[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline

Search Results (170)

Search Parameters:
Keywords = naphthol

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 4311 KiB  
Article
Synthesis of Tumor Selective Indole and 8-Hydroxyquinoline Skeleton Containing Di-, or Triarylmethanes with Improved Cytotoxic Activity
by Dóra Hegedűs, Nikoletta Szemerédi, Krisztina Petrinca, Róbert Berkecz, Gabriella Spengler and István Szatmári
Molecules 2024, 29(17), 4176; https://doi.org/10.3390/molecules29174176 - 3 Sep 2024
Viewed by 396
Abstract
The reaction between glycine-type aminonaphthol derivatives substituted with 2- or 1-naphthol and indole or 7-azaindole has been tested. Starting from 2-naphthol as a precursor, the reaction led to the formation of ring-closed products, while in the case of a 1-naphthol-type precursor, the desired [...] Read more.
The reaction between glycine-type aminonaphthol derivatives substituted with 2- or 1-naphthol and indole or 7-azaindole has been tested. Starting from 2-naphthol as a precursor, the reaction led to the formation of ring-closed products, while in the case of a 1-naphthol-type precursor, the desired biaryl ester was isolated. The synthesis of a bifunctional precursor starting from 5-chloro-8-hydroxyquinoline, morpholine, and ethyl glyoxylate via modified Mannich reaction is reported. The formed Mannich base 10 was subjected to give bioconjugates with indole and 7-azaindole. The effect of the aldehyde component and the amine part of the Mannich base on the synthetic pathway was also investigated. In favor of having a preliminary overview of the structure-activity relationships, the derivatives have been tested on cancer and normal cell lines. In the case of bioconjugate 16, as the most powerful scaffold in the series bearing indole and a 5-chloro-8-hydroxyquinoline skeleton, a potent toxic activity against the resistant Colo320 colon adenocarcinoma cell line was observed. Furthermore, this derivative was selective towards cancer cell lines showing no toxicity on non-tumor fibroblast cells. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Systematic investigation of compounds <b>15</b> and <b>18</b> to determine the effect of the leaving groups.</p>
Full article ">Scheme 1
<p>Reactions of 2-(2-hydroxynaphthalen-1-yl)-2-morpholinoacetate with indole and 7-azaindole.</p>
Full article ">Scheme 2
<p>Preparation of target compound <b>7</b> from 1-naphthol-substituted precursor <b>6</b>.</p>
Full article ">Scheme 3
<p>Synthesis of ethyl 2-(5-chloro-8-hydroxyquinolin-7-yl)-2-morpholinoacetate (<b>10</b>).</p>
Full article ">Scheme 4
<p>The transformation of compound <b>10</b> bearing the 5-chloro-8-hydroxyquinoline moiety with indole and 7-azaindole.</p>
Full article ">Scheme 5
<p>The reaction of precursor <b>13</b> with indole and 7-azaindole.</p>
Full article ">Scheme 6
<p>The transformation of <b>15</b> containing paraformaldehyde as the aldehyde component.</p>
Full article ">Scheme 7
<p>Synthesis of compound <b>16</b> starting from different precursors.</p>
Full article ">
17 pages, 2616 KiB  
Article
Antifungal Mechanism of Ruta graveolens Essential Oil: A Colombian Traditional Alternative against Anthracnose Caused by Colletotrichum gloeosporioides
by Yeimmy Peralta-Ruiz, Junior Bernardo Molina Hernandez, Carlos David Grande-Tovar, Annalisa Serio, Luca Valbonetti and Clemencia Chaves-López
Molecules 2024, 29(15), 3516; https://doi.org/10.3390/molecules29153516 - 26 Jul 2024
Viewed by 553
Abstract
Here, we report for the first time on the mechanisms of action of the essential oil of Ruta graveolens (REO) against the plant pathogen Colletotrichum gloeosporioides. In particular, the presence of REO drastically affected the morphology of hyphae by inducing changes in [...] Read more.
Here, we report for the first time on the mechanisms of action of the essential oil of Ruta graveolens (REO) against the plant pathogen Colletotrichum gloeosporioides. In particular, the presence of REO drastically affected the morphology of hyphae by inducing changes in the cytoplasmic membrane, such as depolarization and changes in the fatty acid profile where straight-chain fatty acids (SCFAs) increased by up to 92.1%. In addition, REO induced changes in fungal metabolism and triggered apoptosis-like responses to cell death, such as DNA fragmentation and the accumulation of reactive oxygen species (ROS). The production of essential enzymes involved in fungal metabolism, such as acid phosphatase, β-galactosidase, β-glucosidase, and N-acetyl-β-glucosaminidase, was significantly reduced in the presence of REO. In addition, C. gloeosporioides activated naphthol-As-BI phosphohydrolase as a mechanism of response to REO stress. The data obtained here have shown that the essential oil of Ruta graveolens has a strong antifungal effect on C. gloeosporioides. Therefore, it has the potential to be used as a surface disinfectant and as a viable replacement for fungicides commonly used to treat anthracnose in the postharvest testing phase. Full article
(This article belongs to the Special Issue Plant Bioactive Compounds in Pharmaceuticals)
Show Figures

Figure 1

Figure 1
<p>Effect of REO on membrane potential of <span class="html-italic">C. gloeosporioides</span>: (<b>a</b>,<b>b</b>) control sample, (<b>c</b>) sample treated with 8.2 µg/mL REO after one hour. Scale bars represent 10 µm.</p>
Full article ">Figure 2
<p>Effect of REO on the cell constituents release over time: (<b>A</b>) concentration of material released in <span class="html-italic">C. gloeosporioides</span> treated with REO, (<b>B</b>) concentration of released proteins, (<b>C</b>) extracellular pH. Values are the averages of the replicates for all the analyses. Error bars are ±SD of the means. Different lowercase letters in the same treatment mean significant differences among the times (<span class="html-italic">p</span> &lt; 0.05). Different uppercase letters at the same time mean significant differences between treatments (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p><span class="html-italic">C. gloeosporioides</span> (<b>a</b>) control, (<b>b</b>) treated with REO. The germinated conidium after one hour of REO treatment was fixed and stained with Hoechst 33258. The nuclei are fluorescently stained in blue. Scale bars, 10 mm.</p>
Full article ">Figure 4
<p>Effect of REO in ROS generation in <span class="html-italic">Colletotrichum gloeosporioides</span>: (<b>a</b>) control sample, (<b>b</b>) sample exposed to REO. Scale bar represents 10 mm.</p>
Full article ">Figure 5
<p>Enzymatic production by <span class="html-italic">C. gloeosporioides</span> in Czapek broth treated with REO (8.2 µg/mL). Different lowercase letters in the column mean a significant difference during time of incubation (<span class="html-italic">p</span> &lt; 0.05), and different uppercase letters in the column mean a significant difference between the treatments.</p>
Full article ">
17 pages, 6715 KiB  
Article
A [3+3] Aldol-SNAr-Dehydration Approach to 2-Naphthol and 7-Hydroxyquinoline Derivatives
by Kwabena Fobi, Ebenezer Ametsetor and Richard A. Bunce
Molecules 2024, 29(14), 3406; https://doi.org/10.3390/molecules29143406 - 20 Jul 2024
Viewed by 591
Abstract
A one-pot [3+3] aldol-SNAr-dehydration annulation sequence was utilized to fuse hindered phenols onto aromatic substrates. The transformation joins doubly activated 1,3-disubstituted acetone derivatives (dinucleophiles) with C5-activated 2-fluorobenzaldehyde SNAr acceptors (dielectrophiles) in the presence of K2CO3 in [...] Read more.
