[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (154)

Search Parameters:
Keywords = dimethylsulfoxide

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 8938 KiB  
Article
Designing Microparticles of Luteolin and Naringenin in Different Carriers via Supercritical Antisolvent Process
by Stefania Mottola and Iolanda De Marco
Polymers 2024, 16(24), 3600; https://doi.org/10.3390/polym16243600 - 23 Dec 2024
Abstract
Antioxidants are contained in fruits and vegetables and are commonly obtained through food. However, it is frequently necessary to supplement the diet with substances that are often poorly soluble in water and sensitive to light and oxygen. For this reason, in this work, [...] Read more.
Antioxidants are contained in fruits and vegetables and are commonly obtained through food. However, it is frequently necessary to supplement the diet with substances that are often poorly soluble in water and sensitive to light and oxygen. For this reason, in this work, luteolin (LUT) and naringenin (NAR), two compounds with antioxidant activity and potential health benefits, were precipitated through the supercritical antisolvent technique using polyvinylpyrrolidone and β-cyclodextrin as the carriers. The precipitation occurred from dimethylsulfoxide using supercritical carbon dioxide as the antisolvent. The influence of pressure (9–12 MPa), active substance/carrier concentration in the solution (20–200 mg/mL), and their ratio (1/1 and 1/2 mol/mol) on morphology, particle mean size, and distribution were investigated. Under the optimized operating conditions, spherical microparticles with a mean diameter equal to 2.7 ± 0.9 μm (for LUT) and 5.5 ± 1.9 μm (for NAR) were obtained. The active ingredients were protected from the external environment by the presence of the carrier, and the dissolution rate was notably increased by processing them with β-cyclodextrin. It was sixty times faster and three times faster than that of the antioxidant alone for LUT and NAR, respectively. Full article
(This article belongs to the Section Polymer Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the SAS apparatus. BPV: back-pressure valve; Cr: cryostat; HPP: high-pressure pump; L: liquid; LB: liquid burette; LS: liquid separator; LSP: liquid solution pump; M: manometer; MV: micrometric valve; PV: precipitation vessel.</p>
Full article ">Figure 2
<p>Luteolin/β-CD powder precipitated at 150 mg/mL and 1/2 LUT/β-CD molar ratio. The effect of pressure: (<b>a</b>) FESEM image at 9 MPa; (<b>b</b>) FESEM image at 12 MPa; (<b>c</b>) comparison of PSDs.</p>
Full article ">Figure 3
<p>Effect of luteolin/β-CD molar ratio on the powders precipitated at 150 mg/mL and 12 MPa: (<b>a</b>) FESEM image at 1/2 mol/mol; (<b>b</b>) FESEM image at 1/1 mol/mol; (<b>c</b>) comparison of the PSDs obtained at equimolar ratio (X equal to 0.5) and 1/2 molar ratio (X equal to 0.33).</p>
Full article ">Figure 4
<p>Luteolin/PVP powder precipitated at 20 mg/mL and 12 MPa; (<b>a</b>) FESEM image at 1/2 mol/mol; (<b>b</b>) particle size distribution at 1/2 mol/mol; (<b>c</b>) FESEM image at 1/1 mol/mol.</p>
Full article ">Figure 5
<p>Naringenin/β-CD powder precipitated at 200 mg/mL and 40 °C; (<b>a</b>) FESEM image and PSD at 9 MPa and 1/2 mol/mol; (<b>b</b>) FESEM image and PSD at 12 MPa and 1/2 mol/mol; (<b>c</b>) FESEM image and PSD at 9 MPa and 1/1 mol/mol.</p>
Full article ">Figure 6
<p>FESEM images of naringenin/PVP powders precipitated at 50 mg/mL, 1/2 mol/mol, and 40 °C; (<b>a</b>) microparticles obtained at 9 MPa; (<b>b</b>) crystals obtained at 9 MPa; (<b>c</b>) microparticles obtained at 12 MPa; (<b>d</b>) crystals obtained at 12 MPa.</p>
Full article ">Figure 7
<p>FT-IR spectra of the pure compounds, physical mixtures, and coprecipitated powders: (<b>a</b>) LUT/β-CD; (<b>b</b>) LUT/PVP; (<b>c</b>) NAR/β-CD; (<b>d</b>) NAR/PVP.</p>
Full article ">Figure 8
<p>Comparison of release kinetics: (<b>a</b>) luteolin and (<b>b</b>) naringenin.</p>
Full article ">Figure 9
<p>Job’s plots of the inclusion complexes: (<b>a</b>) luteolin and (<b>b</b>) naringenin.</p>
Full article ">
20 pages, 6567 KiB  
Article
Calixarene-like Lanthanide Single-Ion Magnets Based on NdIII, GdIII, TbIII and DyIII Oxamato Complexes
by Tamyris T. da Cunha, João Honorato de Araujo-Neto, Meiry E. Alvarenga, Felipe Terra Martins, Emerson F. Pedroso, Davor L. Mariano, Wallace C. Nunes, Nicolás Moliner, Francesc Lloret, Miguel Julve and Cynthia L. M. Pereira
Magnetochemistry 2024, 10(12), 103; https://doi.org/10.3390/magnetochemistry10120103 - 12 Dec 2024
Viewed by 519
Abstract
In this work, we describe the synthesis, crystal structures and magnetic properties of four air-stable mononuclear lanthanide(III) complexes with the N-(2,4,6-trimethylphenyl)oxamate (Htmpa) of formula: n-Bu4N[Nd(Htmpa)4(H2O)]·4H2O (1), n-Bu4N[Gd(Htmpa)4 [...] Read more.
