blue color corresponds to the lowest density, dark red to the highest density).</p>
Full article ">Figure 2
<p>The principle of mapping the atomistic structure of PAN, DMSO, and water to coarse-grained representations. Particles of type P correspond to a segment of the PAN chain containing the comonomers acrylonitrile and itaconic acid (not explicitly shown), D—DMSO molecules, W—water, and C—a piece of MWCNT. Colors of atoms in atomistic models: carbon—gray; oxygen—red; sulfur—yellow; nitrogen—blue; hydrogen—white. The colors of the coarse-grained particles also correspond to the colors of the density fields ρ<sub>α</sub> in other figures. The parameters <span class="html-italic">N</span><sub>P</sub> and <span class="html-italic">N</span><sub>C</sub> denote the number of coarse-grained particles in the polymer chain and CNT models, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) Free energy density <span class="html-italic">F</span>, (<b>b</b>) order parameters Λ<sub>α</sub> (P—PAN, D—DMSO, W—water, C—CNTs), as a function of time for systems without CNTs (<span class="html-italic">C</span><sub>P</sub> = 80 vol%, <span class="html-italic">f</span><sub>W</sub> = 0.5, dotted lines), in the presence of CNTs (<span class="html-italic">C</span><sub>P</sub> = 75 vol%, <span class="html-italic">C</span><sub>C</sub> = 5 vol%, <span class="html-italic">f</span><sub>W</sub> = 0.5, solid lines). Instantaneous snapshots show the PAN density field distribution ρ<sub>P</sub> ≥ 0.75 (red color) on the cross-section of the simulation cell: (<b>c</b>) without filler and (<b>d</b>) with filler. The values of the time points, <span class="html-italic">t</span>, when they were obtained are given between snapshots. The completion time of the polymer/water phase separation, <span class="html-italic">t</span><sub>PS</sub>, is determined by linear extrapolation of the region of a sharp decrease of <span class="html-italic">F</span>(<span class="html-italic">t</span>) up to the intersection with the time axis (shown by the dotted line), which is schematically shown in part (<b>a</b>) of this figure. <span class="html-italic">T</span> = 300 K.</p>
Full article ">Figure 4
<p>Distribution profiles of the fraction of PAN domains, <span class="html-italic">N</span><sub>P</sub>(ρ<sub>P</sub>,<span class="html-italic">t</span>)/<span class="html-italic">N</span><sub>total</sub>, in the simulation cell with density ρ<sub>P</sub> at different simulation times and filler contents. <span class="html-italic">N</span><sub>P</sub>(ρ<sub>P</sub>,<span class="html-italic">t</span>) is the number of nodes with density equal to ρ<sub>P</sub>, and <span class="html-italic">N</span><sub>total</sub> is the total number of grid nodes in the simulation cell). <span class="html-italic">C</span><sub>P</sub> = (80 − <span class="html-italic">C</span><sub>C</sub>) vol%; <span class="html-italic">f</span><sub>W</sub> = 0.5. <span class="html-italic">N</span><sub>total</sub> is the total number of grid points in the modeling cell.</p>
Full article ">Figure 5
<p>Density field distribution ρ<sub>α</sub>(<b>r</b>,<span class="html-italic">t</span><sub>max</sub>) for (<b>a</b>) PAN (red, ρ<sub>P</sub> &gt; 0.75; green color shows the surface of pores in a polymer matrix); (<b>b</b>) DMSO (green, ρ<sub>D</sub> &gt; 0.02) and water (blue, ρ<sub>W</sub> &gt; 0.8); (<b>c</b>) CNT (black, ρ<sub>C</sub> &gt; 0.8); and (<b>d</b>) combined plot (PAN—dark and bright red [bright red—regions with high-density ρ<sub>P</sub> &gt; 1.3], DMSO—green, water—blue, CNT—black). <span class="html-italic">C</span><sub>P</sub> = 75 vol%, <span class="html-italic">C</span><sub>C</sub> = 5 vol%, <span class="html-italic">f</span><sub>W</sub> = 0.5, <span class="html-italic">T</span> = 300 K, <span class="html-italic">t</span><sub>max</sub> = 200 Δ<span class="html-italic">t</span>. Letters correspond to model component designations shown in <a href="#ChemEngineering-08-00097-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 6
<p>PAN/water phase separation time as a function of the CNT content and the mixed solvent composition in the system. <span class="html-italic">C</span><sub>P</sub> = 80−75 vol%, <span class="html-italic">C</span><sub>C</sub> = 0−5 vol%, <span class="html-italic">T</span> = 300 K.</p>
Full article ">Figure 7
<p>Visualization of the density field distribution: PAN + CNT (ρ<sub>P</sub> &gt; 1.3, ρ<sub>C</sub> &gt; 0.8) and water (ρ<sub>W</sub> &gt; 0.8) at different values of <span class="html-italic">f</span><sub>W</sub> and volume fraction of filler in the system. The following ranges of <span class="html-italic">f</span><sub>W</sub> are conventionally denoted by Roman numerals: (I) PAN forms a homogeneous structure in the absence of filler, (II) water forms discrete spherical domains, (III) water forms elongated domains, and (IV) water domains form a percolating network of pores. Letters correspond to model component designations shown in <a href="#ChemEngineering-08-00097-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 8
<p>Distribution profiles of the fraction of domains, <span class="html-italic">N</span><sub>α</sub>(ρ)/<span class="html-italic">N</span><sub>total</sub> ≡ <span class="html-italic">N</span><sub>α</sub>(ρ,<span class="html-italic">t</span><sub>max</sub>)/<span class="html-italic">N</span><sub>total</sub> (α = P, C, W), in the simulation cell with density ρ at different filler contents and <span class="html-italic">f</span><sub>w</sub> for: (<b>a</b>) PAN, (<b>b</b>) CNT, and (<b>c</b>) water. <span class="html-italic">C</span><sub>P</sub> = (80 − <span class="html-italic">C</span><sub>C</sub>) vol%. <span class="html-italic">N</span><sub>total</sub> is the total number of grid points in the simulation cell.</p>
