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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (3,121)

Search Parameters:
Keywords = 1H NMR analysis

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 5170 KiB  
Article
Green Sulfation of Arabinogalactan in the Melt of a Sulfamic Acid–Urea Mixture
by Vladimir A. Levdansky, Alexander V. Levdansky, Yuriy N. Malyar, Timur Yu. Ivanenko, Olga Yu. Fetisova, Aleksandr S. Kazachenko and Boris N. Kuznetsov
Polymers 2025, 17(5), 642; https://doi.org/10.3390/polym17050642 (registering DOI) - 27 Feb 2025
Abstract
Sulfation of arabinogalactan (AG) from larch wood (Larix sibirica Ledeb.) in the melt of a sulfamic acid–urea mixture has been first examined. The impact of the AG sulfation temperature on the AG sulfate yield and the sulfur content has been established. [...] Read more.
Sulfation of arabinogalactan (AG) from larch wood (Larix sibirica Ledeb.) in the melt of a sulfamic acid–urea mixture has been first examined. The impact of the AG sulfation temperature on the AG sulfate yield and the sulfur content has been established. The high sulfur content (11.3–11.6%) in sulfated AG has been obtained in the temperature range of 115–120 °C for a sulfation time of 0.5 h. The process effectively prevents molecular degradation under these conditions. The incorporation of sulfate groups into the arabinogalactan structure has been confirmed by the appearance of absorption bands in the FTIR spectrum that are typical of sulfate group vibrations. The 13C NMR spectroscopy study has proven that the AG sulfation in the melt of a sulfamic acid–urea mixture leads to the substitution of some free hydroxyl groups for C6, C4, and C2 carbon atoms of the AG β-D-galactopyranose units. The advantage of the proposed AG sulfation method is that the reaction occurs without solvent, and the reaction time is only 0.5 h. The kinetics of the thermal decomposition of the initial AG and sulfated AG samples have been studied. It has been found that the sulfated AG samples have a lower thermal resistance than the initial AG. The kinetic analysis has revealed a decrease in the activation energy of the thermal degradation of the sulfated samples as compared to the initial AG. Full article
(This article belongs to the Special Issue Degradation and Stability of Polymer-Based Systems: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic of arabinogalactan sulfation with the melt of sulfamic acid–urea mixture.</p>
Full article ">Figure 2
<p>FTIR spectra of AG (1) and AG sulfate with a sulfur content of 11.6 wt % (2).</p>
Full article ">Figure 3
<p><sup>13</sup>C NMR spectrum of arabinogalactan.</p>
Full article ">Figure 4
<p>Main structural units of AG macromolecule.</p>
Full article ">Figure 5
<p>HSQC spectrum of arabinogalactan. The red color shows the proton-carbon correlations in –CH<sub>2</sub>– groups while the blue color shows the proton-carbon correlations in –CH– and –CH<sub>3</sub> groups.</p>
Full article ">Figure 6
<p><sup>13</sup>C NMR spectrum of the sulfated AG sample (a sulfur content of 11.6 wt %) obtained in the melt of a sulfamic acid–urea mixture.</p>
Full article ">Figure 7
<p>HSQC spectrum of the sulfated AG sample (a sulfur content of 11.6 wt %) obtained in the melt of a sulfamic acid–urea mixture. The red color shows the proton-carbon correlations in –CH<sub>2</sub>– groups while the blue color shows the proton-carbon correlations in –CH– and –CH<sub>3</sub> groups.</p>
Full article ">Figure 8
<p>Molecular weight distributions for initial AG (1) and its sulfated derivatives obtained at temperatures of 110 (2), 115 (3), 120 (4), 125 (5), 130 (6), and 140 °C (7).</p>
Full article ">Figure 9
<p>TG (<b>a</b>) and DTG (<b>b</b>) curves of thermal decomposition of initial AG (1) and sulfated AG obtained at temperatures of 110 (2), 125 (3), and 140 °C (4).</p>
Full article ">Figure 10
<p>DSC curves of thermal decomposition of initial AG (1) and sulfated AG obtained at temperatures of 110 (2), 125 (3), and 140 °C (4).</p>
Full article ">Figure 11
<p>Coats—Redfern linearization of the TGA results. AG (1) and sulfated AG obtained at temperatures of 110 (2), 125 (3), and 140 °C (4).</p>
Full article ">
19 pages, 5187 KiB  
Article
Self-Healing Hydrogels with Intrinsic Antioxidant and Antibacterial Properties Based on Oxidized Hydroxybutanoyl Glycan and Quaternized Carboxymethyl Chitosan for pH-Responsive Drug Delivery
by Jae-pil Jeong, Kyungho Kim, Eunkyung Oh, Sohyun Park and Seunho Jung
Gels 2025, 11(3), 169; https://doi.org/10.3390/gels11030169 - 26 Feb 2025
Viewed by 99
Abstract
In this study, self-healing hydrogels were created using oxidized hydroxybutanoyl glycan (OHbG) and quaternized carboxymethyl chitosan (QCMCS), displaying antioxidant and antibacterial properties for pH-responsive drug delivery. The structures of the modified polysaccharides were confirmed through 1H NMR analysis. Double crosslinking in the [...] Read more.
In this study, self-healing hydrogels were created using oxidized hydroxybutanoyl glycan (OHbG) and quaternized carboxymethyl chitosan (QCMCS), displaying antioxidant and antibacterial properties for pH-responsive drug delivery. The structures of the modified polysaccharides were confirmed through 1H NMR analysis. Double crosslinking in the hydrogel occurred via imine bonds (between the aldehyde group of OHbG and the amine group of QCMCS) and ionic interactions (between the carboxyl group of OHbG and the quaternized group of QCMCS). The hydrogel exhibited self-healing properties and improved thermal stability with an increase in OHbG concentration. The OHbG/QCMCS hydrogel demonstrated high compressive strength, significant swelling, and large pore size. Drug release profiles varied between pH 2.0 (96.57%) and pH 7.4 (63.22%). Additionally, the hydrogel displayed antioxidant and antibacterial effects without compromising the polysaccharides’ inherent characteristics. No cytotoxicity was observed in any hydrogel samples. These findings indicate that the OHbG/QCMCS hydrogel is a biocompatible and stimuli-responsive drug carrier, with potential for various pharmaceutical, biomedical, and biotechnological applications. Full article
(This article belongs to the Special Issue Recent Advances in Gels Engineering for Drug Delivery (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><sup>1</sup>H NMR spectrum of (<b>a</b>) oxidized 3-hydroxylbutanoyl glycan (OHbG) and (<b>b</b>) quaternized carboxymethyl chitosan (QCMCS).</p>
Full article ">Figure 2
<p>(<b>a</b>) FTIR spectra from 4000 to 650 cm<sup>−1</sup> of OHbG/QCMCS hydrogels. (<b>b</b>) FTIR spectra from 2000 to 650 cm<sup>−1</sup> of OHbG/QCMCS hydrogels. (<b>c</b>) DSC curves of OHbG/QCMCS hydrogels.</p>
Full article ">Figure 3
<p>(<b>a</b>) Angular frequency sweep test (0.1 rad/s to 100 rad/s) of OHbG/QCMCS hydrogels at a constant strain (1%). (<b>b</b>) Oscillation strain amplitude sweep test (from 0.1% to 1000%) of OHbG/QCMCS hydrogels at a constant angular frequency (1 Hz). (<b>c</b>) Compressive stress–strain test of OHbG/QCMCS hydrogels.</p>
Full article ">Figure 4
<p>Self-healing ability via the step strain measurements of OHbG/QCMCS hydrogels. (<b>a</b>) OHbG/QCMCS 5, (<b>b</b>) OHbG/QCMCS 7, (<b>c</b>) OHbG/QCMCS 9, and (<b>d</b>) OHbG/QCMCS 11. (<b>e</b>) Photo of self-healed OHbG/QCMCS hydrogel through the integration of two cleaved hydrogels with different colors.</p>
Full article ">Figure 5
<p>(<b>a</b>) Swelling ratio of OHbG/QCMCS hydrogels at physiological conditions (pH 7.4) and (<b>b</b>) the percentage of the remaining weight of OHbG/QCMCS hydrogels at acidic conditions (pH 2.0).</p>
Full article ">Figure 6
<p>SEM image of OHbG/QCMCS hydrogels. The white scale bar means 100 μm.</p>
Full article ">Figure 7
<p>pH-dependent drug release of OHbG/QCMCS hydrogels: cumulative drug release of 5-FU under (<b>a</b>) pH 2.0 at 37 °C, (<b>b</b>) pH 7.4 at 37 °C, (<b>c</b>) pH shock from 7.4 to 2.0. In the pH shock test, the pH was adjusted to 2.0 at 9 hours after the release began at an initial pH of 7.4.</p>
Full article ">Figure 8
<p>DPPH-scavenging activity of OHbG/QCMCS hydrogels.</p>
Full article ">Figure 9
<p>(<b>a</b>) Photographs of the antibacterial effect of OHbG/QCMCS hydrogels against <span class="html-italic">E. coli</span> and <span class="html-italic">S. aureus.</span> (<b>b</b>) Survival ratio of OHbG/QCMCS hydrogels against <span class="html-italic">E. coli</span> and <span class="html-italic">S. aureus</span> (n = 3).</p>
Full article ">Figure 10
<p>Cytotoxicity test of OHbG/QCMCS hydrogels against HEK-293 cell.</p>
Full article ">
11 pages, 1256 KiB  
Article
Structural Characterization of Pinnatoxin Isomers
by Andrew I. Selwood, Christopher O. Miles, Alistair L. Wilkins, Frode Rise, Sarah C. Finch and Roel van Ginkel
Mar. Drugs 2025, 23(3), 103; https://doi.org/10.3390/md23030103 - 26 Feb 2025
Viewed by 191
Abstract
Pinnatoxins, a group of marine biotoxins primarily produced by the dinoflagellate Vulcanodinium rugosum, have garnered significant attention due to their potent toxic effects and widespread distribution in marine ecosystems. LC–MS analysis of shellfish and V. rugosum cultures revealed the presence of previously [...] Read more.