In this work, we describe the synthesis, crystal structures and magnetic properties of four air-stable mononuclear lanthanide(III) complexes with the N-(2,4,6-trimethylphenyl)oxamate (Htmpa) of formula: n-Bu4N[Nd(Htmpa)4(H2O)]·4H2O (1), n-Bu4N[Gd(Htmpa)4(H2O)]·3DMSO·2H2O (2), n-Bu4N[Tb(Htmpa)4(H2O)]·3DMSO·1H2O (3) and n-Bu4N[Dy(Htmpa)4(H2O)]·3DMSO·2H2O (4) (n-Bu4N+ = n-tetrabutylammonium; DMSO = dimethylsulfoxide). Their crystal structures reveal the occurrence of calixarene-type monoanionic species containing all-cis-disposed Htmpa ligands and one water molecule coordinated with the respective LnIII ion (Ln = Nd, Gd, Tb and Dy), featuring a nine-coordinated environment with muffin (MFF-9) (1) or spherical-capped square antiprism (CSAPR-9) (24) geometry. The major difference between their crystal structures is related to the nature of crystallization solvent molecules, either water (1) or both DMSO and water (24). The intermolecular hydrogen bonds among the self-complementary Htmpa ligands in all four compounds mediated a 2 D supramolecular network in the solid state. Direct-current (dc) magnetic properties for 14 show typical behavior for the ground state terms of the LnIII ions [4I9/2 (Nd); 8S7/2(Gd), 7F6 (Tb), 6H15/2 (Dy)]. Alternating-current (ac) magnetic measurements reveal the presence of slow magnetic relaxation without the presence of a dc field only for 4. In contrast, field-induced slow magnetic relaxation behavior was found in complexes 1, 2 and 3. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>ORTEP view of the crystal structures of (<b>a</b>) <b>1</b> and (<b>b</b>) <b>3</b>. Crystallization water molecules and hydrogen atoms have been omitted for clarity, and ellipsoids represent 50% probability levels.</p>
Full article ">Figure 2
<p>(<b>a</b>) Top and side views of the coordination polyhedron around the Nd1 metal center of <b>1</b> with the atom numbering scheme for the donor atoms, showing the distorted muffin geometry. The O1D/O3C/O3D atoms compose the triangular base, and the O1A/O1C/O3A/O3B/O1w atoms compose the pentagonal plane with the O1B atom in the capped position. (<b>b</b>) Top and side views of the coordination polyhedron around the Tb1 metal center of <b>3</b> with the atom numbering scheme for the donor atoms, showing the distorted capped square antiprism geometry. The O3A/O3B/O3C/O3D atoms compose the square base of the prism, and the O1A/O1B/O1C/O1D atoms compose the quadratic plane of the capped face with the O1W atom in the capped position.</p>
Full article ">Figure 3
<p>(<b>a</b>) View of the supramolecular layers of hydrogen-bonded mononuclear units of <b>1</b> along the crystallographic <span class="html-italic">ab</span> plane. (<b>b</b>) View of the supramolecular layers of hydrogen-bonded mononuclear units of <b>3</b> along the crystallographic <span class="html-italic">ac</span> plane, showing the presence of additional hydrogen-bonded DMSO molecules. Cyan dashed lines represent hydrogen bonds. CH hydrogen atoms, remaining solvent molecules and tetrabutylammonium cations were omitted for the sake of clarity.</p>
Full article ">Figure 4
<p>Views perpendicular to the layers growing onto the (<b>a</b>) <span class="html-italic">ab</span> plane in <b>1</b> and (<b>b</b>) <span class="html-italic">ac</span> plane in <b>3</b>. Three layers are depicted in each panel, while hydrogen atoms and solvent molecules were omitted for the sake of clarity. Color codes: carbon, gray; nitrogen, light blue; oxygen, red; Nd, light green; and terbium, green.</p>
Full article ">Figure 5
<p><span class="html-italic">χ</span><sub>M</sub><span class="html-italic">T</span> vs. <span class="html-italic">T</span> curves for <b>1</b>–<b>4</b> (<b>a</b>–<b>d</b>). Inset: <span class="html-italic">M</span> vs. <span class="html-italic">H</span> curves. Solid lines represent the best-fit curves according to the text.</p>
Full article ">Figure 6
<p>(<b>a</b>) Representation of the easy magnetization axis of dysprosium(III) complex <b>4</b> as a red arrow in each single-ion magnet (color code: Dy in violet, N in blue, oxygen in red, carbon in grey). (<b>b</b>) The easy magnetization axis is represented by red arrows in the crystal packing of <b>4</b> along the crystallographic <span class="html-italic">ac</span> direction. For clarity, the hydrogen atoms, solvent molecules and counterions were omitted.</p>
Full article ">Figure 7
<p>(<b>a</b>) <span class="html-italic">χ</span><sub>M</sub>′ and <span class="html-italic">χ</span><sub>M</sub>″ vs. <span class="html-italic">ν</span> curves of <b>1</b> under 1.0 kOe dc magnetic field. (<b>b</b>) Arrhenius plot of <b>1</b> under 1.0 kOe dc magnetic field. Solid lines represent fits to the data (see text).</p>
Full article ">Figure 8
<p>(<b>a</b>) <span class="html-italic">χ</span><sub>M</sub>′ and <span class="html-italic">χ</span><sub>M</sub>″ vs<span class="html-italic">. ν</span> curves of <b>2</b> under 1.0 kOe dc magnetic field. (<b>b</b>) Arrhenius plot of <b>2</b> under a 1.0 kOe dc magnetic field. Solid lines represent fits to the data (see text).</p>
Full article ">Figure 9
<p>(<b>a</b>) <span class="html-italic">χ</span><sub>M</sub>′ and <span class="html-italic">χ</span><sub>M</sub>″ vs. <span class="html-italic">ν</span> curves of <b>3</b> under 5.0 kOe dc magnetic field. (<b>b</b>) Arrhenius plot of <b>3</b> under 5.0 kOe dc magnetic field. Solid lines represent fits to the data (see text).</p>
Full article ">Figure 10
<p><span class="html-italic">χ</span><sub>M</sub>′ and <span class="html-italic">χ</span><sub>M</sub>″ vs. <span class="html-italic">ν</span> curves of compound <b>4</b> under (<b>a</b>) zero dc magnetic field and (<b>b</b>) 1.0 kOe dc magnetic field. Arrhenius plots of <b>4</b> (<b>c</b>) under zero dc field and (<b>d</b>) 1.0 kOe dc field. Solid lines represent fits to the data (see text).</p>
Full article ">Scheme 1
<p>(<b>a</b>) Chemical structure of the proligand <span class="html-italic">N</span>-(2,4,6-trimethylphenyl)oxamic acid ethyl ester (EtHtmpa) and (<b>b</b>) their mononuclear lanthanide(III) oxamate complexes.</p>
Full article ">Scheme 2
<p>Synthetic procedure for complexes <b>1</b>–<b>4</b>.</p>
Full article ">
14 pages, 5228 KiB  
Article
Effect of Selected Organic Solvents on Hydroxyl Radical-Dependent Light Emission in the Fe2+-EGTA-H2O2 System
by Krzysztof Sasak, Michał Nowak, Anna Wlodarczyk, Agata Sarniak and Dariusz Nowak
Molecules 2024, 29(23), 5635; https://doi.org/10.3390/molecules29235635 - 28 Nov 2024
Viewed by 527
Abstract
Numerous compounds that are scavengers of hydroxyl radicals (•OH) in Fenton systems have low solubility in water. Therefore, they are dissolved in organic solvents to reach suitable concentrations in the reaction milieu of the Fenton system. However, these solvents may react with •OH [...] Read more.