Full article ">Figure 9
<p>The average number of (<b>a</b>) high-density (ρ<sub>P</sub> &gt; 1.3) areas in the PAN matrix &lt;<span class="html-italic">N</span><sub>P</sub>&gt; and (<b>b</b>) their average radius &lt;<span class="html-italic">R</span><sub>P</sub>&gt; as a function of the water content <span class="html-italic">f</span><sub>W</sub> and the volume fraction of filler in the system. <span class="html-italic">C</span><sub>P</sub> = 80−75 vol%, <span class="html-italic">C</span><sub>C</sub> = 0−5 vol%, <span class="html-italic">f</span><sub>W</sub> = 0−1, <span class="html-italic">T</span> = 300 K.</p>
Full article ">Figure 10
<p>The average number of (<b>a</b>) ρ<sub>C</sub> &gt; 0.8 areas formed by the filler and (<b>b</b>) their average radius &lt;<span class="html-italic">R</span><sub>C</sub>&gt; as a function of the water content <span class="html-italic">f</span><sub>W</sub> and the volume fraction of the filler in the system. <span class="html-italic">C</span><sub>P</sub> = 80−75 vol%, <span class="html-italic">C</span><sub>C</sub> = 0−5 vol%, <span class="html-italic">f</span><sub>W</sub> = 0−1, <span class="html-italic">T</span> = 300 K.</p>
Full article ">Figure 11
<p>(<b>a</b>) Average number &lt;<span class="html-italic">N</span>&gt; and (<b>b</b>) average radius &lt;<span class="html-italic">R</span>&gt; of the nuclei of the crystalline phase (ρ<sub>P</sub> &gt; 1.3), and high-density area formed by the filler (ρ<sub>C</sub> &gt; 0.8) as a function of CNT solubility parameter. <span class="html-italic">C</span><sub>P</sub> = 77 vol%, <span class="html-italic">C</span><sub>C</sub> = 3 vol%, <span class="html-italic">f</span><sub>W</sub> = 0.8.</p>
Full article ">
8 pages, 2371 KiB  
Short Note
Bis [4,4′-(1,3-Phenylenebis(azanylylidene))-bis(3,6-di-tert-butyl-2-oxycyclohexa-2,5-dien-1-one)-bis(dimethylsulfoxide)nickel(II)]
by Irina N. Meshcheryakova, Nikolay O. Druzhkov, Ilya A. Yakushev, Kseniya V. Arsenyeva, Anastasiya V. Klimashevskaya and Alexandr V. Piskunov
Molbank 2024, 2024(4), M1890; https://doi.org/10.3390/M1890 - 26 Sep 2024
Viewed by 345
Abstract
A new cage-like dimeric nickel(II) complex Ni2L2(DMSO)4 based on a ditopic redox-active hydroxy-para-iminobenzoquinone type ligand LH2 (L is 4,4′-(1,3-phenylene-bis(azaneylylidene))-bis(3,6-di-tert-butyl-2-oxycyclohexa-2,5-dien-1-one dianion) was synthesized in DMSO at 120 °C. The molecular structure of [...] Read more.
A new cage-like dimeric nickel(II) complex Ni2L2(DMSO)4 based on a ditopic redox-active hydroxy-para-iminobenzoquinone type ligand LH2 (L is 4,4′-(1,3-phenylene-bis(azaneylylidene))-bis(3,6-di-tert-butyl-2-oxycyclohexa-2,5-dien-1-one dianion) was synthesized in DMSO at 120 °C. The molecular structure of the synthesized compound was determined by X-ray diffraction analysis. The complex Ni2L2(DMSO)4 is almost insoluble in all organic solvents, probably due to the presence of a large number of intermolecular contacts in its structure. The electronic spectrum and thermal stability of the crystalline compound have been studied. Full article
Show Figures

Figure 1

Figure 1
<p>Molecular structure of dimeric complex <b>Ni<sub>2</sub>L<sub>2</sub>(DMSO)<sub>4</sub></b>·2DMSO. Solvated DMSO molecules and all hydrogen atoms are omitted for clarity. Thermal ellipsoids of 50% probability are given. Basic bond lengths (Å) and angles (°) are: Ni(1)-O(1) 2.013(3); Ni(1)-O(2) 2.077(3); Ni(1)-O(3A) 2.086(3); Ni(1)-O(4A) 2.017(3); Ni(1)-O(5) 2.050(3); Ni(1)-O(6) 2.059(3); O(1)-C(1) 1.303(5); O(2)-C(6) 1.245(5); O(3)-C(24) 1.239(5); O(4)-C(25) 1.299(5); N(1)-C(3) 1.272(6); N(1)-C(15) 1.423(7); N(2)-C(21) 1.315(6); N(2)-C(17) 1.423(6); O(1)-Ni(1)-O(4A) 90.27(14); O(1)-Ni(1)-O(5) 173.37(14); O(4A)-Ni(1)-O(5) 95.99(14); O(1)-Ni(1)-O(6) 91.56(12); O(4A)-Ni(1)-O(6) 98.55(13); O(5)-Ni(1)-O(6) 85.35(12); O(1)-Ni(1)-O(2) 78.67(13); O(4A)-Ni(1)-O(2) 161.80(13); O(5)-Ni(1)-O(2) 95.81(14); O(6)-Ni(1)-O(2) 96.18(13); O(1)-Ni(1)-O(3A) 87.98(13); O(4A)-Ni(1)-O(3A) 78.62(12); O(5)-Ni(1)-O(3A) 95.39(12); O(6)-Ni(1)-O(3A) 177.12(13); O(2)-Ni(1)-O(3A) 86.52(13).</p>
Full article ">Figure 2
<p>The fragment of crystal packing of <b>Ni<sub>2</sub>L<sub>2</sub>(DMSO)<sub>4</sub></b>·2DMSO. Solvated DMSO molecules and all hydrogen atoms are omitted for clarity. Color code: Ni, green; N, blue; O, red; S, yellow; C, grey.</p>
Full article ">Figure 3
<p>TG curve of <b>Ni<sub>2</sub>L<sub>2</sub>(DMSO)<sub>4</sub></b>.</p>
Full article ">Figure 4
<p>Electronic absorption spectrum of Nujol mull of complex <b>Ni<sub>2</sub>L<sub>2</sub>(DMSO)<sub>4</sub></b>.</p>
Full article ">Scheme 1
<p>Synthesis of the dimeric complex <b>Ni<sub>2</sub>L<sub>2</sub>(DMSO)<sub>4</sub></b>.</p>
Full article ">
15 pages, 3580 KiB  
Article
NIR-Sensitive Squaraine Dye—Peptide Conjugate for Trypsin Fluorogenic Detection
by Priyanka Balyan, Shekhar Gupta, Sai Kiran Mavileti, Shyam S. Pandey and Tamaki Kato
Biosensors 2024, 14(10), 458; https://doi.org/10.3390/bios14100458 - 25 Sep 2024
Cited by 1 | Viewed by 513
Abstract
Trypsin enzyme has gained recognition as a potential biomarker in several tumors, such as colorectal, gastric, and pancreatic cancer, highlighting its importance in disease diagnosis. In response to the demand for rapid, cost-effective, and real-time detection methods, we present an innovative strategy utilizing [...] Read more.