Pinnatoxins, a group of marine biotoxins primarily produced by the dinoflagellate Vulcanodinium rugosum, have garnered significant attention due to their potent toxic effects and widespread distribution in marine ecosystems. LC–MS analysis of shellfish and V. rugosum cultures revealed the presence of previously unidentified isomers of pinnatoxins D, E, F, and H, at levels approximately six times lower than those of known isomers. The chemical structures of these isopinnatoxins were determined using a combination of LC–MS/MS and NMR spectroscopy, which demonstrated that the isomerization of each pinnatoxin occurred through the opening and recyclization of the spiro-linked tetrahydropyranyl D-ring to form a smaller tetrahydrofuranyl ring. The acute toxicity of isopinnatoxin E was determined by intraperitoneal injection into mice and was found to be significantly lower than that of pinnatoxin E. Given their low toxicity and low abundance, it is unlikely that isopinnatoxins contribute significantly to the overall toxicity of pinnatoxins. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of pinnatoxins and spirolide C. The substructures colored blue and red, and orange and pink indicate the two interconverting substructures of these molecules (see Figure 4).</p>
Full article ">Figure 2
<p>Collison-induced fragmentation spectra of <b>5</b> (<b>a</b>) and <b>1</b> (<b>b</b>).</p>
Full article ">Figure 3
<p>Structurally important HMBC (colored arrows) and NOESY (black arrow) correlations linking ether rings in compound <b>5</b>.</p>
Full article ">Figure 4
<p>Mole fractions (%) of the four structural forms of pinnatoxin E (<b>1</b>) in the pinnatoxin (<b>1</b> and <b>2</b>) versus isopinnatoxin (<b>5</b> and <b>6</b>) forms, and in the lactone (<b>2</b> and <b>6</b>) versus γ-hydroxycarboxylic acid (<b>1</b> and <b>5</b>) forms, with time, obtained after the addition of isopinnatoxin E (<b>5</b>) to aqueous acetonitrile containing TFA. Data were fitted to three-parameter exponential curves to provide equilibrium positions (<a href="#marinedrugs-23-00103-t002" class="html-table">Table 2</a>) and half-lives for the isomerization reactions. Similar experiments were also performed to evaluate the equilibration of <b>1</b>, <b>2</b>, <b>3</b>, <b>4</b>, and <b>9</b> (<a href="#marinedrugs-23-00103-t002" class="html-table">Table 2</a>, <a href="#app1-marinedrugs-23-00103" class="html-app">Figures S10–S14</a>).</p>
Full article ">Figure 5
<p>LC–MS/MS MRM chromatograms of extracts from three strains of <span class="html-italic">V. rugosum</span> showing pinnatoxin profiles: (<b>a</b>) strain CAWD167, containing pinnatoxins E (<b>1</b>), F (<b>2</b>), and their isomers (<b>5</b> and <b>6</b>); (<b>b</b>) CAWD188, contains pinnatoxin G (<b>9</b>), and; (<b>c</b>) CAWD198, containing pinnatoxin H (<b>4</b>) and its isomer (<b>8</b>). In addition to the spiroketal rearrangement, the γ-butyrolactone–γ-hydroxybutanoic acid moiety in compounds <b>1</b>, <b>2</b>, <b>5,</b> and <b>6</b> opened and closed to an equilibrated mixture of 94:6 lactone–hydroxyacid (<a href="#marinedrugs-23-00103-t002" class="html-table">Table 2</a>, <a href="#marinedrugs-23-00103-f004" class="html-fig">Figure 4</a> and <a href="#app1-marinedrugs-23-00103" class="html-app">Figure S15</a>). Due to the occurrence of these acid-catalyzed rearrangements, compounds <b>1</b>, <b>2,</b> and <b>5</b> ended up as an equilibrated mixture of compounds <b>2</b>, <b>6</b>, <b>1</b>, and <b>5</b> at a ratio of approximately 80:14:5:1. The rate of the D-ring rearrangement was similar to or perhaps slightly faster than the ring closure of the hydroxy acid to the lactone under these conditions (<a href="#marinedrugs-23-00103-f004" class="html-fig">Figure 4</a> and <a href="#app1-marinedrugs-23-00103" class="html-app">Figures S10–S14</a>).</p>
Full article ">
18 pages, 6389 KiB  
Article
Synthesis, Physicochemical Properties and Anti-Fungal Activities of New Meso-Arylporphyrins
by Hayfa Mkacher, Raja Chaâbane-Banaoues, Soukaina Hrichi, Philippe Arnoux, Hamouda Babba, Céline Frochot, Habib Nasri and Samir Acherar
Int. J. Mol. Sci. 2025, 26(5), 1991; https://doi.org/10.3390/ijms26051991 - 25 Feb 2025
Viewed by 157
Abstract
In this work, we describe the synthesis of three new meso-arylporphyrins, named meso-tetrakis [4-(nicotinoyloxy)phenyl] porphyrin (H2TNPP), meso-tetrakis [4-(picolinoyloxy)phenyl] porphyrin (H2TPPP), and meso-tetrakis [4-(isonicotinoyloxy) phenyl] porphyrin (H2TIPP). These [...] Read more.
In this work, we describe the synthesis of three new meso-arylporphyrins, named meso-tetrakis [4-(nicotinoyloxy)phenyl] porphyrin (H2TNPP), meso-tetrakis [4-(picolinoyloxy)phenyl] porphyrin (H2TPPP), and meso-tetrakis [4-(isonicotinoyloxy) phenyl] porphyrin (H2TIPP). These new synthesized meso-arylporphyrins are characterized using spectroscopic analysis: Fourier Transform Infrared Spectroscopy (FTIR) and One-dimensional Nuclear Magnetic Resonance (1D NMR), and mass spectrometry (MS). The photophysical studies (UV–visible absorption, singlet oxygen (1O2) luminescence, and fluorescence emissions) demonstrate their potential uses as photosensitizers (PSs) in photodynamic therapy (PDT) applications. An in vitro investigation of the anti-fungal activity of H2TNPP, H2TPPP, and H2TIPP against Candida (C.) species (C. albicans, C. glabrata, and C. tropicalis) reveals that their minimum inhibitory concentration (MIC) values ranged from 1.25 to 5 mg/mL. In addition, their in vitro anti-fungal susceptibilities against three dermatophyte clinical isolates (Trichophyton rubrum, Microsporum canis, and Trichophyton mentagrophytes) are also evaluated and they demonstrate good anti-fungal activities. A molecular docking study of these meso-arylporphyrins as anti-fungal agents against C. tropicalis extracellular aspartic proteinases, Protein data Bank in Europe (PDBe code: 1J71) and Trichophyton rubrum Sialidases (PDBe code: 7P1D) underlines the possible interactions of H2TNPP, H2TPPP, and H2TIPP with the key amino acid residues of these fungal target proteins. Full article
(This article belongs to the Special Issue Advances in Research on Antifungal Resistance)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of the three new aldehydes <b>AL1</b>, <b>AL2</b>, and <b>AL3</b> (<b>a</b>) and <span class="html-italic">meso</span>-arylporphyrins <b>H<sub>2</sub>TNPP</b>, <b>H<sub>2</sub>TPPP</b>, and <b>H<sub>2</sub>TIPP</b> (<b>b</b>).</p>
Full article ">Figure 2
<p>In vitro anti-fungal activity of <b>H<sub>2</sub>TNPP</b>, <b>H<sub>2</sub>TPPP</b>, and <b>H<sub>2</sub>TIPP</b> against the three dermatophyte strains.</p>
Full article ">Figure 3
<p><b>Three-dimensional</b> structure image of <b>H<sub>2</sub>TIPP</b>, <b>H<sub>2</sub>TNPP</b>, and <b>H<sub>2</sub>TPPP</b> in anti-fungal complex with <span class="html-italic">C. tropicalis</span> (PDBe: 1j71) (<b>a</b>) or <span class="html-italic">T. rubrum</span> (PDBe: 7P1D) (<b>b</b>) protein using Hermes program.</p>
Full article ">Figure 4
<p>Two-dimensional representation illustrating bond lengths, amino acid residues, and the nature of interactions present with <b>H<sub>2</sub>TIPP</b>, <b>H<sub>2</sub>TNPP</b>, and <b>H<sub>2</sub>TPPP</b> in the active site of <span class="html-italic">C. tropicalis</span> (PDBe: 1j71) (<b>a</b>) and <span class="html-italic">T. rubrum</span> (PDBe: 7P1D) (<b>b</b>) using BIOVIA Discovery Studio Visualizer.</p>
Full article ">Scheme 1
<p>Main chemical preparation steps for aldehydes <b>AL1</b>, <b>AL2</b>, and <b>AL3</b>.</p>
Full article ">Scheme 2
<p>Main chemical preparation steps of <span class="html-italic">meso</span>-arylporphyrins <b>H<sub>2</sub>TNPP</b>, <b>H<sub>2</sub>TPPP</b>, and <b>H<sub>2</sub>TIPP</b>.</p>
Full article ">
19 pages, 6778 KiB  
Article
Beyond the Phenothiazine Core: Mechanistic Insights into the Three-Electron Oxidation of Chlorpromazine
by Kiara T. Miller, Ashwin K. V. Mruthunjaya and Angel A. J. Torriero
Molecules 2025, 30(5), 1050; https://doi.org/10.3390/molecules30051050 - 25 Feb 2025
Viewed by 149
Abstract
This study investigates the electrochemical oxidation mechanisms of chlorpromazine (CPZ), revealing a novel three-electron oxidation pathway that challenges the traditionally accepted two-electron paradigm, offering new insights into CPZ oxidation pathways. Using an integrated approach combining cyclic voltammetry, bulk electrolysis, UV-Vis, FT-IR, 1H-NMR [...] Read more.