A one-pot [3+3] aldol-SNAr-dehydration annulation sequence was utilized to fuse hindered phenols onto aromatic substrates. The transformation joins doubly activated 1,3-disubstituted acetone derivatives (dinucleophiles) with C5-activated 2-fluorobenzaldehyde SNAr acceptors (dielectrophiles) in the presence of K2CO3 in DMF at 65–70 °C to form polysubstituted 2-naphthols and 7-hydroxyquinolines. The reaction is regioselective in adding the most stable anionic center to the aldehyde followed by SNAr closure of the less stabilized anion to the electron-deficient aromatic ring. Twenty-seven examples are reported, and a probable mechanism is presented. In two cases where SNAr activation on the acceptor ring was lower (a C5 trifluoromethyl group on the aromatic ring or a 2-fluoropyridine), diethyl 1,3-acetonedicarboxylate initiated an interesting Grob-type fragmentation to give cinnamate esters as the products. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Recent drug candidates incorporating 2-naphthol.</p>
Full article ">Scheme 1
<p>Representative [3+3] annulation of 2-fluoro-5-nitrobenzaldehyde (dielectrophile) with dimethyl 1,3-acetonedicarboxylate (dinucleophile).</p>
Full article ">Scheme 2
<p>Synthesis of hindered hydroxyaromatics via [3+3] annulation.</p>
Full article ">Scheme 3
<p>A probable mechanism for the reaction of <b>1a</b> with methyl 4-phenyl-3-oxobutanoate (<b>8</b>) to give <b>20</b>.</p>
Full article ">Scheme 4
<p>Plausible mechanism for elimination from adduct of <b>5d</b> with <b>7</b> to obtain cinnamic ester derivative <b>40</b>. The numbers in green indicate the sequence of steps involved.</p>
Full article ">
13 pages, 2733 KiB  
Article
Insight into the Binding Interaction between PEDCs and hERRγ Utilizing Molecular Docking and Molecular Dynamics Simulations
by Fanqiang Bu, Lin Chen, Ying Sun, Bing Zhao and Ruige Wang
Molecules 2024, 29(14), 3256; https://doi.org/10.3390/molecules29143256 - 10 Jul 2024
Viewed by 690
Abstract
Phenolic environmental endocrine-disrupting chemicals (PEDCs) are persistent EDCs that are widely found in food packaging materials and environmental media and seriously threaten human health and ecological security. Human estrogen-related receptor γ (hERRγ) has been proposed as a mediator for the low-dose effects of [...] Read more.
Phenolic environmental endocrine-disrupting chemicals (PEDCs) are persistent EDCs that are widely found in food packaging materials and environmental media and seriously threaten human health and ecological security. Human estrogen-related receptor γ (hERRγ) has been proposed as a mediator for the low-dose effects of many environmental PEDCs; however, the atomic-level descriptions of dynamical structural features and interactions of hERRγ and PEDCs are still unclarified. Herein, how three PEDCs, 4-(1-methylpropyl)phenol (4-sec-butylphenol), 5,6,7,8-tetrahydro-2-naphthol (tetrahydro-2-napthol), and 2,2-bis(4-hydroxy-3,5-dimethoxyphenyl)propane (BP(2,2)(Me)), interact with hERRγ to produce its estrogenic disruption effects was studied. Molecular docking and multiple molecular dynamics (MD) simulations were first conducted to distinguish the detailed interaction pattern of hERRγ with PEDCs. These binding structures revealed that residues around Leu271, Leu309, Leu345, and Phe435 are important when binding with PEDCs. Furthermore, the binding energies of PEDCs with hERRγ were also characterized using the molecular mechanics/Poisson Boltzmann surface area (MM-PBSA) and solvated interaction energy (SIE) methods, and the results showed that the interactions of CH-π, π-π, and hydrogen bonds are the major contributors for hERRγ binding to these three PEDCs. What is striking is that the methoxide groups of BP(2,2)(Me), as hydrophobic groups, can help to reduce the binding energy of PEDCs binding with hERRγ. These results provide important guidance for further understanding the influence of PEDCs on human health problems. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of the complex of hERRγ with 4-sec-butylphenol ((<b>A</b>), PDB ID: 6I66) and the three PEDCs (<b>B</b>) in our study.</p>
Full article ">Figure 2
<p>Contributions of key residues when interacting with different PEDCs: (<b>A</b>) 4-sec-butylphenol; (<b>B</b>) tetrahydro-2-napthol; (<b>C</b>) BP(2,2) (Me).</p>
Full article ">Figure 3
<p>Nonpolar interaction and polar interaction energy of important residues when interacting with different PEDCs: (<b>A</b>) 4-sec-butylphenol; (<b>B</b>) tetrahydro-2-napthol; (<b>C</b>) BP(2,2)(Me).</p>
Full article ">Figure 4
<p>Interactions of different key residues when interacting with different PEDCs: (<b>A</b>) 4-sec-butylphenol (blue); (<b>B</b>) tetrahydro-2-napthol (magentas); (<b>C</b>) BP(2,2)(Me) (oranges).</p>
Full article ">
13 pages, 2114 KiB  
Article
Effects of Organic Xenobiotics on Tenebrio molitor Larvae and Their Parasite Gregarina polymorpha
by Viktoriia Lazurska and Viktor Brygadyrenko
Biology 2024, 13(7), 513; https://doi.org/10.3390/biology13070513 - 10 Jul 2024
Viewed by 788
Abstract
Environmental contamination with xenobiotics affects organisms and the symbiotic relations between them. A convenient object to study relationships between parasites and their hosts is the host–parasite system “Tenebrio molitor Linnaeus, 1758 (Coleoptera, Tenebrionidae)—Gregarina polymorpha (Hammerschmidt, 1838) Stein, 1848 (Eugregarinorida, Gregarinidae)”. For [...] Read more.