Trypsin enzyme has gained recognition as a potential biomarker in several tumors, such as colorectal, gastric, and pancreatic cancer, highlighting its importance in disease diagnosis. In response to the demand for rapid, cost-effective, and real-time detection methods, we present an innovative strategy utilizing the design and synthesis of NIR-sensitive dye–peptide conjugate (SQ-3 PC) for the sensitive and selective monitoring of trypsin activity by fluorescence ON/OFF sensing. The current research deals with the design and synthesis of three unsymmetrical squaraine dyes SQ-1, SQ-2, and SQ-3 along with a dye–peptide conjugate SQ-3-PC as a trypsin-specific probe followed by their photophysical characterizations. The absorption spectral investigation conducted on both the dye alone and its corresponding dye–peptide conjugates in water, utilizing SQ-3 and SQ-3 PC respectively, reveals enhanced dye aggregation and pronounced fluorescence quenching compared to observations in DMSO solution. The absorption spectral investigation conducted on dye only and corresponding dye–peptide conjugates in water utilizing SQ-3 and SQ-3 PC, respectively, reveals not only the enhanced dye aggregation but also pronounced fluorescence quenching compared to that observed in the DMSO solution. The trypsin-specific probe SQ-3 PC demonstrated a fluorescence quenching efficiency of 61.8% in water attributed to the combined effect of aggregation-induced quenching (AIQ) and fluorescence resonance energy transfer (FRET). FRET was found to be dominant over AIQ. The trypsin-mediated hydrolysis of SQ-3 PC led to a rapid and efficient recovery of quenched fluorescence (5-fold increase in 30 min). Concentration-dependent changes in the fluorescence at the emission maximum of the dyes reveal that SQ-3 PC works as a trypsin enzyme-specific fluorescence biosensor with linearity up to 30 nM along with the limit of detection and limit of quantification of 1.07 nM and 3.25 nM, respectively. Full article
(This article belongs to the Special Issue Photonics for Bioapplications: Sensors and Technology)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of squaraine dyes (<b>SQ-1</b>, <b>SQ-2</b>, and <b>SQ-3</b>).</p>
Full article ">Figure 2
<p>Chemical structure of (<b>a</b>) trypsin enzyme-specific peptide and (<b>b</b>) dye–peptide conjugate probe (<b>SQ-3 PC</b>) for the trypsin enzyme sensing.</p>
Full article ">Figure 3
<p>(<b>a</b>) Optical absorption and (<b>b</b>) fluorescence emission spectra of <b>SQ-1</b>, <b>SQ-2</b>, <b>SQ-3</b>, and <b>SQ-3 PC</b> in DMSO (5 μM).</p>
Full article ">Figure 4
<p>(<b>a</b>) Absorption and (<b>b</b>) emission Spectra of <b>SQ-3</b> and <b>SQ-3 PC</b> (5 μM) in H<sub>2</sub>O (2% DMSO) solvent.</p>
Full article ">Figure 5
<p>Schematic representation of H- and J-aggregate formation in squaraine dyes.</p>
Full article ">Figure 6
<p>(<b>a</b>) Normalized absorption and emission spectra of <b>SQ-3</b> in DMSO and (<b>b</b>) optimized molecular structure of <b>SQ-3 PC</b> calculated using Gaussian G16 program.</p>
Full article ">Figure 7
<p>(<b>a</b>) Concentration-dependent fluorescence spectra of <b>SQ-3 PC</b> (5 µM) in H<sub>2</sub>O (2% DMSO) for 30 min with the addition of 0 to 75 nM Trypsin, (<b>b</b>) time-dependent fluorescence spectra of <b>SQ-3PC</b> (5 µM) in H<sub>2</sub>O (2% DMSO) for 30 min with the addition of 75 nM of trypsin enzyme, (<b>c</b>) fluorescence spectra of <b>SQ-3 PC</b> with change in fluorescence intensity as a function of time with different concentrations of enzyme for a fixed concentration of <b>SQ-3 PC</b> (5 μM), and (<b>d</b>) plot of ratio of fluorescence intensity (F/F<sub>0</sub>) as a function of enzyme concentration.</p>
Full article ">Figure 8
<p>Fluorescence response (F/F<sub>0</sub>) of the probe <b>SQ-3 PC</b> (5 μM) in H<sub>2</sub>O (2% DMSO) towards different potential and competing enzymes (75 nM).</p>
Full article ">Scheme 1
<p>Schematic representation for the synthesis of dye–peptide conjugate (<b>SQ-3 PC</b>); R, Rink Amide MBHA Resin; PG<sup>1</sup>, Boc; PG<sup>2</sup>, o’tBu; PG<sup>3</sup>, Pbf; PG<sup>4</sup>, trt.</p>
Full article ">
15 pages, 879 KiB  
Article
Hormone-Driven Temperature Optimization for Elevated Reproduction in Goldfish (Carassius auratus) under Laboratory Conditions
by Zeynab Taheri-Khas, Ahmad Gharzi, Somaye Vaissi, Pouria Heshmatzad and Zahra Kalhori
Animals 2024, 14(18), 2701; https://doi.org/10.3390/ani14182701 - 18 Sep 2024
Viewed by 416
Abstract
This study investigates the efficacy of hormone-induced artificial reproduction in goldfish (Carassius auratus) under controlled temperatures. Ovaprim injections significantly enhanced ovulation and sperm production compared to controls. Medium temperature (22 °C) produced the highest ovulation rates, fastest ovulation timing, and optimal [...] Read more.
This study investigates the efficacy of hormone-induced artificial reproduction in goldfish (Carassius auratus) under controlled temperatures. Ovaprim injections significantly enhanced ovulation and sperm production compared to controls. Medium temperature (22 °C) produced the highest ovulation rates, fastest ovulation timing, and optimal sperm quality (motility and morphology) compared to high (28 °C) and low (16 °C) temperature groups. The low-temperature group exhibited reduced sperm motility duration and higher rates of sperm and larvae damage. The sperm volume of the high-temperature group was higher, but their post-injection survival rates were lower. Furthermore, the lowest spawning rate and low egg quality were noted in the high temperature. Cryopreservation using extender E4 (15% DMSO) exhibited superior post-thaw sperm motility and achieved higher fertilization rates. Fertilization rates, embryo development, and larval survival were all highest at the medium temperature. Larvae hatched from fresh sperm at medium temperature exhibited faster growth and fewer deformities. These findings suggest that hormone stimulation coupled with a medium temperature regimen is critical for successful artificial reproduction in goldfish. Cryopreservation with extender E4 holds promise for sperm banking; however, further optimization is necessary to improve fertilization success with thawed sperm. Future research could explore the influence of temperature on sperm physiology and refine cryopreservation protocols to enhance fertilization rates. Full article
(This article belongs to the Special Issue Animal Reproduction: Semen Quality Assessment, Volume II)
Show Figures

Figure 1

Figure 1
<p>Percentage of goldfish (<span class="html-italic">Carassius auratus</span>) sperm motility at various temperatures. The temperatures are categorized as follows: 28 ± 1 °C (HT); 22 ± 1 °C (MT); and 16 ± 1 °C (LT). The values are mean ± SD. (*) <span class="html-italic">p</span>-value &lt; 0.05; (**) <span class="html-italic">p</span>-value &lt; 0.00<span style="lang:ar">1</span>.</p>
Full article ">Figure 2
<p>Sperm morphology of goldfish (<span class="html-italic">Carassius auratus</span>) in different temperature treatments. The temperatures are categorized as follows: 16 ± 1 °C (LT); 22 ± 1 °C (MT); and, 28 ± 1 °C (HT). (a) Head wrinkled; (b) detached head; (c) bent tail; (d) normal sperm; (e) coiled tail. (40× magnification).</p>
Full article ">
15 pages, 1519 KiB  
Article
Caffeine—Legal Natural Stimulant with Open Research Perspective: Spectroscopic and Theoretical Characterization
by Teobald Kupka, Natalina Makieieva, Michał Jewgiński, Magdalena Witek, Barbara Blicharska, Oimahmad Rahmonov, Karel Doležal and Tomáš Pospíšil
Molecules 2024, 29(18), 4382; https://doi.org/10.3390/molecules29184382 - 14 Sep 2024
Viewed by 1018
Abstract
Caffeine is an alkaloid with a purine structure and has been well known for centuries due to its presence in popular drinks—tea and coffee. However, the structural and spectroscopic parameters of this compound, as well as its chemical and biological activities, are still [...] Read more.