This study investigates the electrochemical oxidation mechanisms of chlorpromazine (CPZ), revealing a novel three-electron oxidation pathway that challenges the traditionally accepted two-electron paradigm, offering new insights into CPZ oxidation pathways. Using an integrated approach combining cyclic voltammetry, bulk electrolysis, UV-Vis, FT-IR, 1H-NMR spectroscopy, and LC-MS/MS analysis, we demonstrate that CPZ undergoes sequential oxidation processes involving both the phenothiazine core and the tertiary amine-containing side chain. Our results highlight the critical role of side-chain oxidation in forming nor-CPZ sulfoxide, an often-overlooked metabolite, which may influence CPZ’s metabolic and pharmacological behaviour. Spectroelectrochemical data reveal stable intermediate species, providing insight into the structural rearrangements accompanying oxidation. This work offers a detailed mechanistic understanding of CPZ redox behaviour, contributing to improved interpretations of its pharmacological and metabolic properties. Full article
Show Figures

Figure 1

Figure 1
<p>Cyclic voltammograms recorded at a 1.0 mm diameter GC electrode in CH<sub>3</sub>CN containing 0.1 M [Bu<sub>4</sub>N][PF<sub>6</sub>] as the supporting electrolyte: (<b>A</b>) at a scan rate of 0.1 Vs<sup>−1</sup> in the presence of 2.61 mM PTZ; (<b>B</b>) in the presence of 2.61 mM PTZ at scan rates of (<span class="html-italic">a</span>) 0.05, (<span class="html-italic">b</span>) 0.10, (<span class="html-italic">c</span>) 0.20, (<span class="html-italic">d</span>) 0.30, (<span class="html-italic">e</span>) 0.40, (<span class="html-italic">f</span>) 0.50, (<span class="html-italic">g</span>) 0.70, and (<span class="html-italic">h</span>) 1.00 Vs<sup>−1</sup>; (<b>C</b>) at a scan rate of 0.1 Vs<sup>−1</sup> in the absence (<span class="html-italic">a</span>) and presence (<span class="html-italic">b</span>, <span class="html-italic">c</span>) of 1.27 mM CPZ, with switching potentials of (<span class="html-italic">a</span>) 1.46, (<span class="html-italic">b</span>) 1.04, and (<span class="html-italic">c</span>) 1.38 V vs. DmFc<sup>0/+</sup>; and (<b>D</b>) at a scan rate of 0.1 Vs<sup>−1</sup> in the presence of 1.27 mM CPZ under air-saturated conditions (<span class="html-italic">a</span>) and nitrogen (<span class="html-italic">b</span>) atmosphere. T = 21 ± 1 °C.</p>
Full article ">Figure 2
<p>(<b>A</b>) Cyclic voltammograms of 1.65 mM CPZ in CH<sub>3</sub>CN containing 0.1 M [Bu<sub>4</sub>N][PF<sub>6</sub>] as the supporting electrolyte, recorded at a 1.0 mm diameter GC electrode and scan rates of (<span class="html-italic">a</span>) 0.02, (<span class="html-italic">b</span>) 0.05, (<span class="html-italic">c)</span> 0.10, (<span class="html-italic">d</span>) 0.20, (<span class="html-italic">e</span>) 0.30, (<span class="html-italic">f)</span> 0.40, (<span class="html-italic">g</span>) 0.50, (<span class="html-italic">h</span>) 0.70, and (<span class="html-italic">i</span>) 1.00 Vs<sup>−1</sup>. (<b>B</b>) Dependence of current function (Ψ) on scan rate for CPZ under nitrogen (<span class="html-italic">a</span>) and air (<span class="html-italic">b</span>) atmosphere. The DmFc<sup>0/+</sup> process is used as the reversible one-electron transfer reference compound. T = 21 ± 1 °C.</p>
Full article ">Figure 3
<p>Electrochemical data obtained for oxidation of 1.0 mM CPZ in CH<sub>3</sub>CN containing 0.1 M [Bu<sub>4</sub>N][PF<sub>6</sub>] as the supporting electrolyte in the absence and presence of pyridine: (<b>A</b>) cyclic voltammograms recorded at 1.0 mm diameter GC electrode at a scan rate of 0.10 Vs<sup>−1</sup> in the absence (<span class="html-italic">a</span>) and presence of (<span class="html-italic">b)</span> 0.2, (<span class="html-italic">c</span>) 0.4, (<span class="html-italic">d</span>) 0.6, (<span class="html-italic">e</span>) 0.8, (<span class="html-italic">f</span>) 1.0, (<span class="html-italic">g</span>) 1.6 equivalents of pyridine, and (<span class="html-italic">h</span>) 0.2 mM pyridine in the absence of CPZ; and (<b>B</b>) dependence of current change (∆<span class="html-italic">I</span> = peak current of process II obtained from cyclic voltammetry) for CPZ upon addition of designated concentrations of pyridine. T = 21 ± 1 °C.</p>
Full article ">Figure 4
<p>Cyclic voltammograms of 4.74 mM CPZ in CH<sub>3</sub>CN containing 0.1 M [Bu<sub>4</sub>N][PF<sub>6</sub>] as the supporting electrolyte obtained before (<span class="html-italic">a</span>) and after (<span class="html-italic">b</span>) exhaustive controlled-potential bulk electrolysis at (<b>A</b>) <span class="html-italic">E</span><sub>app</sub> = 0.950 V vs. DmFc<sup>0/+</sup> and (<b>B</b>) <span class="html-italic">E</span><sub>app</sub> = 1.250 V vs. DmFc<sup>0/+</sup>. The arrows indicate the potential scan direction. Scan rate = 0.10 Vs<sup>−1</sup> T = 21 ± 1 °C.</p>
Full article ">Figure 5
<p>UV-Vis spectroelectrochemical experiment for oxidation of 0.39 mM CPZ in CH<sub>3</sub>CN containing 0.1 M [Bu<sub>4</sub>N][PF<sub>6</sub>] as the supporting electrolyte: scans recorded every 1 min during electrolysis at (<b>A</b>) <span class="html-italic">E</span><sub>app</sub> = 0.950 V vs. DmFc<sup>0/+</sup> and (<b>B</b>) <span class="html-italic">E</span><sub>app</sub> = 1.250 V vs. DmFc<sup>0/+</sup>. T = 21 ± 1 °C. Inset: 3D plot of data in the 350–900 nm range that enhances the bands at 526, 773, and 861 nm.</p>
Full article ">Figure 6
<p>FT-IR spectroelectrochemical monitoring of the oxidation of 60 mM CPZ in CH<sub>3</sub>CN containing 0.1 M [Bu<sub>4</sub>N][PF<sub>6</sub>] as the supporting electrolyte. Samples of BE solution were drop-cast on KBr disks. (<b>A</b>) Full FT-IR spectra recorded before and during BE at <span class="html-italic">E</span><sub>app</sub> = 1.250 V vs. DmFc<sup>0/+</sup>, T = 21 ± 1 °C: (<span class="html-italic">a</span>) spectra of CPZ mixed with [Bu<sub>4</sub>N][PF<sub>6</sub>] and (<span class="html-italic">b</span>) [Bu<sub>4</sub>N][PF<sub>6</sub>] recorded before BE, (<span class="html-italic">c</span>–<span class="html-italic">f</span>) correspond to spectra recorded after 1, 3, 5, and 10 min of BE, respectively, and (<span class="html-italic">g</span>) represents the spectrum at the end of the BE. (<b>B</b>) Expanded view of the 1600–1520 cm<sup>−1</sup> region from the spectra shown in (A).</p>
Full article ">Figure 7
<p>Comparison of <sup>1</sup>H-NMR spectra in D<sub>2</sub>O before and after electrochemical oxidation of CPZ.HCl: (<b>A</b>) <sup>1</sup>H-NMR spectrum of the untreated CPZ.HCl and (<b>B</b>) <sup>1</sup>H-NMR spectrum of the products obtained following exhaustive controlled-potential bulk electrolysis at <span class="html-italic">E</span><sub>app</sub> = 1.268 V vs. DmFc<sup>0/+</sup>. Reported chemical shift values correspond to the average (mean) values for multiplets.</p>
Full article ">
12 pages, 2388 KiB  
Article
Acyclic Cucurbit[n]uril-Enabled Detection of Aflatoxin B1 via Host–Guest Chemistry and Bioluminescent Immunoassay
by Shaowen Wu, Ke Feng, Jinlu Niu, Jintao Xu, Hualian Mo, Xiaoman She, Shang-Bo Yu, Zhan-Ting Li and Shijuan Yan
Toxins 2025, 17(3), 104; https://doi.org/10.3390/toxins17030104 - 25 Feb 2025
Viewed by 158
Abstract
Aflatoxin B1 (AFB1), a highly toxic secondary metabolite produced by Aspergillus species, represents a significant health hazard due to its widespread contamination of agricultural products. The urgent need for sensitive and sustainable detection methods has driven the development of diverse analytical approaches, most [...] Read more.