Environmental contamination with xenobiotics affects organisms and the symbiotic relations between them. A convenient object to study relationships between parasites and their hosts is the host–parasite system “Tenebrio molitor Linnaeus, 1758 (Coleoptera, Tenebrionidae)—Gregarina polymorpha (Hammerschmidt, 1838) Stein, 1848 (Eugregarinorida, Gregarinidae)”. For this experiment, we took 390 T. molitor larvae and 24 organic compounds. Groups of mealworms, 15 in each, were subjected to those compounds for 10 days. Then, we recorded the vitality of both the larvae of T. molitor and G. polymorpha. To assess how G. polymorpha had affected the hosts’ wellbeing, we looked for changes in the larvae’s body mass and compared them to the number of gregarines in their intestines. The vitality of the larvae was inhibited by cyclopentanol and 2-naphthol. The intensity of gregarine invasion was reduced by diphenyl ether, benzyl alcohol, catechol, and 3-aminobenzoic acid. No effect on the number of gregarines was produced by 3,4,5-trihydroxybenzoic acid, cyclohexanemethanol, phenol, benzalkonium chloride, maleic anhydride, cyclohexanol, resorcin, benzoic acid, 2-methylfuran, terpinen-4-ol, 1-phenylethylamine, dibutyl phthalate, 3-furancarboxylic acid, 5-methyl furfural, 6-aminohexanoic acid, succinic anhydride, o-xylene, and benzaldehyde. In the infected T. molitor individuals, the mean number of G. polymorpha equaled 45 specimens per host. The groups of smaller mealworms had fewer gregarines. Positive correlation was seen between growth rates of T. molitor larvae and the intensity of invasion by gregarines. Full article
(This article belongs to the Section Microbiology)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Gregarina polymorpha</span> in the intestines of <span class="html-italic">T. molitor</span> larvae: bar—200 µm; the sizes of the gamonts and syzygies of gregarines, as well as the ratio of morphometric indices for the primite and satellite (the length and width of the protomerite and deutomerite, the location and size of the nucleus, etc.) correspond to the data indicated in the taxonomic literature.</p>
Full article ">Figure 2
<p><span class="html-italic">Gregarina steini</span> in the intestines of <span class="html-italic">T. molitor</span> larvae: bar—100 µm; the body sizes of greragins and their morphometric indices correspond to the characteristics indicated in the work of Geus [<a href="#B56-biology-13-00513" class="html-bibr">56</a>].</p>
Full article ">Figure 3
<p>Number of <span class="html-italic">G. polymorpha</span> in the intestines of <span class="html-italic">T. molitor</span> at the end of the 10-day experiment (ordinate axis, spec.) in different variants of the experiment (N = 323): small square—median; lower and upper thresholds of the rectangle—the first and third quartiles, respectively; the lower and upper parts of the vertical line correspond to the minimum and maximum values of samplings; circles and stars <math display="inline"><semantics> <menclose notation="horizontalstrike"> <mo>ж</mo> </menclose> </semantics></math>—releases; different letters indicate values which reliably differed one from another within one line of the table according to the results of comparison using the Tukey test with Bonferroni correction.</p>
Full article ">Figure 4
<p>Dependence between the number of <span class="html-italic">G. polymorpha</span> (ordinate axis, number of specimens of gregarines per one <span class="html-italic">T. molitor</span>) and body mass of the insects (axis abscissa, mg of body mass): N = 323; legend—see <a href="#biology-13-00513-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>Dependence between the number of <span class="html-italic">G. polymorpha</span> (ordinate axis, number of gregarine specimens in one host) and change in body mass of the <span class="html-italic">T. molitor</span> larvae (abscissa axis, mg/day): N = 323; legend—see <a href="#biology-13-00513-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 6
<p>Changes in the body weight of <span class="html-italic">Tenebrio molitor</span> larvae at the end of a 10-day experiment (ordinate axis, mg) under the influence of various organic substances: N = 323; legend—see <a href="#biology-13-00513-f003" class="html-fig">Figure 3</a>.</p>
Full article ">
16 pages, 6723 KiB  
Article
Garlic Cellulosic Powders with Immobilized AgO and CuO Nanoparticles: Preparation, Characterization of the Nanocomposites, and Application to the Catalytic Degradation of Azo Dyes
by Nouha Sebeia, Mahjoub Jabli and Faridah Sonsudin
Polymers 2024, 16(12), 1661; https://doi.org/10.3390/polym16121661 - 12 Jun 2024
Viewed by 633
Abstract
Nanomaterials have attracted specific consideration due to their specific characteristics and uses in several promising fields. In the present study, Chondrilla juncea was employed as a biological extract to facilitate the reduction of copper and silver ions within garlic peel powders. The resulting [...] Read more.
Nanomaterials have attracted specific consideration due to their specific characteristics and uses in several promising fields. In the present study, Chondrilla juncea was employed as a biological extract to facilitate the reduction of copper and silver ions within garlic peel powders. The resulting garlic-CuO and garlic-AgO nanocomposites were characterized using several analytical methods including FTIR, TGA/DTG, SEM, TEM, and XRD analyses. The garlic peel exhibited a rough surface. The nanoparticles were evenly dispersed across its surface. The incorporation of CuO and AgO nanoparticles affected the crystal structure of garlic peel. The establishment of CuO and AgO nanoparticles was evidenced by the highest residual mass values observed for the prepared nanocomposites. The thermogravimetric analysis showed that the prepared nanocomposites had lower thermal stability compared with garlic peel powders. The prepared nanocomposites were used for catalytic degradation of naphthol blue black B and calmagite. The decolorization process depended on the quantity of H2O2, initial concentration of azo dyes, duration of contact, and temperature of the bath. The calculated activation energy (Ea) values for the garlic-CuO nanocomposites were found to be 18.44 kJ mol−1 and 23.28 kJ mol−1 for calmagite and naphthol solutions, respectively. However, those calculated for garlic-AgO nanocomposites were found to be 50.01 kJ mol−1 and 12.44 kJ mol−1 for calmagite and naphthol, respectively. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of (<b>a</b>) calmagite and (<b>b</b>) naphthol blue black B.</p>
Full article ">Figure 2
<p>Preparation of extract from <span class="html-italic">Chondrilla juncea.</span></p>
Full article ">Figure 3
<p>FT-IR of garlic powders, garlic powders-CuO, and garlic-AgO nanocomposites.</p>
Full article ">Figure 4
<p>SEM images (magnifications ×500 and ×1000) from top to bottom: (<b>a</b>) garlic peel, (<b>b</b>) garlic-CuO, and (<b>c</b>) garlic-AgO nanocomposites.</p>
Full article ">Figure 4 Cont.
<p>SEM images (magnifications ×500 and ×1000) from top to bottom: (<b>a</b>) garlic peel, (<b>b</b>) garlic-CuO, and (<b>c</b>) garlic-AgO nanocomposites.</p>
Full article ">Figure 5
<p>SEM images of garlic-AgO nanocomposites (magnification ×5000).</p>
Full article ">Figure 6
<p>XRD patterns of garlic peel powders, garlic-CuO (JCPDS No: 98-009-2367), and garlic-AgO (JCPDS No. 74-1750) nanocomposites.</p>
Full article ">Figure 7
<p>TEM pictures of (<b>a</b>) garlic-CuO and (<b>b</b>) garlic-AgO nanocomposites.</p>
Full article ">Figure 8
<p>TGA/DTG of (<b>a</b>) garlic peel powders, (<b>b</b>) garlic-CuO, and (<b>c</b>) garlic-AgO nanocomposites.</p>
Full article ">Figure 8 Cont.
<p>TGA/DTG of (<b>a</b>) garlic peel powders, (<b>b</b>) garlic-CuO, and (<b>c</b>) garlic-AgO nanocomposites.</p>
Full article ">Figure 9
<p>Effect of H<sub>2</sub>O<sub>2</sub> on the catalytic degradation of calmagite and naphthol (T = 20 °C, pH = 6, C<sub>0</sub> = 30 mg/L) solutions in the presence of garlic-CuO and garlic-AgO nanocomposites.</p>
Full article ">Figure 10
<p>Effect of initial concentrations of calmagite and naphthol solutions (T = 20 °C, pH = 6, H<sub>2</sub>O<sub>2</sub> = 10 mL/L) on catalytic degradation in the presence of garlic-CuO and garlic-AgO nanocomposites.</p>
Full article ">Figure 10 Cont.
<p>Effect of initial concentrations of calmagite and naphthol solutions (T = 20 °C, pH = 6, H<sub>2</sub>O<sub>2</sub> = 10 mL/L) on catalytic degradation in the presence of garlic-CuO and garlic-AgO nanocomposites.</p>
Full article ">Figure 11
<p>Effect of temperature on the catalytic degradation of calmagite and naphthol solutions (pH = 6, C<sub>0</sub> = 30 mg/L, H<sub>2</sub>O<sub>2</sub> = 10 mL/L) in the presence of garlic-CuO and garlic-AgO nanocomposites.</p>
Full article ">Figure 12
<p>(<b>a</b>) Arrhenius plot and (<b>b</b>) plot of Ln (<span class="html-italic">K</span><sub>2</sub>/<span class="html-italic">T</span>) vs. 1/<span class="html-italic">T</span> for calmagite and naphthol degradation using garlic-CuO nanocomposites.</p>
Full article ">Figure 13
<p>(<b>a</b>) Arrhenius plot and (<b>b</b>) plot of Ln (<span class="html-italic">K</span><sub>2</sub>/<span class="html-italic">T</span>) vs. 1/<span class="html-italic">T</span> for calmagite and naphthol degrading using garlic-AgO nanocomposites.</p>
Full article ">
19 pages, 3770 KiB  
Article
Using the Assembly Time as a Tool to Control the Surface Morphology and Separation Performance of Membranes with a Tannic Acid–Fe3+ Selective Layer
by Hluf Hailu Kinfu, Md. Mushfequr Rahman, Erik S. Schneider, Nicolás Cevallos-Cueva and Volker Abetz
Membranes 2024, 14(6), 133; https://doi.org/10.3390/membranes14060133 - 6 Jun 2024
Viewed by 997
Abstract
Thin-film composite (TFC) membranes containing a metal–polyphenol network (MPN)-based selective layer were fabricated on a porous polyacrylonitrile support. The MPN layer was formed through coordination-based self-assembly between plant-based tannic acid (TA) and an Fe3+ ion. For the first time, we demonstrate that [...] Read more.