Caffeine is an alkaloid with a purine structure and has been well known for centuries due to its presence in popular drinks—tea and coffee. However, the structural and spectroscopic parameters of this compound, as well as its chemical and biological activities, are still not fully known. In this study, for the first time, we report on the measured oxygen-17 NMR spectra of this stimulant. To support the assignment of our experimental NMR data, extensive quantum chemical calculations of NMR parameters, including nuclear magnetic shielding constants and indirect spin–spin coupling constants, were performed. In a theoretical study, using nine efficient density functionals (B3LYP, BLYP, BP86, CAM-B3LYP, LC-BLYP, M06, PBE0, TPSSh, wB97x), and in combination with a large and flexible correlation-consistent aug-cc-pVTZ basis set, the structure and NMR parameters were predicted for a free molecule of caffeine and in chloroform, DMSO and water. A polarized continuum model (PCM) was used to include a solvent effect. As a result, an optimal methodology was developed for predicting reliable NMR data, suitable for studies of known, as well as newly discovered, purines and similar alkaloids. The results of the current work could be used in future basic and applied studies, including NMR identification and intermolecular interactions of caffeine in various raw materials, like plants and food, as well as in the structural and spectroscopic characterization of new compounds with similar structures. Full article
(This article belongs to the Section Bioorganic Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Caffeine structure with atom labeling.</p>
Full article ">Figure 2
<p><sup>17</sup>O NMR spectrum of caffeine in CDCl<sub>3</sub> measured at 35 °C (100 Hz line broadening).</p>
Full article ">Figure 3
<p>RMS deviations (in ppm) of chemical shifts in (<b>A</b>) <sup>13</sup>C, (<b>B</b>) <sup>1</sup>H, (<b>C</b>) <sup>15</sup>N and (<b>D</b>) <sup>17</sup>O calculated with selected density functionals for free caffeine in vacuum and in solution (numbers are assigned to used functionals in computational part).</p>
Full article ">Figure 3 Cont.
<p>RMS deviations (in ppm) of chemical shifts in (<b>A</b>) <sup>13</sup>C, (<b>B</b>) <sup>1</sup>H, (<b>C</b>) <sup>15</sup>N and (<b>D</b>) <sup>17</sup>O calculated with selected density functionals for free caffeine in vacuum and in solution (numbers are assigned to used functionals in computational part).</p>
Full article ">Figure 4
<p>RMS deviations of DFT calculated caffeine’s <sup>1</sup>J(C-H) in different environments from experiment [<a href="#B25-molecules-29-04382" class="html-bibr">25</a>] (numbers are assigned to used functionals in computational part).</p>
Full article ">
15 pages, 5427 KiB  
Article
Polymorph Screening of Core-Chlorinated Naphthalene Diimides with Different Fluoroalkyl Side-Chain Lengths
by Inês de Oliveira Martins, Marianna Marchini, Lucia Maini and Enrico Modena
Molecules 2024, 29(18), 4376; https://doi.org/10.3390/molecules29184376 - 14 Sep 2024
Viewed by 443
Abstract
In this work, naphthalenediimide (NDI) derivatives are widely studied for their semiconducting properties and the influence of the side-chain length on the crystal packing is reported, along with the thermal properties of three core-chlorinated NDIs with different fluoroalkyl side-chain lengths (CF3-NDI, [...] Read more.
In this work, naphthalenediimide (NDI) derivatives are widely studied for their semiconducting properties and the influence of the side-chain length on the crystal packing is reported, along with the thermal properties of three core-chlorinated NDIs with different fluoroalkyl side-chain lengths (CF3-NDI, C3F7-NDI and C4F9-NDI). The introduction of fluorinated substituents at the imide nitrogen and addition of strong electron-withdrawing groups at the NDI core are used to improve the NDI derivatives air stability. The new compound, CF3-NDI, was deeply analyzed and compared to the well-known C3F7-NDI and C4F9-NDI, leading to the discovery and solution of two different crystal phases, form α and solvate form, and a solid solution of CF3-NDI and CF3-NDI-OH, formed by the decomposition in DMSO. Full article
(This article belongs to the Special Issue Covalent and Noncovalent Interactions in Crystal Chemistry II)
Show Figures

Figure 1

Figure 1
<p>Crystal habits of CF<sub>3</sub>-NDI crystal forms.</p>
Full article ">Figure 2
<p>Packing of NDI derivative core along <span class="html-italic">a</span> axis (<b>a</b>). View of layer separation of CF<sub>3</sub>-NDI Form α (<b>b</b>) and interdigitation of the molecules of C<sub>3</sub>F<sub>7</sub>-NDI Form α (<b>c</b>) and C<sub>4</sub>F<sub>9</sub>-NDI Form α (<b>d</b>).</p>
Full article ">Figure 3
<p>Alternated slip-stack of NDI and PXY molecule (<b>a</b>) and view of layer separation of CF<sub>3</sub>-NDI·PXY (<b>b</b>).</p>
Full article ">Figure 4
<p>π–π stacking plane distance (<b>a</b>) and crystal packing of CF<sub>3</sub>-NDI·SS (<b>b</b>); only the CF<sub>3</sub>-NDI-OH are shown, but it should be considered that 20% of the solid solution has only Cl in the core.</p>
Full article ">Figure 5
<p>Left, TGA curve of CF<sub>3</sub>-NDI·PXY, which shows the loss of solvent; right, pictures of crystal before (orange crystal) and after (yellow crystal) desolvation.</p>
Full article ">Figure 6
<p>DSC analysis of C<sub>3</sub>F<sub>7</sub>-NDI, showing the presence of melting and crystallization followed by a complete melting, 2 °C/min).</p>
Full article ">Figure 7