Aflatoxin B1 (AFB1), a highly toxic secondary metabolite produced by Aspergillus species, represents a significant health hazard due to its widespread contamination of agricultural products. The urgent need for sensitive and sustainable detection methods has driven the development of diverse analytical approaches, most of which heavily rely on organic solvents, posing environmental challenges for routine food safety analysis. Here, we introduce a supramolecular platform leveraging acyclic cucurbit[n]uril (acCB) as a host molecule for environmentally sustainable AFB1 detection. Screening various acCB derivatives identified acCB6 as a superior host capable of forming a stable 1:1 complex with AFB1 in an aqueous solution, exhibiting a high binding affinity. Proton nuclear magnetic resonance (1H NMR) spectroscopy confirmed that AFB1 was deeply encapsulated within the host cavity, with isothermal titration calorimetry (ITC) experiments and molecular dynamics simulations further substantiating the stability of the interaction, driven by enthalpic and entropic contributions. This supramolecular host was incorporated into a scaffold-assembly-based bioluminescent enzyme immunoassay (SA-BLEIA), providing a green detection platform that rivals the performance of traditional organic solvent-based assays. Our findings highlight the potential of supramolecular chemistry as a foundation for eco-friendly mycotoxin detection and offer valuable insights into designing environmentally sustainable analytical methods. Full article
(This article belongs to the Special Issue Aspergillus flavus and Aflatoxins (Volume III))
Show Figures

Figure 1

Figure 1
<p>Chemical structures of acyclic cucurbit[n]urils and their ability to enhance the solubility of aflatoxin B1 (AFB1). (<b>a</b>) Chemical structure of AFB1 alongside the structures of acCB derivatives, illustrating their substituents and cavity dimensions. (<b>b</b>,<b>c</b>) Visual representation of AFB1 dissolution enhancement in aqueous solution. Sample vials containing AFB1 with various acCBs (vials 1–7 corresponding to acCB 1–7) at 0 molar equivalents (<b>b</b>) and 1.0 molar equivalents (<b>c</b>). Top and side views of the vials are provided.</p>
Full article ">Figure 2
<p>NMR, UV-vis, and fluorescence spectroscopic characterization of the acCB6-AFB1 complex. (<b>a</b>) <sup>1</sup>H NMR titration spectra showing the evolution of chemical shift changes in the 3.6–5.0 ppm region upon the stepwise addition of acCB6 (0.8–25.6 mM, 0.25–8.0 equivalents) to AFB1 (3.2 mM). (<b>b</b>) Signal shift analysis for A and B as a function of [acCB6]/[AFB1] ratio, highlighting a 1:1 binding stoichiometry. (<b>c</b>) Comparative <sup>1</sup>H NMR spectra of free acCB6 (3.2 mM, D<sub>2</sub>O), free AFB1 (3.2 mM, DMSO-d6), and their 1:1 mixture (3.2 mM, D<sub>2</sub>O) at 25 °C, with key proton assignments of AFB1 (a–g) indicated. (<b>d</b>) UV–visible absorption spectra and (<b>e</b>) fluorescence emission spectra of acCB6 (1 μM), AFB1 (1 μM), and their 1:1 complex (1 μM) in H<sub>2</sub>O at 25 °C, demonstrating spectral changes upon complex formation.</p>
Full article ">Figure 3
<p>ITC and molecular dynamics analyses of acCB6-AFB1 interactions. (<b>a</b>) ITC titration curve showing raw calorimetric data and integrated heat changes for the addition of acCB6 (0.1 mM) to AFB1 (0.01 mM) at 25 °C. The solid line represents the best fit to a 1:1 binding model. (<b>b</b>) Molecular dynamics simulation results, including an RMSD plot of the acCB6-AFB1 complex (top) and the center-of-mass distance between acCB6 and AFB1 (bottom) over 1000 ns across three independent trajectories. (<b>c</b>) Representative structures from cluster analysis of MD trajectories showing three dominant binding modes (Clusters 1–3). AFB1 is depicted in cyan, acCB6 in gray, with polar interactions (magenta) and π-π stacking interactions (gray) shown as dashed lines. Polar O/N atoms are colored red in AFB1 and blue in acCB6. The host cavity surface is rendered in transparent white to illustrate the encapsulation of AFB1.</p>
Full article ">Figure 4
<p>Performance of the scaffold-assembly-based bioluminescent enzyme immunoassay (SA-BLEIA) analysis with AFB1 standards dissolved in different solutions. (<b>a</b>) Schematic representation of the SA-BLEIA detection principle, comparing traditional methanol-dissolved AFB1 standards with acCB6-dissolved AFB1 standards. (<b>b</b>) Standard curves illustrating the detection performance of methanol-dissolved AFB1 (gray) and acCB6-dissolved AFB1 (blue) standards, highlighting differences in sensitivity. (<b>c</b>) Influence of ionic strength on SA-BLEIA performance using acCB6-dissolved AFB1 standards at varying NaCl concentrations (50, 100, and 200 mM).</p>
Full article ">
20 pages, 2965 KiB  
Article
Morpholine-Substituted Tetrahydroquinoline Derivatives as Potential mTOR Inhibitors: Synthesis, Computational Insights, and Cellular Analysis
by Rajdeep Dey, Suman Shaw, Ruchi Yadav, Bhumika D. Patel, Hardik G. Bhatt, Gopal Natesan, Abhishek B. Jha and Udit Chaube
Cancers 2025, 17(5), 759; https://doi.org/10.3390/cancers17050759 - 23 Feb 2025
Viewed by 381
Abstract
Backgrounds: This study explores the design of substituted tetrahydroquinoline (THQ) derivatives and their synthesis as possible inhibitors of mTOR inhibitors for targeted cancer therapy. Methods: Inspired by the structural characteristics of known mTOR inhibitors, eight novel derivatives were synthesized, characterized using mass spectroscopy, [...] Read more.
Backgrounds: This study explores the design of substituted tetrahydroquinoline (THQ) derivatives and their synthesis as possible inhibitors of mTOR inhibitors for targeted cancer therapy. Methods: Inspired by the structural characteristics of known mTOR inhibitors, eight novel derivatives were synthesized, characterized using mass spectroscopy, 1H, and 13C NMR, and evaluated for anticancer activity. Results: Computational studies, including molecular docking and molecular dynamics (MD) simulations, highlighted the derivative’s strong binding interaction and stability within the mTOR active site. Assays for in vitro cytotoxicity showed strong and specific anticancer action against cell lines of triple-negative breast cancer, lung cancer, and breast cancer while causing negligible impact on healthy cells. Conclusions: Compound 10e emerged as the most promising candidate, displaying exceptional activity against A549 cells (IC50 = 0.033 µM) and inducing apoptosis in a dose-dependent manner, surpassing standard agents, like Everolimus and 5-flurouracil. Structure–activity relationship analysis revealed that incorporating trifluoromethyl and morpholine moieties significantly enhanced selectivity and potency. MD simulations further validated these findings, confirming stable protein-ligand interactions and favorable dynamics over a 100-ns simulation period. Collectively, this study underscores the therapeutic potential of THQ derivatives, particularly compound 10e, as promising mTOR inhibitors with potential applications in lung cancer treatment. Full article
(This article belongs to the Section Cancer Drug Development)
Show Figures

Figure 1

Figure 1
<p>Strategic design of tetrahydroquinoline analogs, green circle represents morpholine-like substitutions that are crucial for mTOR activity.</p>
Full article ">Figure 2
<p>Synthetic scheme and procedure: reagents and Conditions: (<b>i</b>) nitration: KNO<sub>3</sub>/H<sub>2</sub>SO<sub>4</sub> in DCM, stirred at 0 °C to room temperature (RT) for 2 h, followed by deprotection with pyrrolidine at RT. (<b>ii</b>) Amidation/coupling: triethylamine in DCM, stirred for 24 h. (<b>iii</b>) Reduction: Zn/NH<sub>4</sub>Cl in methanol, heated at 60 °C for 6 h. (<b>iv</b>) Final step: triethylamine in DCM, stirred for 1–2 h.</p>
Full article ">Figure 3
<p>SAR insights: effect of functional groups on the activity of synthesized THQ derivatives.</p>
Full article ">Figure 4
<p>(<b>A</b>) DMSO control cells of lung cancer (A549); (<b>B</b>) effect of 5-FU at 6 µM; (<b>C</b>) effect of Everolimus at 6 µM; (<b>D</b>) effect of compound <b>10d</b> at 6 µM; (<b>E</b>) effect of compound <b>10d</b> at 3 µM; (<b>F</b>) effect of compound 10d at 6µM; (<b>G</b>) effect of compound <b>10d</b> at 3 µM.</p>
Full article ">Figure 5
<p>Molecular docking view of compound <b>10e</b> in the 4JT6 binding site.</p>
Full article ">Figure 6
<p>Plots of MD simulations of the complex of compound <b>10e</b> with mTOR protein for 100 ns. (<b>A</b>) RMSD; (<b>B</b>) RMSF; (<b>C</b>) Rg; (<b>D</b>) SASA plot; (<b>E</b>) number of hydrogen bonds.</p>
Full article ">
22 pages, 57415 KiB  
Article
Enhanced Nanogel Formulation Combining the Natural Photosensitizer Curcumin and Pectis brevipedunculata (Asteraceae) Essential Oil for Synergistic Daylight Photodynamic Therapy in Leishmaniasis Treatment
by Lara Maria Oliveira Campos, Estela Mesquita Marques, Daniele Stéfanie Sara Lopes Lera-Nonose, Maria Julia Schiavon Gonçalves, Maria Valdrinez Campana Lonardoni, Glécilla Colombelli de Souza Nunes, Gustavo Braga and Renato Sonchini Gonçalves
Pharmaceutics 2025, 17(3), 286; https://doi.org/10.3390/pharmaceutics17030286 - 21 Feb 2025
Viewed by 126
Abstract
Background/Objectives: Neglected tropical diseases (NTDs), such as leishmaniasis, remain a global health challenge due to limited therapeutic options and rising drug resistance. In this study, we developed an advanced nanogel formulation incorporating curcumin (CUR) and Pectis brevipedunculata essential oil (EOPb) [...] Read more.