Numerous compounds that are scavengers of hydroxyl radicals (•OH) in Fenton systems have low solubility in water. Therefore, they are dissolved in organic solvents to reach suitable concentrations in the reaction milieu of the Fenton system. However, these solvents may react with •OH and iron, leading to significant errors in the results. We evaluated 11 solvents (4 alcohols, acetone, 4 esters, dimethyl-sulfoxide, and acetonitrile) at concentrations ranging from 0.105 µmol/L to 0.42 µmol/L to assess their effects on light emission, a recognized measure of •OH radical activity, in the Fe2+-EGTA-H2O2 system. Six solvents inhibited and four solvents enhanced light emission at all tested concentrations. Acetonitrile, which initially suppressed light emission, lost this effect at a concentration of 0.105 µmol/L, (−1 ± 13 (2; 0) %, p > 0.05). Methanol, at the lowest tested concentration, inhibited light emission by 62 ± 4% (p < 0.05), while butyl butyrate enhanced it by 93 ± 16% (p < 0.05). These effects may be explained by solvent-driven •OH-scavenging, inhibition or acceleration of Fe2+ regeneration, or photon emission from excited solvent molecules. Our findings suggest that acetonitrile seems suitable for preparing stock solutions to evaluate antioxidant activity in the Fe2+-EGTA-H2O2 system, provided that the final concentration of this solvent in the reaction milieu is kept below 0.105 µmol/L. Full article
Show Figures

Figure 1

Figure 1
<p>Plausible mechanisms through which organic solvents affect UPE in the Fe<sup>2</sup>⁺-EGTA-H<sub>2</sub>O<sub>2</sub> System. (<b>A</b>) Without the addition of solvent. In this case, the •OH radicals generated by the Fe<sup>2</sup>⁺-EGTA-H<sub>2</sub>O<sub>2</sub> system attack ether bonds in the EGTA structure leading to the formation of products containing triplet-excited carbonyl groups (R-CH-O*), which results in light emission (UPE—ultra-weak photon emission). Simultaneously, Fe<sup>2</sup>⁺ is regenerated according to the following chemical equations: Fe<sup>3+</sup>-EGTA + O<sub>2</sub>•- → Fe<sup>2+</sup>-EGTA + O<sub>2</sub> (1); H<sub>2</sub>O<sub>2</sub> + Fe<sup>3+</sup>-EGTA → HO<sub>2</sub>• + H<sup>+</sup> + Fe<sup>2+</sup>- EGTA (2); and HO<sub>2</sub>• + Fe<sup>3+</sup>-EGTA → O<sub>2</sub> + H<sup>+</sup> + Fe<sup>2+</sup>-EGTA (3). This regenerated Fe<sup>2</sup>⁺ contributes to the further production of •OH radicals and UPE. (<b>B</b>) With the addition of an organic solvent that inhibits UPE in the Fe<sup>2</sup>⁺-EGTA-H<sub>2</sub>O<sub>2</sub> system. The solvent directly scavenges •OH radicals, thereby protecting the ether bonds in EGTA and inhibiting the formation of triplet-excited carbonyl-containing products, which reduces light emission. Some solvents (e.g., ethanol) may also inhibit Fe<sup>2</sup>⁺ regeneration (Equation (1)), further suppressing UPE. (<b>C</b>) With the addition of an organic solvent that enhances UPE in the Fe<sup>2</sup>⁺-EGTA-H<sub>2</sub>O<sub>2</sub> system. The solvent accelerates Fe<sup>2</sup>⁺ regeneration. In some cases, such as with acetone, the solvent itself may emit light due to excitation to a triplet state. However, these solvents can simultaneously scavenge •OH radicals; therefore, the enhancing effect is observed when the rate of •OH radical generation, driven by accelerated Fe<sup>2</sup>⁺ regeneration, exceeds the loss of radicals due to the solvent’s scavenging activity.</p>
Full article ">
9 pages, 2551 KiB  
Article
Effect of the Protic vs. Non-Protic Molecular Environment on the cis to trans Conformation Change of Phototrexate Drug
by Flórián Bencze, László Kiss, Heng Li, Hui Yan, László Kollár and Sándor Kunsági-Máté
Int. J. Mol. Sci. 2024, 25(23), 12703; https://doi.org/10.3390/ijms252312703 - 26 Nov 2024
Viewed by 444
Abstract
The therapeutical applicability of the anticancer drug phototrexate, a photoswitchable derivative of the antimetabolite dihydrofolate reductase inhibitor methotrexate, highly depends on the stability of its bioactive isomer. Considering that only the cis configuration of phototrexate is bioactive, in this work, the effect of [...] Read more.
The therapeutical applicability of the anticancer drug phototrexate, a photoswitchable derivative of the antimetabolite dihydrofolate reductase inhibitor methotrexate, highly depends on the stability of its bioactive isomer. Considering that only the cis configuration of phototrexate is bioactive, in this work, the effect of the molecular environment on the stability of the cis isomer of this drug has been investigated. UV-vis absorption and fluorescence-based solvent relaxation methods have been used. Protic methanol and non-protic dimethylsulfoxide were used as medium-ranged permittivity solvents. The results showed a decreased rate of cis → trans conversion and enhanced stabilities of the cis isomer in methanol. Temperature-dependent measurements of the isomerization rate reflect the increased activation energy in methanol. Full article
(This article belongs to the Section Molecular Biophysics)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of <span class="html-italic">trans</span>-Phototrexate (<span class="html-italic">trans</span>-PHX) and <span class="html-italic">cis</span>-Phototrexate (<span class="html-italic">cis</span>-PHX) and the reversible isomerization of PHX. Bottom: 3D structure of the two isomers of the PHX: <span class="html-italic">trans</span>-PHX (left) and <span class="html-italic">cis</span>-PHX (right).</p>
Full article ">Figure 2
<p>Changes of absorption spectra of PHX with 50 μM concentrations in differently composed methanol–DMSO mixtures. (<b>Left</b>) DMSO concentration varies within the 10 vol % and 50 vol % range. The spectra remain unchanged above 50 vol % DMSO. (<b>Right</b>) DMSO concentration varies within the range of 10 vol % and 28 vol %.</p>
Full article ">Figure 3
<p>Correlation of the solvent relaxation times and the composition of the solvation shell of phototrexate molecules estimated from the absorption spectra. DMSO concentration varies within the range of 6 vol % and 28 vol %.</p>
Full article ">Figure 4
<p>Shows the time-dependent representative absorption spectra of PHX in two cases: (<b>a</b>) the solvent composed of pure DMSO and (<b>b</b>) the solvent composed of 90 vol % methanol and 10 vol % DMSO.</p>
Full article ">Figure 5
<p>(<b>Left</b>) Representative <span class="html-italic">cis</span> → <span class="html-italic">trans</span> conversion curves of PHX in two cases: the solvent is pure DMSO or consists of 90 vol % methanol and 10 vol % DMSO. (<b>Right</b>): k<sub>2</sub> plotted against vol % DMSO.</p>
Full article ">Figure 6
<p>Arrhenian plots of the rates of the second conversion step associated with the <span class="html-italic">cis</span> → <span class="html-italic">trans</span> conversion of PHX in two cases: (<b>a</b>) the solvent composed of pure DMSO and (<b>b</b>) the solvent composed of 90 vol % methanol and 10 vol % DMSO. The slopes determine 86 kJ/mol and 234 kJ/mol activation energy for the (<b>a</b>,<b>b</b>) processes, respectively.</p>
Full article ">Figure 7
<p>Optimized structure of the coordination of two methanol molecules to the <span class="html-italic">cis</span>-PHX. (HF/6-31G* level of calculation) The methanol molecules are stabilized preferably by the hydrogen bonds.</p>
Full article ">Figure 8
<p>Schematic scheme of the synthesis of PHX.</p>
Full article ">
15 pages, 2650 KiB  
Article
Evaluation of Cryopreservation of Bovine Ovarian Tissue by Analysis of Reactive Species of Oxygen, Toxicity, Morphometry, and Morphology
by Camila Bizarro-Silva, Larissa Zamparone Bergamo, Camila Bortoliero Costa, Suellen Miguez González, Deborah Nakayama Yokomizo, Ana Carolina Rossaneis, Waldiceu Aparecido Verri Junior, Mateus José Sudano, Evelyn Rabelo Andrade, Amauri Alcindo Alfieri and Marcelo Marcondes Seneda
Vet. Sci. 2024, 11(11), 579; https://doi.org/10.3390/vetsci11110579 - 19 Nov 2024
Viewed by 645
Abstract
Ovarian tissue cryopreservation has been widely investigated for preserving female fertility. In the present study, we aimed to compare the effects of three concentrations (1, 1.5, and 3 M) of dimethylsulfoxide (DMSO) on the vitrification of ovarian tissue. The ovarian cortex was divided [...] Read more.