Thin-film composite (TFC) membranes containing a metal–polyphenol network (MPN)-based selective layer were fabricated on a porous polyacrylonitrile support. The MPN layer was formed through coordination-based self-assembly between plant-based tannic acid (TA) and an Fe3+ ion. For the first time, we demonstrate that TFC membranes containing TA-Fe3+ selective layers can separate small organic solutes in aqueous media from equimolar mixtures of solutes. The effect of the assembly time on the characteristics and performance of the fabricated selective layer was investigated. An increase in the assembly time led to the formation of selective layers with smaller effective pore sizes. The tannic acid–Fe3+ selective layer exhibited a low rejection towards neutral solutes riboflavin and poly(ethylene glycol) while high rejections were observed for anionic dyes of orange II and naphthol green B. Permeation selectivities in the range of 2–27 were achieved between neutral and charged dyes in both single- and mixed-solute experiments, indicating the significant role of Donnan exclusion and the charge-selective nature of the membranes. The rejection efficiency improved with an increasing assembly time. Overall, this study demonstrates that the assembly time is a vital casting parameter for controlling the permeance, rejection and selectivity of thin-film composite membranes with a tannic acid–Fe3+ selective layer. Full article
(This article belongs to the Collection New Challenges in Membranes for Water and Wastewater Application)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the fabrication process of the membrane selective layer. TFC membrane containing TA-Fe<sup>3+</sup> is fabricated by depositing tannic acid and iron aqueous solutions over a porous PAN support in a layer-by-layer technique.</p>
Full article ">Figure 2
<p>(<b>a</b>) FTIR spectra of the PAN and TA-Fe<sup>3+</sup> membranes fabricated at different assembly times. (<b>b</b>) Backscattered electron SEM image of a TA-Fe<sup>3+</sup> TFC membrane sample showing a continuous, thin, Fe-rich selective layer deposited over the porous substrate. Elemental analysis of the PAN support and the TFC membranes: (<b>c</b>) EDX spectra and (<b>d</b>) Fe/O ratio of the TFC membranes. Inset in (<b>c</b>) displays a zoomed−in figure showing the peaks for oxygen and iron. (<b>e</b>) Surface zeta potential as a function of pH for the PAN support and TA-Fe<sup>3+</sup> membranes fabricated at different assembly times. (<b>f</b>) A schematic illustration of the TA–metal ion coordination for thin-film formation by the LBL technique. MPN-1, MPN-2.5, MPN-4 and MPN-6 represent TA-Fe<sup>3+</sup> TFC membranes fabricated at 1 min, 2.5 min, 4 min and 6 min of assembly time, respectively.</p>
Full article ">Figure 3
<p>Top surfaces and corresponding cross-sectional SEM images of (<b>a</b>) the PAN support, and TA-Fe<sup>3+</sup> TFC membranes synthesized at (<b>b</b>) 1 min (MPN-1), (<b>c</b>) 2.5 min (MPN-2.5), (<b>d</b>) 4 min (MPN-4) and (<b>e</b>) 6 min (MPN-6) of assembly time.</p>
Full article ">Figure 4
<p>(<b>a</b>) Water contact angle and (<b>b</b>) pure water permeance of the PAN support and TA-Fe<sup>3+</sup> membranes fabricated at different assembly times. Retention of a single solute—(<b>c</b>) poly(ethylene glycol) (neutral), (<b>d</b>) orange II (–1 charge), (<b>e</b>) riboflavin (neutral) and (<b>f</b>) naphthol green B (–3 charge)—by the TA-Fe<sup>3+</sup> membranes as a function of the assembly time of the membranes. The solid line is added to guide the eye.</p>
Full article ">Figure 5
<p>(<b>a</b>) UV-vis absorption spectra of the feed and permeate solutions and (<b>b</b>) retention of the 1:1 molar mixture of riboflavin and naphthol green B dyes, with peaks at 444 nm and 715 nm, respectively, through the TA-Fe<sup>3+</sup> membranes synthesized at different assembly times. Inset: photographic images of the mixed dye solutions before (feed, F) and after (permeate, P) filtration experiments. Numerical values of 1, 2.5, 4 and 6 represent the assembly time (min) durations at which the membranes were fabricated by exposing the PAN surface to the TA and FeCl<sub>3</sub>·6H<sub>2</sub>O solutions. (<b>c</b>) UV-vis absorption spectra of the feed and permeate solutions and (<b>d</b>) retention of the dyes in the OR-/NGB3- system of the 1:1 molar mixture of orange II and naphthol green B dyes, with UV-vis absorption peaks at 482 nm and 715 nm, respectively, through the TA-Fe<sup>3+</sup> membranes synthesized at different assembly times. (<b>e</b>,<b>f</b>) Evaluation of the antifouling properties of the metal–phenolic selective layers; (<b>e</b>) normalized water and HA solution flux of the TA-Fe<sup>3+</sup> membranes and (<b>f</b>) flux recovery ratios of the synthesized membranes. The investigation was performed by first recording the PWP of a membrane sample for 1 h followed by HA solution filtration for 1.5 h. Then, the membrane was rinsed with pure water to remove fouled HA molecules, after which the PWP of the fouled membrane was measured for one hour again to determine the FRR. This process was repeated for two cycles.</p>
Full article ">
22 pages, 12069 KiB  
Article
Application of Three-Dimensional Porous Aerogel as Adsorbent for Removal of Textile Dyes from Water
by Monika Liugė, Dainius Paliulis and Teresė Leonavičienė
Appl. Sci. 2024, 14(10), 4274; https://doi.org/10.3390/app14104274 - 17 May 2024
Cited by 1 | Viewed by 891
Abstract
The textile industry is one of the most important industries in the European Union. The main environmental problems of the textile industry are the high water consumption, the generated pollution, the variety of chemicals used and the high energy demand. Recently, adsorbents with [...] Read more.