<p>Pictures from hot stage microscopy of C<sub>3</sub>F<sub>7</sub>-NDI: (<b>a</b>) initial state of C<sub>3</sub>F<sub>7</sub>-NDI·Form α crystal, (<b>b</b>) expanded crystal after thermal expansion and respective dimension, (<b>c</b>) melting and crystallization at 302 °C, (<b>d</b>) complete melting at 304 °C, (<b>e</b>) recrystallization at 284 °C and (<b>f</b>) 251 °C.</p>
Full article ">Figure 8
<p>VT-XRPD ofC<sub>3</sub>F<sub>7</sub>-NDI in heating and cooling (<b>top</b>), showing the same behavior for both since the thermal expansion of C<sub>3</sub>F<sub>7</sub>-NDI is reversible, and (<b>bottom</b>) C<sub>4</sub>F<sub>9</sub>-NDI, which shows the non-reversible transition from C<sub>4</sub>F<sub>9</sub>-NDI Form α to C<sub>4</sub>F<sub>9</sub>-NDI Form β.</p>
Full article ">Figure 9
<p>XRPD curve of C<sub>4</sub>F<sub>9</sub>-NDI Form γ transition from C<sub>4</sub>F<sub>9</sub>-NDI Form γ to C<sub>4</sub>F<sub>9</sub>-NDI Form α.</p>
Full article ">Scheme 1
<p>Molecular structure of studied NDI derivatives CF<sub>3</sub>-NDI (4,5,9,10-tetrachloro-2,7-bis (2,2,2-trifluoroethyl) benzo[lmn][3,8] phenanthroline-1,3,6,8(2H,7H)-tetraone), C<sub>3</sub>F<sub>7</sub>-NDI, and C<sub>4</sub>F<sub>9</sub>-NDI.</p>
Full article ">
17 pages, 3647 KiB  
Article
Profoxydim in Focus: A Structural Examination of Herbicide Behavior in Gas and Aqueous Phases
by María Cobos-Escudero, Paula Pla, Álvaro Cervantes-Diaz, José Luis Alonso-Prados, Pilar Sandín-España, Manuel Alcamí and Al Mokhtar Lamsabhi
Molecules 2024, 29(18), 4371; https://doi.org/10.3390/molecules29184371 - 14 Sep 2024
Viewed by 410
Abstract
This study investigates the chemical structure of profoxydim, focusing on its E–isomer, the main commercial form. The research aimed to determine the predominant tautomeric forms under various environmental conditions. Using proton and carbon–13 NMR spectroscopy alongside theoretical modeling, we examined tautomers and their [...] Read more.
This study investigates the chemical structure of profoxydim, focusing on its E–isomer, the main commercial form. The research aimed to determine the predominant tautomeric forms under various environmental conditions. Using proton and carbon–13 NMR spectroscopy alongside theoretical modeling, we examined tautomers and their conformers in different solvents (MeOD, DMSO, CDCl3, benzene) to mimic gas and aqueous phases. The findings reveal that the enolic form dominates in the gas phase, while the ketonic form prevails in aqueous environments, providing key insights into the herbicide’s environmental behavior. We also observed an isomeric transition from E to Z under acidic conditions, which could affect profoxydim’s reactivity in natural environments. The theoretical calculations indicated that in acidic conditions, the E and Z forms are nearly degenerate, with the E form remaining dominant in neutral environments. Additionally, QSAR models assessed the toxicity of various tautomers, revealing significant differences that could impact bioactivity and environmental fate. This research offers crucial insights into the structural dynamics of profoxydim, contributing to cyclohexanedione chemistry and the development of more effective herbicides. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of the different profoxydim’s tautomers and isomers considered in this study.</p>
Full article ">Figure 2
<p>Relative stability of all the tautomer within the range of 5 kcal/mol.</p>
Full article ">Figure 3
<p>Optimized molecular geometries of the most stable conformers of each profoxydim’s tautomer. Relative free energy with respect to the most stable structure (<span class="html-italic">E</span>–diketoenamine–RS/SR) obtained in water as a solvent is given in kcal/mol. Hydrogen bonds are highlighted with a magenta oval. Color coding of atoms: hydrogen (white), carbon (gray), nitrogen (blue), oxygen (red), chlorine (green) and sulfur (yellow).</p>
Full article ">Figure 4
<p>NCI analysis of the most stable conformer of each isomer of profoxydim in water.</p>
Full article ">Figure 5
<p>Superimposed chromatograms of the evolution of <span class="html-italic">E</span>/<span class="html-italic">Z</span> isomer of profoxydim at pH = 3.</p>
Full article ">Figure 6
<p>Positions considered for protonation in profoxydim molecules.</p>
Full article ">Figure 7
<p>Relative stability of all the tautomers within the range of 25 kcal/mol (30 kcal/mol for gas and octanol).</p>
Full article ">Figure 8
<p>The optimized molecular geometries of the most stable protomers of the <span class="html-italic">E</span> compared with <span class="html-italic">Z</span> counterpart. Hydrogen bonds are highlighted with a magenta. Color coding of atoms: hydrogen (white), carbon (gray), nitrogen (blue), oxygen (red), chlorine (green) and sulfur (yellow).</p>
Full article ">
15 pages, 3771 KiB  
Article
An Acid-Responsive Fluorescent Molecule for Erasable Anti-Counterfeiting
by Jiabao Liu, Xiangyu Gao, Qingyu Niu, Mingyuan Jin, Yijin Wang, Thamraa Alshahrani, He-Lue Sun, Banglin Chen, Zhiqiang Li and Peng Li
Molecules 2024, 29(18), 4335; https://doi.org/10.3390/molecules29184335 - 12 Sep 2024
Viewed by 619
Abstract
A tetraphenylethylene (TPE) derivative, TPEPhDAT, modified by diaminotriazine (DAT), was prepared by successive Suzuki–Miyaura coupling and ring-closing reactions. This compound exhibits aggregation-induced emission enhancement (AIEE) properties in the DMSO/MeOH system, with a fluorescence emission intensity in the aggregated state that is 5-fold higher [...] Read more.