Ovarian tissue cryopreservation has been widely investigated for preserving female fertility. In the present study, we aimed to compare the effects of three concentrations (1, 1.5, and 3 M) of dimethylsulfoxide (DMSO) on the vitrification of ovarian tissue. The ovarian cortex was divided into control and vitrified groups: (i) 1 M-DMSO, (ii) 1.5 M-DMSO, and (iii) 3 M-DMSO. Follicles from all fragments were analyzed for DMSO-induced deleterious effects, morphological and morphometric aspects, and concentration of reactive oxygen species. Additionally, the fragments were cultured to assess the integrity and return of follicular development post-vitrification. All DMSO concentrations resulted in a higher percentage of degenerated preantral follicles than before the cryopreservation process. After vitrification, the cryopreserved ovarian fragments showed similar percentages of intact follicles; however, the 3 M DMSO concentration differed from the control. Analyzing free radical production, we found that the 3 M DMSO concentration had higher levels of oxidative stress than the lower DMSO. After in vitro cultivation of the vitrified/warmed fragments, the 1 M DMSO concentration exhibited higher percentages of morphologically intact follicles than the other concentrations. Therefore, we suggest that bovine preantral follicles can be cryopreserved in situ with greater efficiency in 1 M DMSO. Full article
(This article belongs to the Special Issue Assessment of Oxidant and Antioxidant Status in Livestock)
Show Figures

Figure 1

Figure 1
<p>Experimental design of the protocol used for cryopreservation of bovine ovarian tissue.</p>
Full article ">Figure 2
<p>Degenerated (black bars) and intact (white bars) preantral follicles (%) in non-vitrified ovarian fragments (control) and fragments exposed to different DMSO concentrations (1, 1.5, and 3 M) to analyze the deleterious effect of DMSO histologically. Uppercase letters (A, B) indicate significant differences between treatments within the degenerated preantral follicles (<span class="html-italic">p</span> &lt; 0.05). Lowercase letters (a, b) indicate significant differences between treatments within intact preantral follicles (<span class="html-italic">p</span> &lt; 0.05). (*) indicates significant differences between degenerated and intact follicles within the same treatment group (<span class="html-italic">p</span> &lt; 0.05). DMSO, dimethylsulfoxide.</p>
Full article ">Figure 3
<p>Percentage of grade 1, 2, and 3 degenerated bovine preantral follicles upon exposure to DMSO concentrations of 1 M (white bars), 1.5 M (gray bars), and 3 M (black bars) to analyze the deleterious effect of DMSO. Values followed by lowercase letters (α, β, and γ) indicate significant differences between DMSO concentrations within the same degree of degeneration (<span class="html-italic">p</span> &lt; 0.05). Values followed by lowercase letters (a, b) indicate significant differences between the degrees of degeneration (<span class="html-italic">p</span> &lt; 0.05). DMSO, dimethylsulfoxide.</p>
Full article ">Figure 4
<p>Representative histological images of the morphology of vitrified preantral follicles. (<b>A</b>): intact primordial follicle; (<b>B</b>): intact primary follicle; (<b>C</b>): intact secondary follicle; (<b>D</b>): degenerated primordial follicles (cytoplasmic vacuoles); (<b>E</b>): degenerated primordial follicles (granulosa cell disorganization); (<b>F</b>): degenerated primordial follicle (nuclear retraction). Periodic Acid–Schiff (PAS) and hematoxylin, 100×.</p>
Full article ">Figure 5
<p>Percentage of primordial and developing follicles in fresh ovarian tissue (control) and vitrification/warming with different DMSO concentrations (1, 1.5, and 3 M). DMSO, dimethylsulfoxide.</p>
Full article ">Figure 6
<p>Percentage of intact (white bars) and degenerated (black bars) follicles in fresh ovarian tissue (control, day 0) and after vitrification with different concentrations of dimethyl sulfoxide (DMSO). Values followed by lowercase letters (a, b) indicate significant differences between DMSO concentrations in intact preantral follicles (<span class="html-italic">p</span> &lt; 0.05). Values followed by lowercase letters (α, β) indicate significant differences between DMSO concentrations in degenerated preantral follicles (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Average follicular (gray section) and oocyte (white section) diameter (μm) of bovine preantral follicles in tissue samples vitrified with different concentrations of dimethyl sulfoxide (DMSO; 1, 1.5, or 3 M). Values followed by different letters (α, β) indicate significant differences in oocyte diameter between DMSO concentrations (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Percentage of preantral follicles in follicular development classified as primordial (white bars) and developing (primary + secondary; black bars) in un-vitrified ovarian tissue (control) and after in vitro cultured ovarian fragments vitrified with different concentrations of DMSO (1, 1.5, and 3 M) for 10 days. Uppercase letters (A, B) indicate significant differences between treatments within primordial follicles (<span class="html-italic">p</span> &lt; 0.05). Lowercase letters (a, b) indicate significant differences between treatments in the development (primary and secondary) of follicles (<span class="html-italic">p</span> &lt; 0.05). (*) indicates significant differences between primordial and developing follicles within the same treatment group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Effects of DMSO concentrations (1 M, 1.5, and 3 M) on oxidative stress (OS) on vitrified ovarian tissue. (<b>A</b>) The ability to scavenge free radicals (ABTS assay) and (<b>B</b>) the ferric-reducing antioxidant power (FRAP assay) were compared with a Trolox curve. The production of superoxide anion (<b>C</b>) and lipid peroxidation (<b>D</b>) were determined by the NBT and TBARS tests, respectively. Results are means ± SEM (<span class="html-italic">n</span> = 10). One-way ANOVA followed by Tukey’s test. Values followed by lowercase letters (a, b) differ significantly between DMSO concentrations (<span class="html-italic">p</span> &lt; 0.05). DMSO, dimethylsulfoxide; NBT, nitroblue tetrazolium; TBARS, thiobarbituric acid reactive substance.</p>
Full article ">
17 pages, 11316 KiB  
Article
Effects of Moisture Content and Heat Treatment on the Viscoelasticity and Gelation of Polyacrylonitrile/Dimethylsulfoxide Solutions
by Jae-Yeon Yang, Yun-Su Kuk, Byoung-Suhk Kim and Min-Kang Seo
Gels 2024, 10(11), 728; https://doi.org/10.3390/gels10110728 - 10 Nov 2024
Viewed by 725
Abstract
Polyacrylonitrile (PAN) gels create significant obstacles in industrial fiber spinning by forming insoluble networks that compromise solution stability and uniformity. This study investigates the rheological properties of PAN/dimethyl sulfoxide (DMSO) solutions, examining how aging time, moisture content, and polymer concentration affect gelation behavior. [...] Read more.