The textile industry is one of the most important industries in the European Union. The main environmental problems of the textile industry are the high water consumption, the generated pollution, the variety of chemicals used and the high energy demand. Recently, adsorbents with a large specific surface area and low weight, such as aerogels, have attracted great interest as promising materials for removing dyes from polluted water. Cellulose aerogels are inexpensive and non-toxic. Langmuir and Freundlich isotherms were chosen as the best method to describe the performance of the adsorbent. In this study, the adsorption efficiency of Congo red, Naphthol green B, Rhodamine B and Methylene blue were determined by using an adsorbent synthesized from paper and cardboard waste. The total organic carbon concentration was chosen as an indicator of the concentration of the dyes in the solutions. The aerogel capsules had 5% cellulose content. It was found that the adsorption capacity of the aerogel in the solutions of Congo red varied from 0.028 mg/g to 14.483 mg/g; in the solutions of Naphthol green B, from 0.013 mg/g to 7.698 mg/g; in the solutions of Rhodamine B, from 0.020 mg/g to 8.768 mg/g; and in the solutions of Methylene blue, from 0.024 mg/g to 13.538 mg/g. Full article
Show Figures

Figure 1

Figure 1
<p>Photo of aerogel surface, top view.</p>
Full article ">Figure 2
<p>Determination of contact angle of aerogel.</p>
Full article ">Figure 3
<p>Surface area of aerogel captured before adsorption process with SEM Helios Nanolab 650: (<b>a</b>) 400 μm; (<b>b</b>) 5 μm.</p>
Full article ">Figure 4
<p>Surface area of aerogel captured after adsorption process with SEM Helios Nanolab 650: (<b>a</b>) 400 μm; (<b>b</b>) 5 μm.</p>
Full article ">Figure 5
<p>The linearized form of the Langmuir isotherm for Congo red.</p>
Full article ">Figure 6
<p>The linearized form of the Freundlich isotherm for Congo red.</p>
Full article ">Figure 7
<p>Equilibrium analysis for Congo red.</p>
Full article ">Figure 8
<p>The linearized form of the Langmuir isotherm for Naphthol green B.</p>
Full article ">Figure 9
<p>The linearized form of the Freundlich isotherm for Naphthol green B.</p>
Full article ">Figure 10
<p>Equilibrium analysis for Naphthol green B.</p>
Full article ">Figure 11
<p>The linearized form of the Langmuir isotherm for Rhodamine B.</p>
Full article ">Figure 12
<p>The linearized form of the Freundlich isotherm for Rhodamine B.</p>
Full article ">Figure 13
<p>Equilibrium analysis for Rhodamine B.</p>
Full article ">Figure 14
<p>The linearized form of the Langmuir isotherm for Methylene blue.</p>
Full article ">Figure 15
<p>The linearized form of the Freundlich isotherm for Methylene blue.</p>
Full article ">Figure 16
<p>Equilibrium analysis for Methylene blue.</p>
Full article ">Figure 17
<p>The kinetic experimental data with the pseudo-first-order and the pseudo-second-order kinetic models using the obtained constants (<a href="#applsci-14-04274-t004" class="html-table">Table 4</a> and <a href="#applsci-14-04274-t005" class="html-table">Table 5</a>).</p>
Full article ">Figure 18
<p>The kinetic experimental data with the pseudo-first-order and the pseudo-second-order kinetic models using the obtained constants (<a href="#applsci-14-04274-t006" class="html-table">Table 6</a> and <a href="#applsci-14-04274-t007" class="html-table">Table 7</a>).</p>
Full article ">Figure 19
<p>The kinetic experimental data with the pseudo-first-order and the pseudo-second-order kinetic models using the obtained constants (<a href="#applsci-14-04274-t008" class="html-table">Table 8</a> and <a href="#applsci-14-04274-t009" class="html-table">Table 9</a>).</p>
Full article ">Figure 20
<p>The kinetic experimental data with the pseudo-first-order and the pseudo-second-order kinetic models using the obtained constants (<a href="#applsci-14-04274-t010" class="html-table">Table 10</a> and <a href="#applsci-14-04274-t011" class="html-table">Table 11</a>).</p>
Full article ">
11 pages, 2511 KiB  
Article
Highly Stable Hybrid Pigments Prepared from Organic Chromophores and Fluorinated Hydrotalcites
by Magali Hernández, Carlos Felipe, Ariel Guzmán-Vargas, José Luis Rivera and Enrique Lima
Colorants 2024, 3(2), 125-135; https://doi.org/10.3390/colorants3020009 - 9 May 2024
Viewed by 612
Abstract
Structural hydroxide groups in layered magnesium–aluminum double hydroxides were partially replaced by fluoride ions. Fluorinated and fluorine-free materials were used as hosts for two dyes, carminic acid and hydroxyl naphthol blue, resulting in a hybrid pigment color palette. The pigments were produced by [...] Read more.
Structural hydroxide groups in layered magnesium–aluminum double hydroxides were partially replaced by fluoride ions. Fluorinated and fluorine-free materials were used as hosts for two dyes, carminic acid and hydroxyl naphthol blue, resulting in a hybrid pigment color palette. The pigments were produced by two ways, either incorporating chromophore during the synthesis of the layered double hydroxide or in a post-synthesis step through the memory effect of the LDHs. Additionally, the pigments were protected with a magnesium hydroxide phase to prevent the color from fading over time. The pigments were stable for periods as long as 10 years. The color properties of the pigments were significantly influenced by the host of dye since the presence of fluorine directly influences the acid–base properties of the layered double hydroxides. The pigments conferred their color to white cream in the preparation of colored creams. The colored creams acquired the color of the layered pigment. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of pigments generated by sorption of carminic acid onto layered double hydroxides. (<b>a</b>) LDH-CA, (<b>b</b>) FLDH-CA, (<b>c</b>) FLDH-CA + MgO, (<b>d</b>) ME-FLDH-CA, and (<b>e</b>) ME-FLDH-CA + MgO.</p>
Full article ">Figure 2
<p>Multinuclear solid-state NMR spectra of pigments. (<b>a</b>) <sup>1</sup>H→<sup>13</sup>C CP/MAS NMR, (<b>b</b>) <sup>27</sup>Al MAS NMR, and (<b>c</b>) <sup>19</sup>F MAS NMR.</p>
Full article ">Figure 3
<p>Pigments generated by the sorption of carminic acid onto layered double hydroxides. As a reference, a white LDH powder is included.</p>
Full article ">Figure 4
<p>UV-vis spectra of carminic acid and pigments generated by sorption of carminic acid onto layered double hydroxides. (<b>a</b>) CA, (<b>b</b>) LDH-CA, (<b>c</b>) FLDH-CA, (<b>d</b>) FLDH-CA + MgO, (<b>e</b>) ME-FLDH-CA, and (<b>f</b>) ME-FLDH-CA + MgO.</p>
Full article ">Figure 5
<p>XRD patterns of pigments aged for 10 years. (<b>a</b>) LDH-CA, (<b>b</b>) FLDH-CA, (<b>c</b>) FLDH-CA + MgO, (<b>d</b>) ME-FLDH-CA, and (<b>e</b>) ME-FLDH-CA + MgO.</p>
Full article ">Figure 6
<p><sup>1</sup>H→<sup>13</sup>C CP/MAS spectra of aged pigments. (<b>a</b>) LDH-CA, (<b>b</b>) FLDH-CA, (<b>c</b>) FLDH-CA + MgO, and (<b>d</b>) ME-FLDH-CA + MgO.</p>
Full article ">Figure 7
<p>Pigmented emollient creams prepared from white cream (<b>a</b>) and several pigments. (<b>b</b>) LDH-CA, (<b>c</b>) FLDH-CA, (<b>d</b>) FLDH-CA + MgO, (<b>e</b>) ME-FLDH-CA, and (<b>f</b>) ME-FLDH-CA + MgO. As a reference, the white pigment made using FLDH is shown in (<b>g</b>).</p>
Full article ">
22 pages, 6052 KiB  
Article
Photocatalytic Degradation of Tartrazine and Naphthol Blue Black Binary Mixture with the TiO2 Nanosphere under Visible Light: Box-Behnken Experimental Design Optimization and Salt Effect
by Fadimatou Hassan, Bouba Talami, Amira Almansba, Pierre Bonnet, Christophe Caperaa, Sadou Dalhatou, Abdoulaye Kane and Hicham Zeghioud
ChemEngineering 2024, 8(3), 50; https://doi.org/10.3390/chemengineering8030050 - 3 May 2024
Cited by 2 | Viewed by 1294
Abstract
In this study, TiO2 nanospheres (TiO2-NS) were synthesized by the solvothermal method. Firstly, the synthesized nanomaterial was characterized by X-ray diffraction (XRD), Fourier Transformed Infrared (FTIR), scanning electron microscopy (SEM) and UV-Vis Diffuse Reflectance Spectroscopy (DRS). To study the photocatalytic [...] Read more.