Background/Objectives: Neglected tropical diseases (NTDs), such as leishmaniasis, remain a global health challenge due to limited therapeutic options and rising drug resistance. In this study, we developed an advanced nanogel formulation incorporating curcumin (CUR) and Pectis brevipedunculata essential oil (EOPb) within an F127/Carbopol 974P matrix to enhance bioavailability and therapeutic efficacy against Leishmania (Leishmania) amazonensis (LLa) promastigotes. Methods: The chemical profile of EOPb was determined through GC-MS and NMR analyses, confirming the presence of key bioactive monoterpenes such as neral, geranial, α-pinene, and limonene. The nanogel formulation (nGPC) was optimized to ensure thermosensitivity, and stability, exhibiting a sol–gel transition at physiological temperatures. Rheological analysis revealed that nGPC exhibited Newtonian behavior at 5 °C, transitioning to shear-thinning and thixotropic characteristics at 25 and 32 °C, respectively. This behavior facilitates its application and controlled drug release, making it ideal for topical formulations. Dynamic light scattering (DLS) analysis demonstrated that nGPC maintained a stable nanoscale structure with hydrodynamic radius below 300 nm, while Fourier-transform infrared spectroscopy (FTIR) confirmed strong molecular interactions between EOPb, CUR, and the polymer matrix. Biological assays demonstrated that nGPC significantly enhanced anti-promastigote activity compared to free CUR and OEPb. Results: At the highest tested concentration (50 μg/mL EOPb and 17.5 μg/mL CUR) nGPC induced over 88% mortality in LLa promastigotes across 24, 48, and 72 h, indicating sustained efficacy. Even at lower concentrations, nGPC retained dose-dependent activity, suggesting a synergistic effect between CUR and EOPb. These findings highlight the potential of nGPC as an innovative nanocarrier for daylight photodynamic therapy (dPDT) in the treatment of leishmaniasis. Future studies will investigate the underlying mechanisms of this synergism and explore the potential application of photodynamic therapy (PDT) to further enhance therapeutic outcomes. Full article
(This article belongs to the Special Issue Natural Products in Photodynamic Therapy)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Stages of preparation of nG<span class="html-italic">P</span>C (nanogel loaded with EO<span class="html-italic">Pb</span> and CUR), involving the addition of curcumin (CUR) to the solution under controlled temperature conditions to ensure its uniform dispersion within the polymeric matrix, followed by the incorporation of <span class="html-italic">Pectis brevipedunculata</span> essential oil (EO<span class="html-italic">Pb</span>). (<b>B</b>) Molecular structure of the polymers utilized in nanogel formulation and primary chemical constituents composing EO<span class="html-italic">Pb</span>. (<b>C</b>) Photographs of the nG<span class="html-italic">P</span>C1–nG<span class="html-italic">P</span>C5 formulations following the accelerated stability test, where the nG<span class="html-italic">P</span>C5 nanogel exhibits a reaming effect.</p>
Full article ">Figure 2
<p>FTIR spectra of the nanogels: (<b>A</b>) nG (empty nanogel), (<b>B</b>) nGP (EO<span class="html-italic">Pb</span>-loaded nanogel) (<b>C</b>) nG<span class="html-italic">P</span>C (EO<span class="html-italic">Pb</span>/CUR-loaded nanogel) in the range of 4000–3100 cm<sup>−1</sup>, and (<b>D</b>) nG<span class="html-italic">P</span>C in the range of 3100–690 cm<sup>−1</sup>.</p>
Full article ">Figure 3
<p>SEM micrographs of nanogels after the freeze-drying process: (<b>A</b>) and (<b>B</b>) nG (empty nanogel) at magnifications of 1000× and 5000×, respectively; (<b>C</b>), (<b>D</b>), (<b>E</b>), and (<b>F</b>) nG<span class="html-italic">P</span>C (EO<span class="html-italic">Pb</span>-loaded nanogel) at magnifications of 200×, 500×, 1000×, and 2000×, respectively.</p>
Full article ">Figure 4
<p>Comparative analysis of the diffusion coefficient D<sub><span class="html-italic">if</span></sub> as a function of temperature (K): (<b>A</b>,<b>C</b>,<b>E</b>) nG (empty nanogel) and (<b>B</b>,<b>D</b>,<b>F</b>) nG<span class="html-italic">P</span>C (EO<span class="html-italic">Pb</span>/CUR-loaded nanogel).</p>
Full article ">Figure 5
<p>Plot of <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msubsup> <mi>G</mi> <mi>d</mi> <mo>‡</mo> </msubsup> </mrow> </semantics></math> as a function of temperature for the (<b>A</b>) nG (empty nanogel) and (<b>B</b>) nG<span class="html-italic">P</span>C (EO<span class="html-italic">Pb</span>/CUR-loaded nanogel).</p>
Full article ">Figure 6
<p>Rheological behavior of the nG<span class="html-italic">P</span>C (EO<span class="html-italic">Pb</span>/CUR-loaded nanogel) sample at varying temperatures: (<b>A</b>) Newtonian fluid behavior at 5 °C with a linear shear stress versus shear rate curve; (<b>B</b>) shear-thinning behavior at 25 °C, demonstrating viscosity decrease with increasing shear rate; (<b>C</b>) pronounced thixotropic pattern at 32 °C, with viscosity reduction and incomplete structural recovery after shear. The curves demonstrate the nanogel’s dynamic response to shear forces, highlighting its suitability for topical applications; (<b>D</b>) real image of the nG<span class="html-italic">P</span>C nanogel during the sol–gel transition, showing sol at 5 °C and gel at 25 °C. The curves demonstrate the nanogel’s dynamic response to shear forces, highlighting its suitability for topical applications.</p>
Full article ">Figure 7
<p>(<b>A</b>) in vitro viability (%) of <span class="html-italic">LLa</span> promastigote cells in response to different concentrations of nG<span class="html-italic">P</span>C (EO<span class="html-italic">Pb</span>/CUR-loaded nanogel). The concentrations C4–C1 represent the combination of EO<span class="html-italic">Pb</span>/CUR as follows: C4 = 6.65/2.19, C3 = 12.5/4.38, C2 = 25/8.75 and C1 = 50/17.5 (μg/mL). The nG (empty nanogel) and AmB (amphotericin B) were used as negative and positive controls, respectively. nG<span class="html-italic">P</span> (EO<span class="html-italic">Pb</span>-loaded nanogel) and nG<span class="html-italic">C</span> (CUR-loaded nanogel) did not exhibit cytotoxicity against <span class="html-italic">LLa</span> at any of the tested concentrations; therefore, their results were not included in the graph. (<b>B</b>) Representative images of <span class="html-italic">LLa</span> cultures treated with nanogel formulations at different time points: (a) untreated; (b) nG; (c) nGC (50 μg/mL); (d) nG<span class="html-italic">P</span> (17.5 μg/mL); (e) AmB (1.95 μg/mL); (f–i) nG<span class="html-italic">P</span>C (C4–C1).</p>
Full article ">
6 pages, 955 KiB  
Short Note
((1R,3aS,5aR,5bR,7aR,9S,11aR,11bR,13aR,13bR)-9-Acetoxy-5a,5b,8,8,11a-pentamethyl-1-(prop-1-en-2-yl)icosahydro-3aH-cyclopenta[a]chrysen-3a-yl)methyl 2-Bromo-3-methylbenzoate
by Yuzhu Guo, Yang Xiao, Carmine Coluccini and Paolo Coghi
Molbank 2025, 2025(1), M1971; https://doi.org/10.3390/M1971 - 20 Feb 2025
Viewed by 175
Abstract
In this report, we discuss the synthesis of modified betulin through the chemical derivatization of a natural compound. The compound was fully characterized by proton (1H), carbon-13 (13C), heteronuclear single quantum coherence (HSQC) and distortionless enhancement through polarization transfer [...] Read more.