Polyacrylonitrile (PAN) gels create significant obstacles in industrial fiber spinning by forming insoluble networks that compromise solution stability and uniformity. This study investigates the rheological properties of PAN/dimethyl sulfoxide (DMSO) solutions, examining how aging time, moisture content, and polymer concentration affect gelation behavior. Dynamic rheological analysis revealed that both physical and chemical crosslinks play crucial roles in gel formation, with gelation accelerating markedly when moisture content exceeds 3% and aging progresses. Under heat treatment at 80 °C, samples with increased moisture content demonstrated rapid transitions to solid-like states, indicating a critical moisture threshold for enhanced gelation kinetics. Additionally, reductions in polymer concentration disrupted physical crosslink density, thereby mitigating gel formation. These results underscore the importance of precisely controlling moisture and concentration parameters in PAN solutions to stabilize solution properties and minimize gel formation, thus enhancing process efficiency and quality in PAN-based carbon fiber production. Full article
(This article belongs to the Special Issue Gel-Based Materials: Preparations and Characterization (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Complex viscosity (<b>a</b>), storage and loss modulus (<b>b</b>), and loss factors (<b>c</b>) of samples with different moisture contents with respect to time.</p>
Full article ">Figure 2
<p>Complex viscosity (<b>a</b>), storage modulus (<b>b</b>), loss modulus (<b>c</b>), and loss factors (<b>d</b>) of the heat-treated samples with different moisture contents with respect to the time. A schematic of the gelation behavior (<b>e</b>) of PAN/DMSO solutions.</p>
Full article ">Figure 3
<p>Complex viscosity (<b>a</b>), storage and loss modulus (<b>b</b>), and loss factors (<b>c</b>) of the P-WT3-H80-24 at different frequencies with respect to the time.</p>
Full article ">Figure 4
<p>(<b>a</b>) Behavior of ions and dipoles in polymers and (<b>b</b>) ion and dipole conductions under an external electric field.</p>
Full article ">Figure 5
<p>Ion viscosity (<b>a</b>) and complex viscosity (<b>b</b>) of the PAN/DMSO solutions with different moisture contents.</p>
Full article ">Figure 6
<p>UV absorbance spectra of the heat-treated PAN/DMSO solutions with different moisture contents.</p>
Full article ">Figure 7
<p>TGA thermograms of the PAN/DMSO solutions with different moisture contents and heat treatments.</p>
Full article ">Figure 8
<p>Particle size of PAN solutions with respect to the moisture content and heat treatment.</p>
Full article ">Figure 9
<p>Surface roughness analysis of PAN films: (<b>a</b>) P-WT0-H80-24, (<b>b</b>) P-WT3-H80-24, and (<b>c</b>) P-WT5-H80-24.</p>
Full article ">Figure 10
<p>Photographs of the reactor used for the suspension polymerization of PAN polymers (<b>left</b>) and the obtained product powders (<b>right</b>).</p>
Full article ">Figure 11
<p>Nomenclature order of PAN solutions.</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 - 16 Oct 2024
Viewed by 1086
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 ">
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 1118
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 608
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 ">
11 pages, 4752 KiB  
Article
Investigation of the Application of Reduced Graphene Oxide–SPION Quantum Dots for Magnetic Hyperthermia
by Haneen Omar, Yara Ahmed Alkurdi, Arshia Fathima and Edreese H. Alsharaeh
Nanomaterials 2024, 14(19), 1547; https://doi.org/10.3390/nano14191547 - 25 Sep 2024
Viewed by 880
Abstract
Integrating hyperthermia with conventional cancer therapies shows promise in improving treatment efficacy while mitigating their side effects. Nanotechnology-based hyperthermia, particularly using superparamagnetic iron oxide nanoparticles (SPIONs), offers a simplified solution for cancer treatment. In this study, we developed composites of SPION quantum dots [...] Read more.
Integrating hyperthermia with conventional cancer therapies shows promise in improving treatment efficacy while mitigating their side effects. Nanotechnology-based hyperthermia, particularly using superparamagnetic iron oxide nanoparticles (SPIONs), offers a simplified solution for cancer treatment. In this study, we developed composites of SPION quantum dots (Fe3O4) with reduced graphene oxide (Fe3O4/RGO) using the coprecipitation method and investigated their potential application in magnetic hyperthermia. The size of Fe3O4 nanoparticles was controlled within the quantum dot range (≤10 nm) by varying the synthesis parameters, including reaction time as well as the concentration of ammonia and graphene oxide, where their biocompatibility was further improved with the inclusion of polyethylene glycol (PEG). These nanocomposites exhibited low cytotoxic effects on healthy cells (CHO-K1) over an incubation period of 24 h, though the inclusion of PEG enhanced their biocompatibility for longer incubation periods over 48 h. The Fe3O4/RGO composites dispersed in acidic pH buffer (pH 4.66) exhibited considerable heating effects, with the solution temperature increasing by ~10 °C within 5 min of exposure to pulsed magnetic fields, as compared to their dispersions in phosphate buffer and aqueous dimethylsulfoxide solutions. These results demonstrated the feasibility of using quantum dot Fe3O4/RGO composites for magnetic hyperthermia-based therapy to treat cancer, with further studies required to systematically optimize their magnetic properties and evaluate their efficacy for in vitro and in vivo applications. Full article
Show Figures

Figure 1

Figure 1
<p>XRD for Fe<sub>3</sub>O<sub>4</sub> quantum dots synthesized using (<b>a</b>) different amounts of ammonia and (<b>b</b>) different reaction times, and its composites with RGO obtained by (<b>c</b>) varying the amounts of GO solution used during their synthesis.</p>
Full article ">Figure 2
<p>XRD of Fe<sub>3</sub>O<sub>4</sub>/RGO nanocomposites with the inclusion of PEG synthesized at optimal conditions (6 mL NH<sub>3</sub>, 25 µL GO solution and 75 min reaction time).</p>
Full article ">Figure 3
<p>UV-Vis spectra of optimized Fe<sub>3</sub>O<sub>4</sub>, Fe<sub>3</sub>O<sub>4</sub>/RGO and Fe<sub>3</sub>O<sub>4</sub>/RGO/PEG.</p>
Full article ">Figure 4
<p>TEM of (<b>a</b>) Fe<sub>3</sub>O<sub>4</sub> and (<b>b</b>) optimized Fe<sub>3</sub>O<sub>4</sub>/RGO/PEG nanocomposites.</p>
Full article ">Figure 5
<p>Heating profiles for solutions of Fe<sub>3</sub>O<sub>4</sub>/RGO (black) and Fe<sub>3</sub>O<sub>4</sub>/RGO/PEG (red) in (<b>a</b>,<b>d</b>) PBS, (<b>b</b>,<b>e</b>) aqueous DMSO and (<b>c</b>,<b>f</b>) acidic pH buffer upon exposure to magnetic field at 150 A.</p>
Full article ">Figure 6
<p>Cell viability of (<b>a</b>) CHO-K1 cells incubated with Fe<sub>3</sub>O<sub>4</sub>, Fe<sub>3</sub>O<sub>4</sub>/RGO and Fe<sub>3</sub>O<sub>4</sub>/RGO/PEG (200 and 320 µg/mL) over a period of 24 h and 48 h, and (<b>b</b>) HEK-293 cancer cells incubated with Fe<sub>3</sub>O<sub>4</sub>/RGO/PEG (200 µg/mL) at 24 and 48 h incubation time.</p>
Full article ">
21 pages, 7976 KiB  
Article
A3B Zn(II)-Porphyrin-Coated Carbon Electrodes Obtained Using Different Procedures and Tested for Water Electrolysis
by Bogdan-Ovidiu Taranu, Florina Stefania Rus and Eugenia Fagadar-Cosma
Coatings 2024, 14(8), 1048; https://doi.org/10.3390/coatings14081048 - 16 Aug 2024
Viewed by 702
Abstract
In the context of water electrolysis being highlighted as a promising technology for the large-scale sustainable production of hydrogen, the water-splitting electrocatalytic properties of an asymmetrically functionalized A3B zinc metalated porphyrin, namely, Zn(II) 5-(4-pyridyl)-10,15,20-tris(4-phenoxyphenyl)-porphyrin, were evaluated in a wide pH range. [...] Read more.