In this study, TiO2 nanospheres (TiO2-NS) were synthesized by the solvothermal method. Firstly, the synthesized nanomaterial was characterized by X-ray diffraction (XRD), Fourier Transformed Infrared (FTIR), scanning electron microscopy (SEM) and UV-Vis Diffuse Reflectance Spectroscopy (DRS). To study the photocatalytic degradation of Tartrazine (TTZ) and Naphthol Blue Black (NBB) in a binary mixture, the influence of some key parameters such as pH, pollutant concentration and catalyst dose was taken into account under visible and UV light. The results show a 100% degradation efficiency for TTZ after 150 min of UV irradiation and 57% under visible irradiation at 180 min. The kinetic study showed a good pseudo-first-order fit to the Langmuir–Hinshelwood model. Furthermore, in order to get closer to the real conditions of textile wastewater, the influence of the presence of salt on TiO2-NS’s photocatalytic performance was explored by employing NaCl as an inorganic ion. The optimum conditions provided by the Response Surface Methodology (RSM) were low concentrations of TTZ (2 ppm) and NBB (2.33 ppm) and negligible salt (NaCl) interference. The percentage of photodegradation was high at low pollutant and NaCl concentrations. However, this yield became very low as NaCl concentrations increased. The photocatalytic treatment leads to 31% and 53% of mineralization yield after 1 and 3 h of visible light irradiation. The synthesis of TiO2-NS provides new insights that will help to develop an efficient photocatalysts for the remediation of contaminated water. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD diffractogram, (<b>b</b>) Raman spectra, (<b>c</b>) FTIR spectra and (<b>d</b>) Elemental Composition of TiO<sub>2</sub>−NS.</p>
Full article ">Figure 2
<p>SEM images of TiO<sub>2</sub>-NS at different magnifications: (<b>a</b>) ×33,000 (0.5 μm), (<b>b</b>) ×18,000 (1 μm), (<b>c</b>) ×12,000 (1 μm) and (<b>d</b>) ×6000 (2 μm).</p>
Full article ">Figure 3
<p>(<b>a</b>) Diffuse reflectance spectra and (<b>b</b>) Plot of transferred Kubelka-Munk Versus energy of TiO<sub>2</sub>−NS.</p>
Full article ">Figure 4
<p>Adsorption equilibrium and photocatalytic degradation of tartrazine with TiO<sub>2</sub>−NS catalyst under UV and visible light (C<sub>0</sub>: 5 ppm. C<sub>TiO<sub>2</sub>−NS</sub>: 0.2 g/L. V: 200 mL. natural pH: 6).</p>
Full article ">Figure 5
<p>Effect of catalyst dose on tartrazine degradation under visible light (C<sub>0</sub>: 5 ppm. V<sub>solution</sub>: 200 mL. Natural pH: 6).</p>
Full article ">Figure 6
<p>(<b>a</b>) Effect of initial Tartrazine concentration with 200 mL of solution and 40 mg of catalyst at natural pH, under visible light irradiation; (<b>b</b>) PFO kinetics for tartrazine degradation under visible light ([TTZ]<sub>0</sub> = 2–12 ppm. V<sub>solution</sub>: 200 mL. Natural pH: 6. C<sub>TiO<sub>2</sub>-NS</sub>: 0.2 g/L. Reaction time = 180 min), (<b>c</b>) Langmuir–Hinshelwood plot for photodegradation of tartrazine under visible light ([TTZ]<sub>0</sub> = 2–12 ppm. V<sub>solution</sub>: 200 mL.Natural pH: 6. C<sub>TiO<sub>2</sub>-NS</sub>: 0.2 g/L. Reaction time = 180 min).</p>
Full article ">Figure 7
<p>Photocatalytic degradation of binary solution of TTZ and NBB with TiO<sub>2</sub> nanosphere catalyst under visible light at different pH values (C<sub>TTZ</sub>: 2 ppm. C<sub>NBB</sub>: 2.33 ppm. C<sub>TiO<sub>2</sub>-NS</sub>: 0.2 g/L. V<sub>solution</sub>: 200 mL. Treatment duration: 120 min, 10 ppm of NaCl presence).</p>
Full article ">Figure 8
<p>(<b>a</b>) Photocatalytic degradation of Tartrazine (TTZ) (C<sub>NBB</sub>: 2.33 ppm. m<sub>TiO<sub>2</sub>-NS</sub>: 40 mg. V<sub>solution</sub>: 200 mL. Natural pH: 6, 10 ppm of NaCl presence) and of (<b>b</b>) Naphthol Blue Black (NBB) (C<sub>TTZ</sub>: 2 ppm. m<sub>TiO<sub>2</sub>-NS</sub>: 40 mg. V<sub>solution</sub>: 200 mL. natural pH: 6, 10 ppm of NaCl presence) in binary solution under visible light.</p>
Full article ">Figure 9
<p>Predicted vs. experimental results of degradation efficiency in binary system: (<b>a</b>) Tartrazine; (<b>b</b>) Naphthol Blue Black.</p>
Full article ">Figure 10
<p>RSM surfaces plots and 2D contour plots of the interaction effects between: (<b>a</b>) TTZ and NBB concentrations; (<b>b</b>) TTZ and NaCl concentrations and (<b>c</b>) NaCl and NBB concentrations.</p>
Full article ">Figure 10 Cont.
<p>RSM surfaces plots and 2D contour plots of the interaction effects between: (<b>a</b>) TTZ and NBB concentrations; (<b>b</b>) TTZ and NaCl concentrations and (<b>c</b>) NaCl and NBB concentrations.</p>
Full article ">Figure 11
<p>RSM surfaces plots and 2D contour plots of the interaction effects between (<b>a</b>) NBB and TTZ concentrations; (<b>b</b>) NaCl and Tartrazine concentrations; and (<b>c</b>) NBB and NaCl concentrations.</p>
Full article ">Figure 11 Cont.
<p>RSM surfaces plots and 2D contour plots of the interaction effects between (<b>a</b>) NBB and TTZ concentrations; (<b>b</b>) NaCl and Tartrazine concentrations; and (<b>c</b>) NBB and NaCl concentrations.</p>
Full article ">Figure 12
<p>Mineralization of tartrazine with TiO<sub>2</sub> nanosphere catalyst under visible light (C<sub>0</sub>: 6 ppm. C<sub>TiO<sub>2</sub>-NS</sub>: 0.2 g/L. V<sub>solution</sub>: 200 mL. Natural pH: 6).</p>
Full article ">Figure 13
<p>Reusability cycles of TiO<sub>2</sub>-NS for photocatalytic degradation of tartrazine under visible light (C<sub>0</sub>: 6 ppm. m<sub>TiO<sub>2</sub>-NS</sub>: 100 mg. V<sub>solution</sub>: 200 mL. natural pH: 6).</p>
Full article ">
17 pages, 5306 KiB  
Article
Gas Chromatography–Mass Spectrometry Chemical Profiling of Commiphora myrrha Resin Extracts and Evaluation of Larvicidal, Antioxidant, and Cytotoxic Activities
by Naimah Asid H. Alanazi, Abdullah A. Alamri, Abadi M. Mashlawi, Nujud Almuzaini, Gamal Mohamed and Salama A. Salama
Molecules 2024, 29(8), 1778; https://doi.org/10.3390/molecules29081778 - 13 Apr 2024
Viewed by 1742
Abstract
Plant extracts and essential oils can be alternative environmentally friendly agents to combat pathogenic microbes and malaria vectors. Myrrh is an aromatic oligum resin that is extracted from the stem of Commiphora spp. It is used in medicine as an insecticide, cytotoxic, and [...] Read more.