In this report, we discuss the synthesis of modified betulin through the chemical derivatization of a natural compound. The compound was fully characterized by proton (1H), carbon-13 (13C), heteronuclear single quantum coherence (HSQC) and distortionless enhancement through polarization transfer (DEPT) NMR and elemental analysis. We investigated the optical properties through ultraviolet (UV) and Fourier-transform infrared (FTIR) spectroscopy. Full article
(This article belongs to the Section Natural Product Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Example of betulin derivatives reported in the literature: (<b>a</b>–<b>c</b>) [<a href="#B2-molbank-2025-M1971" class="html-bibr">2</a>], (<b>d</b>,<b>e</b>) [<a href="#B12-molbank-2025-M1971" class="html-bibr">12</a>], (<b>f</b>) [<a href="#B13-molbank-2025-M1971" class="html-bibr">13</a>].</p>
Full article ">Scheme 1
<p>(<b>a</b>) Synthesis of ((1<span class="html-italic">R</span>,3a<span class="html-italic">S</span>,5a<span class="html-italic">R</span>,5b<span class="html-italic">R</span>,7a<span class="html-italic">R</span>,9<span class="html-italic">S</span>,11a<span class="html-italic">R</span>,11b<span class="html-italic">R</span>,13a<span class="html-italic">R</span>,13b<span class="html-italic">R</span>)-9-hydroxy-5a,5b,8,8,11a-pentamethyl-1-(prop-1-en-2-yl)icosahydro-3a<span class="html-italic">H</span>-cyclopenta[a]chrysen-3a-yl)methyl 2-bromo-3-methylbenzoate: (i) EDC (1.08 eq.), DMAP (0.48 eq.), DCM, and TEA (0.48 eq.) for 12 h at room temperature. (<b>b</b>) Synthesis of ((1<span class="html-italic">R</span>,3a<span class="html-italic">S</span>,5a<span class="html-italic">R</span>,5b<span class="html-italic">R</span>,7a<span class="html-italic">R</span>,9<span class="html-italic">S</span>,11a<span class="html-italic">R</span>,11b<span class="html-italic">R</span>,13a<span class="html-italic">R</span>,13b<span class="html-italic">R</span>)-9-acetoxy-5a,5b,8,8,11a-pentamethyl-1-(prop-1-en-2-yl)icosahydro-3a<span class="html-italic">H</span>-cyclopenta[a]chrysen-3a-yl)methyl 2-bromo-3-methylbenzoate: (ii) DMAP, dry DCM, and TEA (3 eq.) with ice at 0 °C for 12 h at room temperature.</p>
Full article ">
17 pages, 4338 KiB  
Article
Self-Thickening Materials Derived from Phenylpropanoid Ene Reactions
by Atanu Biswas, Huai N. Cheng, Bret Chisholm, Ryan Beni, Zengshe Liu, Karl Vermillion, Michael Appell, Kelton Forson, Omar El Seoud, Carlucio R. Alves and Roselayne F. Furtado
Molecules 2025, 30(5), 977; https://doi.org/10.3390/molecules30050977 - 20 Feb 2025
Viewed by 241
Abstract
In this work, we report the observation of uncatalyzed ene reactions between several phenylpropanoid compounds and diethyl azodicarboxylate (DEAD). For allylbenzene, the reaction produces the ene product at molar ratios of up to 1:2 of allylbenzene to DEAD. At higher levels of DEAD, [...] Read more.
In this work, we report the observation of uncatalyzed ene reactions between several phenylpropanoid compounds and diethyl azodicarboxylate (DEAD). For allylbenzene, the reaction produces the ene product at molar ratios of up to 1:2 of allylbenzene to DEAD. At higher levels of DEAD, more complex reactions are observed. For the reaction between methyl eugenol and DEAD, similar ene reaction products have been found. However, the reaction of eugenol with DEAD is more complex; in addition to the ene reaction, other reactions happen at the same time. Most of the structures of the resulting products have been elucidated using NMR spectroscopy (1H, 13C, and 2D), and the findings have been further corroborated by FTIR analysis. Interestingly, these products appear to undergo molecular aggregation, which results in self-thickening in their neat form. However, the viscosity significantly decreases upon dilution with a solvent. This self-thickening property suggests their potential use as thickening agents in organic solvent formulations. Full article
(This article belongs to the Special Issue π-Conjugated Functional Molecules & Polymers)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photographs of (<b>a</b>) samples A-2, M-2, and E-2 (at ca. 1:2 molar ratio of alkene/DEAD); (<b>b</b>) the same three samples after overnight heating at 90 °C.</p>
Full article ">Figure 2
<p>NMR spectra of the 1:2 reaction products between allylbenzene and DEAD (sample A-2): (<b>a</b>) <sup>13</sup>C spectrum and (<b>b</b>) <sup>1</sup>H spectrum. U and A denote unreacted and ene products, respectively; the subscripts correspond to the numbering shown in <a href="#molecules-30-00977-sch003" class="html-scheme">Scheme 3</a>. The letter X indicates the CDCl<sub>3</sub> peaks, and the letter s denotes ethyl acetate peaks.</p>
Full article ">Figure 3
<p>NMR spectra of the 1:3 reaction products between AB and DEAD (sample A-5): (<b>a</b>) <sup>13</sup>C spectrum, (<b>b</b>) <sup>1</sup>H spectrum. A and C denote ene products and ethoxyformaldehyde, respectively; the subscripts correspond to the numbering shown in <a href="#molecules-30-00977-sch003" class="html-scheme">Scheme 3</a>. R denotes the peaks from residual DEAD moieties after the reaction and X the CDCl<sub>3</sub> peaks.</p>
Full article ">Figure 4
<p>NMR spectra of the reaction products between methyl eugenol and DEAD (sample M-1): (<b>a</b>) <sup>13</sup>C spectrum and (<b>b</b>) <sup>1</sup>H spectrum. V and G denote unreacted methyl eugenol and ene products; the subscripts correspond to the numbering shown in <a href="#molecules-30-00977-sch004" class="html-scheme">Scheme 4</a>. X denotes the CDCl<sub>3</sub> peak, and s denotes ethyl acetate peaks.</p>
Full article ">Figure 5
<p>NMR spectra of the reaction products between eugenol and DEAD (sample E-1). (<b>a</b>) <sup>13</sup>C spectrum, (<b>b</b>) <sup>1</sup>H spectrum. I and H denote unreacted eugenol and ene products, respectively; the subscripts correspond to the numbering shown in <a href="#molecules-30-00977-sch005" class="html-scheme">Scheme 5</a>. R denotes the peaks from residual DEAD moieties after the reaction, and X represents the CDCl<sub>3</sub> peaks.</p>
Full article ">Figure 6
<p>FTIR spectra of starting alkenes and ene reaction products (at 1:2 molar ratio of alkene/DEAD). From bottom to top: AB and AB-DEAD (sample A-2), ME and ME-DEAD (sample M-2), and eugenol and eugenol/DEAD (sample E-2).</p>
Full article ">Figure 7
<p>Viscosity buildup (in Pa-s) during the reaction of AB and DEAD as a function of reaction time (in seconds) at 87 °C for 19.4 h.</p>
Full article ">Figure 8
<p>Viscosity of the AB-DEAD reaction products at different wt % DEAD measured as a function of increasing temperature; the samples are (from bottom to top) A-R1, A-R2, A-R3, and A-R4.</p>
Full article ">Figure 9
<p>Viscosity of the ME-DEAD reaction products at different wt % DEAD measured as a function of temperature; the samples are (from bottom to top) M-R1, M-R2, M-R3, and M-R4.</p>
Full article ">Figure 10
<p>Viscosity of the eugenol/DEAD reaction products at different wt % DEAD measured as a function of temperature; the samples are (from bottom to top) E-R1, E-R2, and E-R3.</p>
Full article ">Scheme 1
<p>Natural products with allylbenzene functionality: eugenol (I), estragole (II), anethole (III), safrole (IV), and methyl eugenol (V).</p>
Full article ">Scheme 2
<p>Reaction using heat between allylbenzene and DEAD.</p>
Full article ">Scheme 3
<p>Structures for unreacted allylbenzene (U) and the AB-DEAD ene reaction derivative (A) and three other structures (C,D,E). Structures (U) and (A) are numbered to facilitate NMR assignments.</p>
Full article ">Scheme 4
<p>Structures for unreacted methyl eugenol (V) and the methyl eugenol/DEAD ene reaction derivative (G), numbered to facilitate NMR assignments.</p>
Full article ">Scheme 5
<p>Structures for the eugenol/DEAD ene derivative (H), numbered to facilitate NMR assignments, and two other possible structures produced in the eugenol/DEAD reaction.</p>
Full article ">
31 pages, 5710 KiB  
Article
Antitumor Activity, Mechanisms of Action and Phytochemical Profiling of Sub-Fractions Obtained from Ulex gallii Planch. (Fabaceae): A Medicinal Plant from Galicia (Spain)
by Lucía Bada, Hussain Shakeel Butt, Elías Quezada, Aitor Picos, Helle Wangensteen, Kari Tvete Inngjerdingen, José Gil-Longo and Dolores Viña
Molecules 2025, 30(4), 972; https://doi.org/10.3390/molecules30040972 - 19 Feb 2025
Viewed by 233
Abstract
The plant kingdom serves as a valuable resource for cancer drug development. This study explored the antitumor activity of different sub-fractions (hexane, dichloromethane and methanol) of U. gallii (gorse) methanol extract in glioblastoma (U-87MG and U-373MG) and neuroblastoma (SH-SY5Y) cell lines, along with [...] Read more.