A tetraphenylethylene (TPE) derivative, TPEPhDAT, modified by diaminotriazine (DAT), was prepared by successive Suzuki–Miyaura coupling and ring-closing reactions. This compound exhibits aggregation-induced emission enhancement (AIEE) properties in the DMSO/MeOH system, with a fluorescence emission intensity in the aggregated state that is 5-fold higher than that of its counterpart in a dilute solution. Moreover, the DAT structure of the molecule is a good acceptor of protons; thus, the TPEPhDAT molecule exhibits acid-responsive fluorescence. TPEPhDAT was protonated by trifluoroacetic acid (TFA), leading to fluorescence quenching, which was reversibly restored by treatment with ammonia (on–off switch). Time-dependent density functional theory (TDDFT) computational studies have shown that protonation enhances the electron-withdrawing capacity of the triazine nucleus and reduces the bandgap. The protonated TPEPhDAT conformation became more distorted, and the fluorescence lifetime was attenuated, which may have produced a twisted intramolecular charge transfer (TICT) effect, leading to fluorescence redshift and quenching. MeOH can easily remove the protonated TPEPhDAT, and this acid-induced discoloration and erasable property can be applied in anti-counterfeiting. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) In the DMSO/MeOH solvent mixture, the photograph of the volume fraction of MeOH (<span class="html-italic">f</span><sub>MeOH</sub>) increasing from 0 vol% to 95 vol% (under UV light); (<b>b</b>) fluorescence spectra of TPEPhDAT in DMSO/MeOH mixtures with different <span class="html-italic">f</span><sub>MeOH</sub> (concentration: 10 μM; <span class="html-italic">λ</span><sub>ex</sub>: 380 nm); (<b>c</b>) plot of relative PL intensity (<span class="html-italic">I</span>/<span class="html-italic">I</span><sub>0</sub>) vs <span class="html-italic">f</span><sub>MeOH</sub>; (<b>d</b>) time-resolved decay curves of solution (<span class="html-italic">f</span><sub>MeOH</sub> = 0 vol%) and aggregate (<span class="html-italic">f</span><sub>MeOH</sub> = 95 vol%) at <span class="html-italic">λ</span><sub>ex</sub> = 380 nm; (<b>e</b>) hydrodynamic radius distribution of aggregates when <span class="html-italic">f</span><sub>MeOH</sub> = 95 vol%.</p>
Full article ">Figure 2
<p>(<b>a</b>) Pictures of acid-responsive fluorescence color changes (2 mg/mL aqueous solution, under UV light) of TPEPhDAT (i) in its original state, (ii) after dropwise addition of HNO<sub>3</sub>, and (iii) after dropwise addition of ammonium hydroxide; (<b>b</b>) fluorescence spectra of TPEPhDAT under 350 nm excitation before and after dropwise addition of HNO<sub>3</sub>; (<b>c</b>) fluorescence quenching efficiency of TPEPhDAT with the addition of different acids; (<b>d</b>) fluorescence recovery cycle of TPEPhDAT to TFA: the red solid line indicates the quenching process and the purple dashed line indicates the recovery process.</p>
Full article ">Figure 3
<p>(<b>a</b>) Partial <sup>1</sup>H NMR spectra of TPEPhDAT in DMSO-<span class="html-italic">d</span><sub>6</sub> after adding TFA and NH<sub>3</sub>·H<sub>2</sub>O; (<b>b</b>) FT-IR spectra of TPEPhDAT before and after protonation; (<b>c</b>) C−N−C internal angles for the protonated and unprotonated positions of the triazine ring (ellipsoids drawn at 50% probability).</p>
Full article ">Figure 4
<p>(<b>a</b>) Supposed sensing mechanism of TPEPhDAT to TFA; (<b>b</b>) time-resolved decay curves of TPEPhDAT after successive treatment by TFA and ammonia at <span class="html-italic">λ</span><sub>ex</sub> = 350 nm.</p>
Full article ">Figure 5
<p>TDDFT calculations for TPEPhDAT and <span class="html-italic">S</span><sub>0</sub> → <span class="html-italic">S</span><sub>1</sub> electron–hole calculations. Protonation is adopted to simulate TFA stimulation, as previously reported.</p>
Full article ">Figure 6
<p>(<b>a</b>) The “<b>HEBUT</b>” pattern is treated alternately with acid and alkali vapors (under UV light); (<b>b</b>) pictures of reversible fluorescence switching of TPEPhDAT in acidic and alkaline environments and its erasure by MeOH (under UV light).</p>
Full article ">Scheme 1
<p>Synthetic route of TPEPhDAT.</p>
Full article ">
15 pages, 2116 KiB  
Article
Addition of Cryoprotectant DMSO Reduces Damage to Spermatozoa of Yellow Catfish (Pelteobagrus fulvidraco) during Cryopreservation: Ultrastructural Damage, Oxidative Damage and DNA Damage
by Yuxin Zhang, Dongqing Liu, Qinghua Wang, Qingxin Ruan, Sijie Hua, Weiwei Zhang, Sen Yang and Zining Meng
Animals 2024, 14(18), 2652; https://doi.org/10.3390/ani14182652 - 12 Sep 2024
Viewed by 347
Abstract
Spermatozoa cryopreservation protocols have been established for yellow catfish (Pelteobagrus fulvidraco), but cryopreservation can still cause cellular damage and affect spermatozoa viability and fertility. Therefore, the aim of this paper was to evaluate the effects of adding or not adding cryoprotectants [...] Read more.
Spermatozoa cryopreservation protocols have been established for yellow catfish (Pelteobagrus fulvidraco), but cryopreservation can still cause cellular damage and affect spermatozoa viability and fertility. Therefore, the aim of this paper was to evaluate the effects of adding or not adding cryoprotectants during low-temperature storage on the ultrastructural damage, oxidative damage, and DNA damage of thawed yellow catfish spermatozoa. The mixed semen of three male yellow catfish was divided into a fresh spermatozoa group, a frozen spermatozoa group (DMSO+) with a cryoprotectant (10% DMSO), and a frozen spermatozoa group without a cryoprotectant (DMSO). Ultrastructural of the spermatozoa after thawing were observed under an electron microscope and the spermatozoa were assayed for SOD, MDA, and T-AOC enzyme activities, as well as for DNA integrity. In terms of movement parameters, compared with DMSO, the addition of DMSO has significantly improved sperm motility, curve line velocity (VCL), and straight line velocity (VSL). The ultrastructural results showed that although thawed spermatozoa exhibited increased damage than fresh spermatozoa, 10% DMSO effectively reduced the damage to the plasma membrane, mitochondria, and flagellum of spermatozoa by cryopreservation, and most of the spermatozoa were preserved with intact structure. The results of oxidative damage showed that compared with frozen spermatozoa, 10% DMSO significantly increased the activities of SOD and T-AOC enzymes and clearly reduced the activity of the MDA enzyme. The antioxidant capacity of spermatozoa was improved, lipid peroxidation was reduced, and the oxidative damage caused by cryopreservation was mitigated. The DNA integrity of spermatozoa showed that 10% DMSO clearly reduced the DNA fragmentation rate. In conclusion, 10% DMSO can effectively reduce the ultrastructural damage, oxidative damage, and DNA damage of yellow catfish spermatozoa during cryopreservation; it can also further optimize the cryopreservation protocol for yellow catfish spermatozoa. Meanwhile, it also provides a theoretical basis for the future optimization of the cryopreservation protocols. Full article