Plant extracts and essential oils can be alternative environmentally friendly agents to combat pathogenic microbes and malaria vectors. Myrrh is an aromatic oligum resin that is extracted from the stem of Commiphora spp. It is used in medicine as an insecticide, cytotoxic, and aromatic. The current study assessed the effect of Commiphora myrrha resin extracts on the biological potency of the third larval stage of Aedes aegypti, as well as its antioxidant and cytotoxic properties against two types of tumor cells (HepG-2 and Hela cell lines). It also used GC–MS to determine the chemical composition of the C. myrrha resin extracts. Fifty components from the extracted plant were tentatively identified using the GC–MS method, with curzerene (33.57%) typically listed as the primary ingredient, but other compounds also make up a significant portion of the mixture, including 1-Methoxy-3,4,5,7-tetramethylnaphthalene (15.50%), β-Elemene (5.80%), 2-Methoxyfuranodiene (5.42%), 2-Isopropyl-4,7-Dimethyl-1-Naphthol (4.71%), and germacrene B (4.35%). The resin extracts obtained from C. myrrha exhibited significant efficacy in DPPH antioxidant activity, as evidenced by an IC50 value of 26.86 mg/L and a radical scavenging activity percentage of 75.06%. The 50% methanol extract derived from C. myrrha resins exhibited heightened potential for anticancer activity. It demonstrated substantial cytotoxicity against HepG-2 and Hela cells, with IC50 values of 39.73 and 29.41 µg mL−1, respectively. Notably, the extract showed non-cytotoxic activity against WI-38 normal cells, with an IC50 value exceeding 100 µg mL−1. Moreover, the selectivity index for HepG-2 cancer cells (2.52) was lower compared to Hela cancer cells (3.40). Additionally, MeOH resin extracts were more efficient against the different growth stages of the mosquito A. aegypti, with lower LC50, LC90, and LC95 values of 251.83, 923.76, and 1293.35 mg/L, respectively. In comparison to untreated groups (1454 eggs/10 females), the average daily number of eggs deposited (424 eggs/L) decreases at higher doses (1000 mg/L). Finally, we advise continued study into the possible use of C. myrrha resins against additional pests that have medical and veterinary value, and novel chemicals from this extract should be isolated and purified for use in medicines. Full article
(This article belongs to the Section Analytical Chemistry)
Show Figures

Figure 1

Figure 1
<p>Chromatogram of the methanol extract of <span class="html-italic">C. myrrha</span> resins by GC–MS.</p>
Full article ">Figure 2
<p>IC<sub>50</sub> values of the tested plant-resin extracts and doxorubicin as standard against human cancer and normal cells.</p>
Full article ">Figure 3
<p><span class="html-italic">Commiphora myrrha</span> (Nees) Engl. (<b>a</b>) overview of the growing tree, (<b>b</b>) a close-up of tree bark demonstrating resin accumulation from scraping.</p>
Full article ">
12 pages, 6913 KiB  
Article
A “Pincer” Type of Acridine–Triazole Fluorescent Dye for Iodine Detection by Both ‘Naked-Eye’ Colorimetric and Fluorometric Modes
by Mei Yu, Lu Jiang, Lan Mou, Xi Zeng, Ruixiao Wang, Tao Peng, Fuyong Wu and Tianzhu Shi
Molecules 2024, 29(6), 1355; https://doi.org/10.3390/molecules29061355 - 19 Mar 2024
Viewed by 1056
Abstract
Iodine, primarily in the form of iodide (I), is the bioavailable form for the thyroid in the human body. Both deficiency and excess intake of iodide can lead to serious health issues, such as thyroid disease. Selecting iodide ions among anions [...] Read more.
Iodine, primarily in the form of iodide (I), is the bioavailable form for the thyroid in the human body. Both deficiency and excess intake of iodide can lead to serious health issues, such as thyroid disease. Selecting iodide ions among anions has been a significant challenge for decades due to interference from other anions. In this study, we designed and synthesized a new pincer-type acridine–triazole fluorescent probe (probe 1) with an acridine ring as a spacer and a triazole as a linking arm attached to two naphthol groups. This probe can selectively recognize iodide ions in a mixed solvent of THF/H2O (v/v, 9/1), changing its color from colorless to light yellow, making it suitable for highly sensitive and selective colorimetric and fluorescent detection in water systems. We also synthesized another molecular tweezer-type acridine–triazole fluorescent probe (probe 2) that exhibits uniform detection characteristics for iodide ions in the acetonitrile system. Interestingly, compared to probe 2, probe 1 can be detected by the naked eye due to its circulation effect, providing a simple method for iodine detection. The detection limit of probe 1 is determined to be 10−8 mol·L−1 by spectrometric titration and isothermal titration calorimetry measurements. The binding stoichiometry between probe 1 and iodide ions is calculated to be 1:1 by these methods, and the binding constant is 2 × 105 mol·L−1. Full article
Show Figures

Figure 1

Figure 1
<p>UV-Vis spectra of probe <b>1</b> for (<b>a</b>) (10 µmol·L<sup>−1</sup>, THF/H<sub>2</sub>O, <span class="html-italic">v</span>/<span class="html-italic">v</span>, 9/1) and probe <b>2</b> for (<b>b</b>) (10 µmol·L<sup>−1</sup>, Acetonitrile) in different anions (20 µmol·L<sup>−1</sup>); anions include I<sup>−</sup>, F<sup>−</sup>, Cl<sup>−</sup>, Br<sup>−</sup>, NO<sub>3</sub><sup>−</sup>, HSO<sub>4</sub><sup>−</sup>, ClO<sub>4</sub><sup>−</sup>, PF<sub>6</sub><sup>−</sup>, AcO<sup>−</sup>, and H<sub>2</sub>PO<sub>4</sub><sup>−</sup>.</p>
Full article ">Figure 2
<p>Emission spectra of probe <b>1</b> for (<b>a</b>) (10 µmol·L<sup>−1</sup>, THF/H<sub>2</sub>O, <span class="html-italic">v</span>/<span class="html-italic">v</span>, 9/1) and probe <b>2</b> for (<b>b</b>) (10 µmol·L<sup>−1</sup>, Acetonitrile) in different anions (20 µmol·L<sup>−1</sup>); anions include I<sup>−</sup>, F<sup>−</sup>, Cl<sup>−</sup>, Br<sup>−</sup>, NO<sub>3</sub><sup>−</sup>, HSO<sub>4</sub><sup>−</sup>, ClO<sub>4</sub><sup>−</sup>, PF<sub>6</sub><sup>−</sup>, AcO<sup>−</sup>, and H<sub>2</sub>PO<sub>4</sub><sup>−</sup>. Probe 1: λex/λem = 357/427 nm, probe <b>2</b>: λex/λem = 247/425 nm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Curves of absorbance spectrum titration of probe <b>1</b> (10 µmol·L<sup>−1</sup>, THF/H<sub>2</sub>O, <span class="html-italic">v</span>/<span class="html-italic">v</span>, 9/1). (<b>b</b>) Curves of Emission spectra titration of probe <b>1</b> (10 µmol·L<sup>−1</sup>, THF/H<sub>2</sub>O, <span class="html-italic">v</span>/<span class="html-italic">v</span>, 9/1). Inset is the titration profile and Job’s plot. λex/λem = 357/427 nm. (<b>c</b>) Fluorescence response of probe <b>1</b> (10 µmol·L<sup>−1</sup>, THF/H<sub>2</sub>O, <span class="html-italic">v</span>/<span class="html-italic">v</span>, 9/1) in the presence of 0.4 mM different anions. Inset is the comparison photos of the color and fluorescence changes with different anions. (<b>d</b>) Curves of Emission spectra titration of probe <b>2</b> (10 µmol·L<sup>−1</sup>, Acetonitrile). Inset is the molar ratio’s plot (not connected squares) and Job’s plot (connected squares). λex/λem = 247/425 nm.</p>