In the context of water electrolysis being highlighted as a promising technology for the large-scale sustainable production of hydrogen, the water-splitting electrocatalytic properties of an asymmetrically functionalized A3B zinc metalated porphyrin, namely, Zn(II) 5-(4-pyridyl)-10,15,20-tris(4-phenoxyphenyl)-porphyrin, were evaluated in a wide pH range. Two different electrode manufacturing procedures were employed to outline the porphyrin’s applicative potential for the O2 and H2 evolution reactions (OER and HER). The electrode, manufactured by coating the catalyst on a graphite support from a dimethylsulfoxide solution, displayed electrocatalytic activity for the OER in an acidic electrolyte. An overpotential value of 0.44 V (at i = 10 mA/cm2) and a Tafel slope of 0.135 V/dec were obtained. The modified electrode that resulted from applying a Zn(II)-porphyrin-containing catalyst ink onto the same substrate type was identified as a bifunctional water-splitting catalyst in a neutral medium. OER and HER overpotentials of 0.78 and 1.02 V and Tafel slopes of 0.39 and 0.249 V/dec were determined. This is the first Zn(II)-porphyrin to be reported as a heterogenous bifunctional water-splitting electrocatalyst in neutral aqueous electrolyte solution and is one of very few porphyrins behaving as such. The TEM analysis of the porphyrin’s self-assembly behavior revealed a wide variety of architectures. Full article
(This article belongs to the Special Issue Environmentally Friendly Energy Conversion Materials and Thin Films)
Show Figures

Figure 1

Figure 1
<p>STEM images recorded on specimens prepared by applying ZnPyTPPP from DCM (<b>a</b>), THF (<b>b</b>), PhCN (<b>c</b>), DMF (<b>d</b>), and DMSO (<b>e</b>). The insets in (<b>a</b>–<b>c</b>,<b>e</b>) present STEM and TEM images collected at higher magnification.</p>
Full article ">Figure 2
<p>LSVs obtained during OER experiments on the G<sub>0</sub>, G<sub>PZn-DCM</sub>, G<sub>PZn-THF</sub>, G<sub>PZn-PhCN</sub>, G<sub>PZn-DMF</sub>, and G<sub>PZn-DMSO</sub> electrodes immersed in 1 M KOH (<b>a</b>), 0.1 M KCl (<b>b</b>), and 0.5 M H<sub>2</sub>SO<sub>4</sub> (<b>c</b>) electrolyte solutions at <span class="html-italic">v</span> = 5 mV/s and η<sub>OER</sub> bar column graphs measured at i = 10 mA/cm<sup>2</sup> for the alkaline (<b>d</b>), neutral (<b>e</b>), and acidic (<b>f</b>) media.</p>
Full article ">Figure 3
<p>Graphical representations of i<sub>a</sub> and i<sub>c</sub> vs. <span class="html-italic">v</span><sup>1/2</sup> for G<sub>PZn-DMF</sub> (<b>a</b>) and G<sub>PZn-DMSO</sub> (<b>b</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>) The Tafel plot for the G<sub>PZn-DMSO</sub> electrode in 0.5 M H<sub>2</sub>SO<sub>4</sub> solution. (<b>b</b>) Chronoamperometric curves recorded on G<sub>PZn-DMSO</sub> electrodes, in 0.5 M H<sub>2</sub>SO<sub>4</sub> solution, at constant E values corresponding to i = 10 mA/cm<sup>2</sup> and i = 5 mA/cm<sup>2</sup>, respectively. (<b>c</b>) The LSVs obtained on the G<sub>PZn-DMSO</sub> electrode before and after the electrochemical stability test performed at the constant potential corresponding to i = 5 mA/cm<sup>2</sup> (0.5 M H<sub>2</sub>SO<sub>4</sub> solution and <span class="html-italic">v</span> = 5 mV/s).</p>
Full article ">Figure 5
<p>SEM micrographs obtained on the G<sub>PZn-DMSO</sub> electrode: (<b>a</b>,<b>b</b>) before the chronoamperometric stability experiment and (<b>c</b>,<b>d</b>) after the test.</p>
Full article ">Figure 6
<p>LSVs obtained during OER experiments on the G<sub>CB</sub> and G<sub>CB-PZn</sub> electrodes immersed in 0.5 M H<sub>2</sub>SO<sub>4</sub> (<b>a</b>) and in 0.1 M KCl (<b>b</b>) electrolyte solutions at <span class="html-italic">v</span> = 5 mV/s.</p>
Full article ">Figure 7
<p>Graphical representations of i<sub>a</sub> and i<sub>c</sub> vs. <span class="html-italic">v</span><sup>1/2</sup> for G<sub>CB</sub> (<b>a</b>) and G<sub>CB-PZn</sub> (<b>b</b>).</p>
Full article ">Figure 8
<p>(<b>a</b>) The Tafel plot for the G<sub>CB-PZn</sub> electrode in 0.1 M KCl solution. (<b>b</b>) Chronoamperometric curve recorded on G<sub>CB-PZn</sub> in 0.1 M KCl solution and inset with the LSVs obtained on the same electrode before and after the electrochemical stability test (0.1 M KCl solution and <span class="html-italic">v</span> = 5 mV/s).</p>
Full article ">Figure 9
<p>(<b>a</b>) LSVs obtained during HER experiments on the G<sub>CB</sub> and G<sub>CB-PZn</sub> electrodes immersed in 0.1 M KCl electrolyte solution at <span class="html-italic">v</span> = 5 mV/s. (<b>b</b>) The Tafel plot for G<sub>CB-PZn</sub> in 0.1 M KCl solution. (<b>c</b>) Chronoamperometric curve recorded on G<sub>CB-PZn</sub> in 0.1 M KCl solution and inset with the LSVs obtained on the same electrode before and after the electrochemical stability test.</p>
Full article ">Figure 10
<p>SEM images recorded on G<sub>CB-PZn</sub> electrodes before any electrochemical testing (<b>a</b>), after testing at constant anodic potential (<b>b</b>), and after testing at constant cathodic potential (<b>c</b>).</p>
Full article ">Scheme 1
<p>Chemical structure of Zn(II) 5-(4-pyridyl)-10,15,20-tris(4-phenoxyphenyl)-porphyrin.</p>
Full article ">
12 pages, 12253 KiB  
Article
Photocatalytic N-Formylation of CO2 with Amines Catalyzed by Diethyltriamine Pentaacetic Acid
by Xuexin Yuan, Qiqi Zhou, Yu Chen, Hai-Jian Yang, Qingqing Jiang, Juncheng Hu and Cun-Yue Guo
C 2024, 10(3), 62; https://doi.org/10.3390/c10030062 - 11 Jul 2024
Viewed by 1275
Abstract
In the present work, inexpensive and commercially available diethyltriamine pentaacetic acid (DTPA) was used as an initiator to catalyze the N-formylation reaction of CO2 with amines via the construction of C-N bonds in the presence of xanthone as the photosensitizer and PhSiH [...] Read more.