The plant kingdom serves as a valuable resource for cancer drug development. This study explored the antitumor activity of different sub-fractions (hexane, dichloromethane and methanol) of U. gallii (gorse) methanol extract in glioblastoma (U-87MG and U-373MG) and neuroblastoma (SH-SY5Y) cell lines, along with their phytochemical profiles. Cytotoxicity was evaluated using 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) and lactate dehydrogenase (LDH) assays, and cell cycle arrest and apoptosis were assessed through flow cytometry and by measuring reactive oxygen species (ROS) and protein expression levels. D7 and D8 dichloromethane sub-fractions significantly reduced cell viability, triggered early apoptosis in SH-SY5Y and U-87MG cells and specifically increased ROS levels in U-87MG cells. Western blot analyses showed that D7 increased p53, caspase-3, caspase-8 and γH2AX expression in SH-SY5Y and U-87MG cells, while D8 specifically elevated p53 in SH-SY5Y cells and caspase-3 in both cell lines. In U-373 cells, D7 and D8 markedly reduced cell viability, with D8 inducing necrosis. Morphological changes indicative of apoptosis were also observed in all cell lines. Bioinformatic analysis of UHPLC-MS and GC-MS data tentatively identified 20 metabolites in D7 and 15 in D8, primarily flavonoids. HPLC-DAD confirmed isoprunetin and genistein as the most abundant in D7 and D8, respectively, both isolated and identified by NMR spectroscopy. Most of the flavonoids identified have been reported as antitumor agents, suggesting that these compounds may be responsible for the observed pharmacological activity. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of hexane (H1, H3.1–3.6, H4), dichloromethane (D1–D8) and methanol (M1–M4) sub-fractions (0.1 mg/mL) on cell viability after 24 h of treatment of the: (<b>A</b>) SH-SY5Y, (<b>B</b>) U-87MG and (<b>C</b>) U-373MG cell lines, and effect of dichloromethane (D3–D8) sub-fractions (0.1 mg/mL) on cell viability after 24 h of treatment of the (<b>D</b>) MRC-5 cell line. Topotecan (10 µM) was used as a positive control. Data represent the means ± s.e.m. (standard error of mean) from at least three independent experiments (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.0001 compared to the respective control (cells treated with DMSO (vehicle) &lt; 1%).</p>
Full article ">Figure 2
<p>Effect of dichloromethane sub-fractions D7 and D8 (0.05 mg/mL) after 24 h of treatment on the cell viability of cell lines: (<b>A</b>) SH-SY5Y, (<b>B</b>) U-87MG and (<b>C</b>) U-373MG. Data represent the means ± s.e.m. (standard error of mean) from at least three independent experiments (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> &lt; 0.0001 compared to the respective control (cells treated with DMSO (vehicle) &lt; 0.5%).</p>
Full article ">Figure 3
<p>Effect of dichloromethane sub-fractions (D3–D8, 0.05 mg/mL) on LDH release after 24 h of treatment of cell lines: (<b>A</b>) SH-SY5Y, (<b>B</b>) U-87MG and (<b>C</b>) U-373MG. Data represent the means ± s.e.m. (standard error of mean) from at least three independent experiments (<span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> &lt; 0.01 compared to the respective control (cells treated with DMSO (vehicle) &lt; 0.5%).</p>
Full article ">Figure 4
<p>Effect of the dichloromethane sub-fractions (D3–D8, 0.05 mg/mL) on ROS production after 24 h of treatment of cell lines: (<b>A</b>) SH-SY5Y, (<b>B</b>) U-87MG and (<b>C</b>) U-373MG. H<sub>2</sub>O<sub>2</sub> (100 µM) was used as a positive control. Data represent the means ± s.e.m. (standard error of mean) from at least three independent experiments (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared to the respective control (cells treated with DMSO (vehicle) &lt; 0.5%).</p>
Full article ">Figure 5
<p>Effect of the sub-fractions D3–D8 (0.05 mg/mL) on cell cycle progression after 24 h of treatment of cell lines: (<b>A</b>) SH-SY5Y, (<b>B</b>) U-87MG and (<b>C</b>) U-373MG. Data represent the means ± s.e.m. (standard error of mean) from at least three independent experiments (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05,** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.0001 compared to the respective control (cells treated with DMSO (vehicle) &lt; 0.5%).</p>
Full article ">Figure 6
<p>Effect of the sub-fractions D3–D8 (0.05 mg/mL) on cell death profile after 24 h of treatment of cell lines: (<b>A</b>) SH-SY5Y, (<b>B</b>) U-87MG and (<b>C</b>) U-373MG. Triton X-100, TX (1%) was used as a positive control. Data represent the means ± s.e.m. (standard error of mean) from at least three independent experiments (<span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.0001 compared to the respective control (cells treated with DMSO (vehicle) &lt; 0.5%).</p>
Full article ">Figure 7
<p>Effects of the sub-fraction D7 (0.05 mg/mL) on the expression of (<b>A</b>) p53 and (<b>B</b>) caspase-3 in SH-SY5Y, U-87MG and U-373MG cells lines and (<b>C</b>) caspase-8 and (<b>D</b>) ƴ-H2AX in SH-SY5Y and U-87MG cell lines after 24 h of treatment. DMSO: Cells treated with vehicle (&lt;0.5%). Data represent the means ± s.e.m. (standard error of mean) from at least two independent experiments (<span class="html-italic">n</span> = 2). * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.0001 compared to the control.</p>
Full article ">Figure 8
<p>Effects of the sub-fraction D8 (0.05 mg/mL) on the expression of (<b>A</b>) p53 and (<b>B</b>) caspase-3 in SH-SY5Y, U-87MG and U-373MG cells lines and (<b>C</b>) caspase-8 in SH-SY5Y and U-87MG cell lines after 24 h of treatment. DMSO: Cells treated with vehicle (&lt;0.5%). Data represent the means ± s.e.m. (standard error of mean) from at least two independent experiments (<span class="html-italic">n</span> = 2). * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.0001 compared to the control.</p>
Full article ">Figure 9
<p>Effect of the sub-fractions D7 and D8 (0.05 mg/mL) on cell morphology after 24 h of treatment in SH-SY5Y cells using Hoescht (blue, chromatin status) and Phalloidin-iFluorTM 594 (red, cytoplasmatic traits). DMSO: Cells treated with vehicle (&lt;0.5%). 20× magnification. Yellow arrows: apoptotic bodies; white arrow: alteration in nuclear morphology.</p>
Full article ">Figure 10
<p>Effect of the sub-fractions D7 and D8 (0.05 mg/mL) on cell morphology after 24 h of treatment in U-87MG cells using Hoescht (blue, chromatin status) and Phalloidin-iFluor<sup>TM</sup> 594 (red, cytoplasmatic traits). DMSO: Cells treated with vehicle (&lt;0.5%). 20× magnification. Yellow arrows: apoptotic bodies; white arrow: alteration in nuclear morphology.</p>
Full article ">Figure 11
<p>Effect of the sub-fractions D7 and D8 (0.05 mg/mL) on cell morphology after 24 h of treatment in U-373MG cells using Hoescht (blue, chromatin status) and Phalloidin-iFluorTM 594 (red, cytoplasmatic traits). DMSO: Cells treated with vehicle (&lt;0.5%). 20× magnification. White arrow: alteration in nuclear morphology.</p>
Full article ">Figure 12
<p>HPLC-DAD chromatograms at 255 nm of sub-fractions D7 (<b>A</b>) and D8 (<b>B</b>).</p>
Full article ">
7 pages, 1600 KiB  
Short Note
2,4,6-Trichloro-cyclohexa-2,5-dienone
by Guido Gambacorta, Qin Han Teo and Ian R. Baxendale
Molbank 2025, 2025(1), M1969; https://doi.org/10.3390/M1969 - 19 Feb 2025
Viewed by 182
Abstract
A continuous flow process was optimised for the perchlorination of p-cresol to the corresponding 2,4,6-trichloro-cyclohexa-2,5-dienone derivative employing trichloroisocyanuric acid as a green and safer-to-handle chlorinating agent. The system could furnish 200 g of pure material within 5 h of operation (throughput = [...] Read more.
A continuous flow process was optimised for the perchlorination of p-cresol to the corresponding 2,4,6-trichloro-cyclohexa-2,5-dienone derivative employing trichloroisocyanuric acid as a green and safer-to-handle chlorinating agent. The system could furnish 200 g of pure material within 5 h of operation (throughput = 40 g h−1). The compound was easily isolated by filtration and obtained in 95% purity as determined by GC analysis; it could be further purified by crystallisation from a 20:1 Hexane/AcOEt mixture left at −20 °C overnight. The resultant product was characterised by 1H & 13C NMR, MS, IR analyses, with melting point and X-ray single-crystal data being obtained, confirming the structure. Full article
(This article belongs to the Section Organic Synthesis and Biosynthesis)
Show Figures

Figure 1

Figure 1
<p>Examples of biological active 2,4,6-trichloro-cyclohexa-2,5-dienone derivatives.</p>
Full article ">Figure 2
<p>A photo of the setup employed for the perchlorination of <span class="html-italic">p</span>-cresol to <b>2</b>.</p>
Full article ">Scheme 1
<p>Setup employed for the preparation of <b>2</b> under flow conditions.</p>
Full article ">
13 pages, 1535 KiB  
Article
Metabolomic Nuclear Magnetic Resonance Insights into Wine and Grape Ale Maturation
by Dessislava Gerginova, Plamen Chorbadzhiev and Svetlana Simova
Beverages 2025, 11(1), 29; https://doi.org/10.3390/beverages11010029 - 18 Feb 2025
Viewed by 240
Abstract
The chemical profiles of young and mature wines produced from three grape varieties Merlot, Mavrud, and Sauvignon blanc were analyzed using 1H nuclear magnetic resonance (NMR) spectroscopy and advanced statistical methods. Furthermore, grape ales―a hybrid of beer and wine—were subjected to analysis [...] Read more.