Show Figures

Figure 1

Figure 1
<p>Criteria for determining DNA fragments of spermatozoa. A: the minimum diameter of the spermatozoa head. B: width of unilateral halo. 1: Spermatozoa with intact DNA; 2: Spermatozoa with DNA fragments.</p>
Full article ">Figure 2
<p>Scanning electron micrographs of fresh and post-thaw spermatozoa. (<b>A</b>,<b>B</b>) Fresh spermatozoa images under scanning electron microscopy (H: head; MP: midpiece of spermatozoa; F: tail; S: central space of sleeve; a: Whole spermatozoa scanning electron microscopy; b: Localized scanning electron microscopy of spermatozoa); (<b>C</b>,<b>D</b>) the images of thawed spermatozoa stored in 10% DMSO under scanning electron microscopy; (<b>E</b>,<b>F</b>) images of frozen spermatozoa under scanning electron microscopy.</p>
Full article ">Figure 3
<p>Transmission electron micrographs of fresh and post-thaw spermatozoa. (<b>A</b>,<b>B</b>) Images of fresh spermatozoa under transmission electron microscopy (a, b, c: microtubule structures; M: mitochondria; (<b>C</b>,<b>D</b>) images of thawed spermatozoa stored in 10% DMSO under transmission electron microscopy (PM: plasma membrane; NM: nuclear membrane; PC: proximal centriole; BB: basal baby; OM: outer membrane of sleeve cover; IM: inner membrane of the sleeves; S: central space of the sleeve; F: flagellum; TR: transition region); (<b>E</b>,<b>F</b>) images of frozen spermatozoa under transmission electron microscopy.</p>
Full article ">Figure 4
<p>DNA fragments of fresh spermatozoa, DMSO<sup>+</sup>, and DMSO<sup>−</sup> groups of yellow catfish spermatozoa under 100× oil microscope. (<b>A</b>) DNA fragments in fresh spermatozoa; (<b>B</b>) DNA fragments in the DMSO<sup>+</sup> group; (<b>C</b>) DNA fragments in the DMSO<sup>−</sup> group.</p>
Full article ">Figure 5
<p>DNA fragmentation rate of fresh spermatozoa, DMSO<sup>+</sup>, and DMSO<sup>−</sup> groups of yellow catfish spermatozoa (%). FS: fresh spermatozoa; different superscripts refer to significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
17 pages, 2230 KiB  
Article
Myosin Isoform-Dependent Effect of Omecamtiv Mecarbil on the Regulation of Force Generation in Human Cardiac Muscle
by Beatrice Scellini, Nicoletta Piroddi, Marica Dente, J. Manuel Pioner, Cecilia Ferrantini, Corrado Poggesi and Chiara Tesi
Int. J. Mol. Sci. 2024, 25(18), 9784; https://doi.org/10.3390/ijms25189784 - 10 Sep 2024
Viewed by 517
Abstract
Omecamtiv mecarbil (OM) is a small molecule that has been shown to improve the function of the slow human ventricular myosin (MyHC) motor through a complex perturbation of the thin/thick filament regulatory state of the sarcomere mediated by binding to myosin allosteric sites [...] Read more.
Omecamtiv mecarbil (OM) is a small molecule that has been shown to improve the function of the slow human ventricular myosin (MyHC) motor through a complex perturbation of the thin/thick filament regulatory state of the sarcomere mediated by binding to myosin allosteric sites coupled to inorganic phosphate (Pi) release. Here, myofibrils from samples of human left ventricle (β-slow MyHC-7) and left atrium (α-fast MyHC-6) from healthy donors were used to study the differential effects of μmolar [OM] on isometric force in relaxing conditions (pCa 9.0) and at maximal (pCa 4.5) or half-maximal (pCa 5.75) calcium activation, both under control conditions (15 °C; equimolar DMSO; contaminant inorganic phosphate [Pi] ~170 μM) and in the presence of 5 mM [Pi]. The activation state and OM concentration within the contractile lattice were rapidly altered by fast solution switching, demonstrating that the effect of OM was rapid and fully reversible with dose-dependent and myosin isoform-dependent features. In MyHC-7 ventricular myofibrils, OM increased submaximal and maximal Ca2+-activated isometric force with a complex dose-dependent effect peaking (40% increase) at 0.5 μM, whereas in MyHC-6 atrial myofibrils, it had no effect or—at concentrations above 5 µM—decreased the maximum Ca2+-activated force. In both ventricular and atrial myofibrils, OM strongly depressed the kinetics of force development and relaxation up to 90% at 10 μM [OM] and reduced the inhibition of force by inorganic phosphate. Interestingly, in the ventricle, but not in the atrium, OM induced a large dose-dependent Ca2+-independent force development and an increase in basal ATPase that were abolished by the presence of millimolar inorganic phosphate, consistent with the hypothesis that the widely reported Ca2+-sensitising effect of OM may be coupled to a change in the state of the thick filaments that resembles the on–off regulation of thin filaments by Ca2+. The complexity of this scenario may help to understand the disappointing results of clinical trials testing OM as inotropic support in systolic heart failure compared with currently available inotropic drugs that alter the calcium signalling cascade. Full article
(This article belongs to the Special Issue Molecular Motors: Mechanical Properties and Regulation)
Show Figures

Figure 1

Figure 1
<p>Myosin heavy chain isoform composition in human ventricular and atrial myofibrils. (<b>A</b>): Representative 8% SDS-PAGE of the MyHC isoforms in human cardiac samples. HV and HA: samples used to prepare the myofibrils used in this study; cAF: surgical sample from patients with chronic atrial fibrillation used as a reference standard for identification of α-fast MyHC-6 and β-slow MyHC-7 band positions. (<b>B</b>): Representative density profile of the MyHC isoforms in the HA sample (MyHC-6: 0.83; MyHC-7: 0.18). (<b>C</b>): Relative distribution of the MyHC isoforms in HV and HA myofibrils (HV: MyHC-6: 0.04 ± 0.01, β-slow MyHC-7: 0.96 ± 0.01, <span class="html-italic">n</span> = 6; HA: MyHC-6: 0.83 ± 0.01, β-slow MyHC-7: 0.17 ± 0.01, <span class="html-italic">n</span> = 21; means ± SEMs).</p>
Full article ">Figure 2