In the present work, inexpensive and commercially available diethyltriamine pentaacetic acid (DTPA) was used as an initiator to catalyze the N-formylation reaction of CO2 with amines via the construction of C-N bonds in the presence of xanthone as the photosensitizer and PhSiH3 as the reducing agent. After a systematic study of various factors, the optimal conditions for the photocatalytic reaction were obtained: 2.5 mmol of amine, 2.5 mmol of PhSiH3, 10 mol% of DTPA, 20 mol% of xanthone, 1 mL of dimethylsulfoxide (DMSO), atmospheric pressure, and 35 W UV lamp irradiation for 48 h. Under the optimal conditions, the catalyst system afforded high performance for the N-formylation of amines (primary and secondary amines) and CO2, and the yields of the N-formylated products of dialkylamines were above 70%. Further studies exhibit that the catalytic system has a wide scope of substrate applications. For various alicyclic secondary amines, heterocyclic secondary amines, aliphatic primary amines, and aromatic primary amines, the corresponding N-formylation products can be obtained efficiently. In addition, the catalyst can be recycled by simple precipitation and filtration. After five cycles of recycling, there was no significant change in the catalytic and structural properties of DTPA. Finally, a possible reaction mechanism is proposed. Full article
(This article belongs to the Section CO2 Utilization and Conversion)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photocatalyzed CO<sub>2</sub> N-formylation reaction.</p>
Full article ">Figure 2
<p>Effect of solvent on the N-formylation of N-methylaniline with CO<sub>2</sub>. Reaction conditions: N-methylaniline (2.5 mmol), xanthone (20 mol%), DTPA (10 mol%), solvent (1 mL), CO<sub>2</sub> (0.1 MPa), PhSiH<sub>3</sub> (2.5 mmol), rt, and 48 h. The yield was determined by <sup>1</sup>H NMR spectra analysis using 1,3,5-trimethoxybenzene as an internal standard.</p>
Full article ">Figure 3
<p>Recycling experiment of DTPA-catalyzed N-formylation reaction of CO<sub>2</sub> with aniline.</p>
Full article ">Figure 4
<p>Infrared comparison spectra of DTPA before and after 5 times of reuse.</p>
Full article ">Figure 5
<p>Reaction mechanism of CO<sub>2</sub> with aniline N-formylation.</p>
Full article ">
13 pages, 1335 KiB  
Article
New Methodology for Modifying Sodium Montmorillonite Using DMSO and Ethyl Alcohol
by Adriana Stoski, Bruno Rafael Machado, Bruno Henrique Vilsinski, Lee Marx Gomes de Carvalho, Edvani Curti Muniz and Carlos Alberto Policiano Almeida
Materials 2024, 17(12), 3029; https://doi.org/10.3390/ma17123029 - 20 Jun 2024
Viewed by 823
Abstract
Modified clays with organic molecules have many applications, such as the adsorption of pollutants, catalysts, and drug delivery systems. Different methodologies for intercalating these structures with organic moieties can be found in the literature with many purposes. In this paper, a new methodology [...] Read more.
Modified clays with organic molecules have many applications, such as the adsorption of pollutants, catalysts, and drug delivery systems. Different methodologies for intercalating these structures with organic moieties can be found in the literature with many purposes. In this paper, a new methodology of modifying Sodium Montmorillonite clays (Na-Mt) with a faster drying time was investigated by X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), BET, and thermogravimetric analysis (TG and DTG). In the modification process, a mixture of ethyl alcohol, DMSO, and Na-Mt were kept under magnetic stirring for one hour. Statistical analysis was applied to evaluate the effects of the amount of DMSO, temperature, and sonication time on the modified clay (DMSO-SMAT) using a 23-factorial design. XRD and FTIR analyses showed the DMSO intercalation into sodium montmorillonite Argel-T (SMAT). An average increase of 0.57 nm for the interplanar distance was found after swelling with DMSO intercalation. BET analysis revealed a decrease in the surface area (from 41.8933 m2/g to 2.1572 m2/g) of Na-Mt when modified with DMSO. The porosity increased from 1.74 (SMAT) to 1.87 nm (DMSO-SMAT) after the application of the methodology. Thermal analysis showed a thermal stability for the DMSO-SMAT material, and this was used to calculate the DMSO-SMAT formula of Na[Al5Mg]Si12O30(OH)6 · 0.54 DMSO. Statistical analysis showed that only the effect of the amount of DMSO was significant for increasing the interlayer space of DMSO-SMAT. In addition, at room temperature, the drying time of the sample using this methodology was 30 min. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of unmodified Mt (SMAT) and modified Mt (DMSO-SMAT) samples.</p>
Full article ">Figure 2
<p>FTIR spectra of modified Mt (DMSO-SMAT) and unmodified Mt (SMAT) samples with emphasis on the region between 960 and 900 cm<sup>−1</sup>.</p>
Full article ">Figure 3
<p>TG and DTG curves for modified Mt (DMSO-SMAT) and unmodified Mt (SMAT) samples.</p>
Full article ">Figure 4
<p>Interaction effects between: (<b>a</b>) temperature and amount of the DMSO; (<b>b</b>) temperature and sonication time and (<b>c</b>) amount of the DMSO and sonication time for the 2<sup>3</sup>-factorial designs for interplanar space increase.</p>
Full article ">Figure 5
<p>Appearance of the samples using the procedure of <a href="#sec2dot2-materials-17-03029" class="html-sec">Section 2.2</a>: (<b>a</b>) 10 min and (<b>b</b>) 30 min of drying with ethanol; (<b>c</b>) 10 min and (<b>d</b>) 30 min without ethanol.</p>
Full article ">
16 pages, 2551 KiB  
Article
From Extraction to Stabilization: Employing a 22 Experimental Design in Developing Nutraceutical-Grade Bixin from Bixa orellana L.
by Christine L. Luna-Finkler, Aralí da C. Gomes, Francisco C. A. de Aguiar Júnior, Ester Ribeiro, Raquel de Melo Barbosa, Patricia Severino, Antonello Santini and Eliana B. Souto
Foods 2024, 13(11), 1622; https://doi.org/10.3390/foods13111622 - 23 May 2024
Viewed by 1422
Abstract
Bixin is the main carotenoid found in the outer portion of the seeds of Bixa orellana L., commercially known as annatto. This compound is industrially employed in pharmaceutical, cosmetic, and food formulations as a natural dye to replace chemical additives. This study aimed [...] Read more.