The chemical profiles of young and mature wines produced from three grape varieties Merlot, Mavrud, and Sauvignon blanc were analyzed using 1H nuclear magnetic resonance (NMR) spectroscopy and advanced statistical methods. Furthermore, grape ales―a hybrid of beer and wine—were subjected to analysis to facilitate a comparison of their composition with that of traditional wines. The analysis yielded a total of 37 compounds, which were identified and quantified. Orthogonal partial least squares discriminant analysis (OPLS-DA) models were employed to distinguish the chemical profiles of young and mature wines, as well as those of grape ales. The findings demonstrate that the fermentation and aging processes result in the formation of distinctive chemical signatures in wines, with key compounds such as shikimic acid and fructose contributing to this differentiation. The identified compounds comprise seven alcohols (2,3-butanediol, glycerol, 2-methylpropan-1-ol, 3-methyl-butan-1-ol, myo-inositol, 1-propanol, 2-phenylethanol), six organic acids (galacturonic, citric, lactic, malic, shikimic, succinic), three amino acids (alanine, proline, tyrosine), four sugars (arabinose, fructose, galactose, glucose), coutaric acid, and acetoin. The levels of these 22 components enabled the successful differentiation of young and mature wines among the three grape varieties. These findings underscore the substantial chemical distinctions between grape ales and wines, thereby emphasizing the potential of grape ales as an innovative fermented beverage. Full article
(This article belongs to the Section Quality, Nutrition, and Chemistry of Beverages)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><sup>1</sup>H NMR spectra of mature wine (red, <b>sb_m1</b>), young wine (olive, <b>sb_y2</b>), and grape ale (orange, <b>ga_3</b>).</p>
Full article ">Figure 2
<p>Contribution plot of wines (yellow) and grape ale (orange).</p>
Full article ">Figure 3
<p>OPLS-DA score plot for the classification of young wine (olive), mature wine (red), and grape ale (orange).</p>
Full article ">Figure 4
<p><span class="html-italic">Nightingale</span> diagram comparing the average content of compounds in young (olive, Y) and mature (red, M) wines.</p>
Full article ">Figure 5
<p>OPLS-DA score plot illustrating the differentiation of young and aged Merlot (light blue and blue), Mavrud (pink and red violet), and Sauvignon blanc (yellow and orange) wines.</p>
Full article ">Figure 6
<p>Concentration changes in Merlot (<b>A</b>), Mavrud (<b>B</b>), and Sauvignon blanc (<b>C</b>) during the aging process from young to mature wine, represented as major (<b>left</b>, ≥480 mg/L) and minor (<b>right</b>, &lt;480 mg/L) groups. Positive changes are indicated in green, while negative changes are in red.</p>
Full article ">Figure 7
<p>PCA biplot illustrating the differentiation of four grape ales based on their characteristic components.</p>
Full article ">
9 pages, 4021 KiB  
Brief Report
Novel Metabolites as Potential Indicators of Recovery After Large Vessel Occlusion Stroke: A Pilot Study
by Evgeny V. Sidorov, Kyle Smith, Chao Xu and Dharambir K. Sanghera
Neurol. Int. 2025, 17(2), 30; https://doi.org/10.3390/neurolint17020030 - 18 Feb 2025
Viewed by 241
Abstract
Introduction: Serum metabolome changes after acute ischemic stroke (AIS), but the significance of this is poorly understood. We evaluated whether this change is associated with AIS outcomes in patients with large vessel occlusion (LVO). To improve validity, we combined cross-sectional and longitudinal designs [...] Read more.
Introduction: Serum metabolome changes after acute ischemic stroke (AIS), but the significance of this is poorly understood. We evaluated whether this change is associated with AIS outcomes in patients with large vessel occlusion (LVO). To improve validity, we combined cross-sectional and longitudinal designs and analyzed serum using Nuclear Magnetic Resonance (NMR) and Liquid Chromatography–Mass Spectrometry (LC-MS). Methodology: In the cross-sectional part, we compared serum metabolome from 48 LVO strokes, collected at 48–72 h, and analyzed with NMR, while in the longitudinal part, we compared metabolome from 15 LVO strokes, collected at <24 h, 48–72 h, 5–7 days, and 80–120 days, and analyzed with LC-MS between patients with modified Rankin Scores (mRS) of 0–3 and 4–6 at 90 days. We hypothesized that compounds elevated in patients with mRS 0–3 in the cross-sectional part would also be elevated in the longitudinal part, and vice versa. We used regression for the analysis and TSBH for multiple testing. Results: In the cross-sectional part, cholesterol, choline, phosphoglycerides, sphingomyelins, and phosphatidylethanolamines had lower levels in patients with an mRS of 0–3 compared to an mRS of 4–6. In the longitudinal part, lower levels of sphingomyelin (d18:1/19:0, d19:1/18:0)* significantly correlated with an mRS of 0–3 in patients with small infarction volume, while lower levels of sphingolipid N-palmitoyl-sphingosine (d18:1/16:0), 1-palmitoyl-2-docosahexaenoyl-GPC (16:0/22:6), 1-palmitoyl-2-docosahexaenoyl-GPE, palmitoyl-docosahexaenoyl-glycerol (16:0/22:6), campesterol, and 3beta-hydroxy-5-cholestenoate correlated with an mRS of 0–3 in patients with large infarction volume. Conclusions: This pilot study showed that lower levels of lipidomic components nerve cell membrane correlate with good AIS outcomes. If proven on large-scale studies, these compounds may become important AIS outcome markers. Full article
(This article belongs to the Collection Biomarkers in Stroke Prognosis)
17 pages, 2851 KiB  
Article
Acetylenic Substituent: Influence on the Structure, Electrochemical, Photophysical, and Thermal Properties of Rhenium(I) and Platinum(II) Complexes
by Bartosz Zowiślok, Anna Świtlicka, Anna Maria Maroń and Sławomir Kula
Molecules 2025, 30(4), 915; https://doi.org/10.3390/molecules30040915 - 16 Feb 2025
Viewed by 271
Abstract
The ‘wire-like’ acetylenic moiety is an important building block in organic and coordination chemistry that can facilitate electron transfer in donor–acceptor compounds, contributing to the enhancement of their photophysical properties. 2,6-Bis-(thiazol-2-yl)pyridine (dtpy) functionalized with a 4-phenylacetylene group (Ph-C≡C-dtpy) was, [...] Read more.
The ‘wire-like’ acetylenic moiety is an important building block in organic and coordination chemistry that can facilitate electron transfer in donor–acceptor compounds, contributing to the enhancement of their photophysical properties. 2,6-Bis-(thiazol-2-yl)pyridine (dtpy) functionalized with a 4-phenylacetylene group (Ph-C≡C-dtpy) was, for the first time, used for the preparation of [ReCl(CO)3(Ph-C≡C-dtpy)] and [Pt(Ph-C≡C-dtpy)Cl]CF3SO3 in order to understand the properties derived from the use of the acetylenic substituent. The coordination ability of Ph-C≡C-dtpy toward Pt(II) and Re(I) centers was determined. All the studied compounds were characterized using FT-IR, 1H NMR, and 13C NMR spectroscopies, elemental analysis and, in the case of the free ligand and rhenium(I) complex, single-crystal X-ray analysis was also used. Their electrochemical, photophysical, and thermal properties were compared with the previously described similar systems. The photoluminescence spectra of Ph-C≡C-dtpy, [ReCl(CO)3(Ph-C≡C-dtpy)] and [Pt(Ph-C≡C-dtpy)Cl]CF3SO3 were investigated in solution and in the solid state at 298 K and 77 K. The experimental results were supported by the DFT and TD-DFT calculations. As a result of the introduction of the -C≡C- moiety into the organic ligand skeleton, the Re(I) and Pt(II) complexes display room-temperature emission. In the case of [Pt(Ph-C≡C-dtpy)Cl]CF3SO3, photoluminescence lifetime in a microsecond regime was observed, whereas nanosecond lifetime for [ReCl(CO)3(Ph-C≡C-dtpy)] in solution is comparable to lifetimes previously observed for rhenium(I) compounds with 4-substituted dtpys. Full article
(This article belongs to the Special Issue Advances in Coordination Chemistry 2.0)
Show Figures

Figure 1

Figure 1
<p>Synthesis of Ph-C≡C-<span class="html-italic">dtpy</span> by condensation.</p>
Full article ">Figure 2
<p>A perspective view showing the asymmetric units of Ph-C≡C-dtpy (<b>a</b>) and [ReCl(CO)<sub>3</sub>(Ph-C≡C-dtpy)] (<b>b</b>) with atom numbering. The displacement ellipsoids are drawn at 50% probability.</p>
Full article ">Figure 3
<p>A view of the packing of Ph-C≡C-<span class="html-italic">dtpy</span> showing intermolecular π–π stacking interactions (<b>a</b>); a view of supramolecular honeycomb-like layers in [ReCl(CO)<sub>3</sub>(Ph-C≡C-<span class="html-italic">dtpy</span>)] (<b>b</b>).</p>
Full article ">Figure 4
<p>UV-VIS spectra together with experimental data.</p>
Full article ">Figure 5
<p>UV-Vis spectra of [ReCl(CO)<sub>3</sub>(Ph-C≡C-<span class="html-italic">dtpy</span>)] in acetonitrile (<b>a</b>) and chloroform (<b>b</b>) recorded once every 25 min for 5 h.</p>
Full article ">Figure 6
<p>Experimental (blue line) absorption spectra in chloroform (<b>a</b>) and acetonitrile (<b>b</b>), along with transitions (black lines) computed at the TD-DFT/PBE1PBE/def2-TZVP/def2-TZVPD level, together with the compositions of selected frontier orbitals for metal complexes.</p>
Full article ">
Back to TopTop