<p>Effects of the rapid perturbation of [OM] during steady calcium-activation on tension development in human ventricular and atrial myofibrils (15 °C). (<b>A</b>): Representative trace from a jump experiment in aHV myofibril activated under control conditions and subjected to a rapid change in [OM] from 0 to 0.5 µM and back. The initial sarcomere length is 2.13 µm. The timing of the solution change is represented by the indexed bar at the top of the tension trace (top trace). Fast length changes (bottom trace) are applied to the myofibril under the condition of steady force generation (P<sub>0</sub>) under control and [OM] conditions to measure <span class="html-italic">k</span><sub>TR</sub>. (<b>B</b>): Mean values of active tension P<sub>0</sub> (dots) and <span class="html-italic">k</span><sub>TR</sub> triangles) measured from jump experiments (and normalised by the mean value of the same parameters measured before and after the ligand jump) plotted as a function of [OM]. Black symbols: jump experiments performed at maximal activation (pCa 4.50); grey symbols: jump experiments performed at submaximal activation (pCa 5.75). Error bars ± SEMs. (<b>C</b>): Representative trace from a jump experiment in a HA myofibril activated under control conditions and subjected to a rapid change in [OM] from 0 to 0.05 µM and back. The initial sarcomere length is 2.19 µm. The timing of the solution change is represented by the indexed bar at the top of the tension trace (top trace). Rapid changes in length (lower trace) are applied to the myofibril under conditions of steady force generation P<sub>0</sub> in control and OM conditions to measure <span class="html-italic">k<sub>TR</sub></span>. (<b>D</b>): Mean values of active tension P<sub>0</sub> (dots) and <span class="html-italic">k</span><sub>TR</sub> (triangles) measured from jump experiments in HA myofibrils (and normalised by the mean value of the same parameters measured before and after the ligand jump) plotted as a function of [OM]. Black symbols: jump experiments performed at maximal activation (pCa 4.50); grey symbols: jump experiments performed at submaximal activation (pCa 5.75). Error bars± SEMs.</p>
Full article ">Figure 3
<p>Effects of OM on the development of resting and calcium-activated tension in human ventricular and atrial myofibrils (15 °C). (<b>A</b>,<b>B</b>): Representative traces of tension generation in two human ventricular myofibrils activated and relaxed by fast solution change in the presence of 5 µM [OM] (A) or 5 µM [OM] and 5 mM [Pi] (<b>B</b>) in both relaxing and activating solutions; the timing of the solution change is represented by the indexed bars at the top of the tension traces (top traces); bottom traces: timing of rapid length changes applied to the myofibril to measure <span class="html-italic">k</span><sub>TR</sub>. Sarcomere length: 2.03 µm (left panel) and 2.10 µm (right panel). (<b>C</b>,<b>D</b>): Mean absolute tension values measured in activationrelaxation cycles in the presence of different [OM]s (<b>C</b>) or in the presence of different [OM]s and 5 mM Pi (<b>D</b>). The [OM]s range from 0 to 10 µM. Filled symbols: active tension generation in pCa 4.50; empty symbols: Ca<sup>2+</sup>-independent force generation in relaxing solution (pCa 9.0) in the presence of [OM]. Error bars ± SEMs. (<b>E</b>): Representative trace of tension generation (top trace) in a HA myofibril activated and relaxed by fast solution change in the presence of 10 µM [OM] in both relaxing and activating solutions. The timing of the solution change is represented by the indexed bars at the top of the tension trace. Bottom trace: rapid length changes applied to the myofibril to measure <span class="html-italic">k</span><sub>TR</sub>. Sarcomere length: 2.09 µm. (<b>F</b>): Mean absolute tension values measured in activation–relaxation cycles in control conditions and in the presence of different [OM]s up to 10 µM. Filled symbols: active tension generation in pCa 4.50; empty symbols: Ca<sup>2+</sup>-independent force generation in relaxing solution (pCa 9.0) in the presence of different [OM]s. Error bar, ±SEM.</p>
Full article ">Figure 4
<p>Effects of OM on the kinetics of tension generation and relaxation in human ventricular and atrial myofibrils (15 °C). (<b>A</b>–<b>D</b>): HV myofibrils. (<b>A</b>): Mean rate of tension generation (<span class="html-italic">k</span><sub>ACT</sub>) at different [OM]s. Filled symbols: rate of tension development induced by switching from relaxing (pCa 9.0) to activating (pCa 4.5) solutions; empty symbols: <span class="html-italic">k</span><sub>OM</sub> rate of Ca<sup>2+</sup>-independent tension development induced by OM in relaxing solution (pCa 9.0). Error bars ± SEM. (<b>B</b>–<b>D</b>): Mean parameters of force relaxation in the presence of different [OM]s. (<b>B</b>): Duration of the slow phase of relaxation. (<b>C</b>): Rate of the slow phase of relaxation <span class="html-italic">slow k</span><sub>REL</sub>. (<b>D</b>): Rate of the fast phase of relaxation <span class="html-italic">fast k</span><sub>REL</sub>. (<b>E</b>–<b>H</b>): As above for HA myofibrils. (<b>E</b>): Mean values of <span class="html-italic">k</span><sub>ACT</sub> or <span class="html-italic">k</span><sub>OM</sub>. (<b>F</b>): Duration of the slow phase of relaxation. (<b>G</b>): <span class="html-italic">slow k</span><sub>REL</sub> and H: <span class="html-italic">fast k</span><sub>REL</sub> in the presence of different [OM]s. Error bars ± SEM.</p>
Full article ">Figure 5
<p>Effect of OM on the force/[Pi] relations of human ventricular and atrial myofibrils. Maximal Ca<sup>2+</sup>-activated isometric force in the presence of different [Pi] normalised over the force developed in the absence of added Pi under control conditions (black symbols) or in the presence of 0.5 µM [OM] (red symbols). (<b>A</b>): HV myofibrils; (<b>B</b>): HA myofibrils. Hyperbolic fitting of force/[Pi] relations in the absence of OM: Pi<sub>50</sub>, 0.83 ± 0.08 mM for HV and 2.12 ± 0.73 mM for HA myofibrils (asymptotes: 0.29 ± 0.01 and 0.15 ± 0.08, respectively).</p>
Full article ">Figure 6
<p>Effect of OM on the Ca<sup>2+</sup>-independent tension and the Ca<sup>2+</sup>-independent ATPase of skinned human left ventricular strips (25 °C). (<b>A</b>): Representative traces of resting tension in a HV skinned strips in pCa 9.0 or in pCa 9.0 added with 5 µM [OM] that causes Ca<sup>2+</sup>-independent force development. The timing of the solution change is represented by the indexed bar at the top of the tension trace. The arrow marks the time slot when the solution is changed to a pCa 9.0 solution added with both 5 µM [OM] and 5 mM [Pi]. The experiment shows that Pi abolishes Ca<sup>2+</sup>-independent/OM-induced force development. (<b>B</b>): ATPase measurements in pCa 9.0 (black) or in pCa 9.0 added with 5 µM [OM] (red) by enzyme coupled assay. (<b>C</b>): Mean values of Ca<sup>2+</sup>-independent tension (<b>left</b>) and Ca<sup>2+</sup>-independent ATPase (<b>right</b>) in control conditions (black bar) or in the presence of 5 µM [OM] (red bar). Error bars ± SEMs. Initial sarcomere length set to 2.2 µm. * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 5, paired <span class="html-italic">t</span> test.</p>
Full article ">
Back to TopTop