Bixin is the main carotenoid found in the outer portion of the seeds of Bixa orellana L., commercially known as annatto. This compound is industrially employed in pharmaceutical, cosmetic, and food formulations as a natural dye to replace chemical additives. This study aimed to extract bixin from annatto seeds and obtain encapsulated bixin in a powder form, using freeze-drying encapsulation and maltodextrin as encapsulating agent. Bixin was extracted from annatto seeds employing successive washing with organic solvents, specifically hexane and methanol (1:1 v/v), followed by ethyl acetate and dichloromethane for subsequent washes, to effectively remove impurities and enhance bixin purity, and subsequent purification by crystallization, reaching 1.5 ± 0.2% yield (or approximately 15 mg of bixin per gram of seeds). Bixin was analyzed spectrophotometrically in different organic solvents (ethanol, isopropyl alcohol, dimethylsulfoxide, chloroform, hexane), and the solvents chosen were chloroform (used to solubilize bixin during microencapsulation) and hexane (used for spectrophotometric determination of bixin). Bixin was encapsulated according to a 22 experimental design to investigate the influence of the concentration of maltodextrin (20 to 40%) and bixin-to-matrix ratio (1:20 to 1:40) on the encapsulation efficiency (EE%) and solubility of the encapsulated powder. Higher encapsulation efficiency was obtained at a maltodextrin concentration of 40% w/v and a bixin/maltodextrin ratio of 1:20, while higher solubility was observed at a maltodextrin concentration of 20% w/v for the same bixin/maltodextrin ratio. The encapsulation of this carotenoid by means of freeze-drying is thus recognized as an innovative and promising approach to improve its stability for further processing in pharmaceutical and food applications. Full article
(This article belongs to the Section Food Nutrition)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of bixin ((2E,4E,6E,8E,10E,12E,14E,16Z,18E)-20-methoxy-4,8,13,17-tetramethyl-20-oxoicosa-2,4,6,8,10,12,14,16,18-nonaenoic acid, IUPAC nomenclature).</p>
Full article ">Figure 2
<p>Bibliometric map obtained by VOSviewer software version 1.6.16 (<a href="https://www.vosviewer.com" target="_blank">https://www.vosviewer.com</a>, accessed on 25 March 2024), using “bixin” and “freeze-drying” or “lyophilization” as keywords, recorded from Scopus database (recorded on the 25 March 2024).</p>
Full article ">Figure 3
<p>Schematic representation of bixin extraction and washing out of interferences.</p>
Full article ">Figure 4
<p>Visual appearance of the annatto seeds in natura (<b>A</b>) and the bixin crystals extracted from the seeds (<b>B</b>).</p>
Full article ">Figure 5
<p>Chromatogram of the bixin standard (&gt;90% purity) (<b>a</b>) and the bixin sample obtained after extraction (<b>b</b>).</p>
Full article ">Figure 6
<p>Scanning spectrophotometry of bixin extracted from annatto seeds in the solvents (ethanol, isopropyl alcohol, DMSO, chloroform, and hexane), recorded in the measurement wavelength range from 300 to 800 nm.</p>
Full article ">Figure 7
<p>Contour curves (<b>a</b>) and Pareto diagram (<b>b</b>) for the microencapsulation efficiency response variable as a function of the variables (maltodextrin (1) concentration and maltodextrin/bixin (2) ratio).</p>
Full article ">Figure 8
<p>Contour curves (<b>a</b>) and Pareto diagram (<b>b</b>) for the variable response solubility of the microencapsulated powder as a function of the variables’ maltodextrin (1) concentration and maltodextrin/bixin (2) ratio.</p>
Full article ">
8 pages, 2707 KiB  
Communication
4-(Tris(4-methyl-1H-pyrazol-1-yl)methyl)aniline
by Bradley B. Garrison, Joseph E. Duhamel, Nehemiah Antoine, Steven J. K. Symes, Kyle A. Grice, Colin D. McMillen and Jared A. Pienkos
Molbank 2024, 2024(2), M1823; https://doi.org/10.3390/M1823 - 17 May 2024
Viewed by 1257
Abstract
4-(tris(4-methyl-1H-pyrazol-1-yl)methyl)aniline was prepared in a 63% yield utilizing a C–F activation strategy from a mixture of 4-(trifluoromethyl)aniline, 4-methylpyrazole, and KOH in dimethylsulfoxide (DMSO). The identity of the product was confirmed by nuclear magnetic resonance spectroscopy, infrared spectroscopy, mass spectrometry, and single-crystal [...] Read more.
4-(tris(4-methyl-1H-pyrazol-1-yl)methyl)aniline was prepared in a 63% yield utilizing a C–F activation strategy from a mixture of 4-(trifluoromethyl)aniline, 4-methylpyrazole, and KOH in dimethylsulfoxide (DMSO). The identity of the product was confirmed by nuclear magnetic resonance spectroscopy, infrared spectroscopy, mass spectrometry, and single-crystal analysis. An analysis of crystals grown from the layering method (CH2Cl2/acetone/pentane) indicated two distinct polymorphs of the title compound. Moreover, density functional theory calculations utilizing the MN15L density functional and the def2-TZVP basis set indicated that 4-(tris(4-methyl-1H-pyrazol-1-yl)methyl)aniline forms with similar energetics to the previously reported unmethylated analog. Full article
(This article belongs to the Section Organic Synthesis and Biosynthesis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Tris(pyrazolyl)methane (<b>a</b>), tris(pyrazolyl)borate (<b>b</b>), and C–F activation strategy (<b>c</b>) exploited by Liddle et al. [<a href="#B13-molbank-2024-M1823" class="html-bibr">13</a>].</p>
Full article ">Figure 2
<p>Structure of the compound associated with the base peak in the mass spectrum.</p>
Full article ">Figure 3
<p>Crystal structure of polymorphs I and II of compound <b>1</b>. The molecules are oriented similarly with respect to the central carbon atom and the aniline substituent. Ellipsoids are shown at a 50% probability level. The overlaid wireframe structures of the polymorphs (polymorph I in orange and polymorph II in purple) again keeps a consistent orientation of the central carbon atom and the aniline substituent.</p>
Full article ">Figure 4
<p>Selected intermolecular interactions (N-H···N as dashed blue lines) in polymorphs I and II of compound <b>1</b>.</p>
Full article ">Figure 5
<p>Comparison of energies of two C–F activation reactions.</p>
Full article ">Figure 6
<p>Transition state calculation involved in compound formation.</p>
Full article ">Scheme 1
<p>Synthesis of title compound.</p>
Full article ">
Back to TopTop