Full article ">Figure 4
<p>The effect of coexisting anions to probe <b>1</b>-I<sup>−</sup> for (<b>a</b>) and probe <b>2</b>-I<sup>−</sup> complex for (<b>b</b>). (<b>a</b>) The dark bars represent the emission intensity of probe <b>1</b> (1.00 × 10<sup>−5</sup> mol·L<sup>−1</sup>, THF/H<sub>2</sub>O, <span class="html-italic">v</span>/<span class="html-italic">v</span>, 9/1) in the presence of 20 equiv. different anions. The white bars represent the emission intensity of probe <b>1</b>-I<sup>−</sup> complex solution after the addition of other anions. Other anions include I<sup>−</sup>, F<sup>−</sup>, Cl<sup>−</sup>, Br<sup>−</sup>, NO<sub>3</sub><sup>−</sup>, HSO<sub>4</sub><sup>−</sup>, ClO<sub>4</sub><sup>−</sup>, PF<sub>6</sub><sup>−</sup>, AcO<sup>−</sup>, and H<sub>2</sub>PO<sub>4</sub><sup>−</sup>; λex /λem = 357/427 nm. (<b>b</b>) The dark bars represent the emission intensity of probe <b>2</b> (1.00 × 10<sup>−5</sup> mol·L<sup>−1</sup>, acetonitrile) in the presence of 20 equiv. different anions. The white bars represent the emission intensity of probe <b>2</b>-I<sup>−</sup> complex solution after the addition of other anions. Other anions include I<sup>−</sup>, F<sup>−</sup>, Cl<sup>−</sup>, Br<sup>−</sup>, NO<sub>3</sub><sup>−</sup>, HSO<sub>4</sub><sup>−</sup>, ClO<sub>4</sub><sup>−</sup>, PF<sub>6</sub><sup>−</sup>, AcO<sup>−</sup>, and H<sub>2</sub>PO<sub>4</sub><sup>−</sup>; λex /λem = 247/425 nm.</p>
Full article ">Figure 5
<p>Fluorescence spectroscopy of reversible reaction of probe <b>1</b> (1.00 × 10<sup>−5</sup> mol·L<sup>−1</sup>, THF/H<sub>2</sub>O, <span class="html-italic">v</span>/<span class="html-italic">v</span>, 9/1) with I<sup>−</sup>.</p>
Full article ">Figure 6
<p>Isothermal titration calorimetry profile of probe <b>1</b> (1.00 × 10<sup>−4</sup> mol·L<sup>−1</sup>) with I<sup>−</sup> (1.00 × 10<sup>−3</sup> mol·L<sup>−1</sup>) (<b>a</b>) and non-linear curve fitting for the binding of probe <b>1</b> (1.00 × 10<sup>−4</sup> mol·L<sup>−1</sup>, 298.15 K) (<b>b</b>).</p>
Full article ">Figure 7
<p>(<b>A</b>) The changes of partial <sup>1</sup>H NMR of <b>1</b> upon titration with I<sup>−</sup>: a, b, c, d, e, and f in the presence of 0, 0.25, 0.50, 1.0, 2.0, 3.0 equiv. of I- at 400 MHz in CDCl<sub>3</sub>/CD<sub>3</sub>CN, 1/1(<span class="html-italic">v</span>/<span class="html-italic">v</span>), respectively. (<b>B</b>) <sup>1</sup>H NMR shift of Ha and Hb in 1 with I<sup>−</sup>.</p>
Full article ">Scheme 1
<p>Synthesis of probes <b>1</b> and <b>2</b>.</p>
Full article ">
17 pages, 5372 KiB  
Article
Comparison between Electrooxidation of 1-Naphthol and 2-Naphthol in Different Non-Aqueous Solvents and Suppression of Layer Growth of Polymers
by László Kiss, Péter Szabó and Sándor Kunsági-Máté
Surfaces 2024, 7(1), 164-180; https://doi.org/10.3390/surfaces7010011 - 15 Mar 2024
Cited by 1 | Viewed by 1367
Abstract
The two naphthol isomers were investigated in different organic solvents by taking cyclic voltammograms, and fouling took place on a platinum electrode surface, except for dimethyl sulfoxide and dimethyl formamide. Studies in allyl alcohol rarely used in electrochemical investigations pointed to the importance [...] Read more.
The two naphthol isomers were investigated in different organic solvents by taking cyclic voltammograms, and fouling took place on a platinum electrode surface, except for dimethyl sulfoxide and dimethyl formamide. Studies in allyl alcohol rarely used in electrochemical investigations pointed to the importance of the carbon–carbon double bond as electrode deactivation was remarkably faster compared with its saturated analog solvent. Similarly, the use of the other unsaturated solvent mesityl oxide in the electropolymerization of naphthols resulted in different findings compared with methyl isobutyl ketone. As dimethyl formamide was the best choice concerning the solubility of products, it was successfully tested in electrode renewal after deactivation in an aqueous solution. The increase in dimethyl formamide content led to more and more improved reproducibility of the currents of the outlined aromatic compounds. Naphthol isomers were assessed in the suppression of layer growth originating from the electrooxidation of another monomer phloroglucinol. Its simultaneous electrooxidation with naphthol monomers had a dramatic effect on layer morphology and it was found that instead of a coherent organic layer originating from the homopolymerization of phloroglucinol, the copolymerization with naphthols led to the development of more porous and rougher deposits. The suppressed electropolymerization thus increased sensitivity towards a chosen redox active compound, 4-methoxyphenol. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Anodic peak currents of 1-naphthol (<b>a</b>) and 2-naphthol (<b>b</b>) in different non-aqueous solvents (c = 20 mM, supporting electrolyte 50 mM TBAP), (<b>c</b>) is related to peak currents recorded in 1-pentanol, and (<b>d</b>) in CH<sub>2</sub>Cl<sub>2.</sub></p>
Full article ">Figure 1 Cont.
<p>Anodic peak currents of 1-naphthol (<b>a</b>) and 2-naphthol (<b>b</b>) in different non-aqueous solvents (c = 20 mM, supporting electrolyte 50 mM TBAP), (<b>c</b>) is related to peak currents recorded in 1-pentanol, and (<b>d</b>) in CH<sub>2</sub>Cl<sub>2.</sub></p>
Full article ">Figure 2
<p>Cyclic voltammograms for 1-naphthol (<b>a</b>) and 2-naphthol (<b>b</b>) in 1-propanol (<span class="html-italic">c</span> = 20 mM, <span class="html-italic">c</span>(TBAP) = 50 mM, in inset graphs: curves taken in allyl alcohol).</p>
Full article ">Figure 2 Cont.
<p>Cyclic voltammograms for 1-naphthol (<b>a</b>) and 2-naphthol (<b>b</b>) in 1-propanol (<span class="html-italic">c</span> = 20 mM, <span class="html-italic">c</span>(TBAP) = 50 mM, in inset graphs: curves taken in allyl alcohol).</p>
Full article ">Figure 3
<p>Ten successive cyclic voltammograms for 1-naphthol (<b>a</b>) and 2-naphthol (<b>b</b>) in mesityl oxide (inset graph: in methyl isobutyl ketone) (<span class="html-italic">c</span> = 20 mM, supporting electrolyte 50 mM TBAP).</p>
Full article ">Figure 4
<p>Scanning electron micrographs of layers after 10 successive voltammograms for 1-naphthol electropolymer deposited from MIBK (<b>a</b>), 2-naphthol from MIBK (<b>b</b>), 1-naphthol from MZO (<b>c</b>), and 2-naphthol from MZO (<b>d</b>) (c = 20 mM, supporting electrolyte 50 mM TBAP).</p>
Full article ">Figure 4 Cont.
<p>Scanning electron micrographs of layers after 10 successive voltammograms for 1-naphthol electropolymer deposited from MIBK (<b>a</b>), 2-naphthol from MIBK (<b>b</b>), 1-naphthol from MZO (<b>c</b>), and 2-naphthol from MZO (<b>d</b>) (c = 20 mM, supporting electrolyte 50 mM TBAP).</p>
Full article ">Figure 5
<p>Normalized peak currents for different <span class="html-italic">v</span>/<span class="html-italic">v</span>% dimethyl formamide in solutions mixed with water containing 1 mM 1-naphthol (<b>a</b>) and 1 mM 2-naphthol (<b>b</b>) (v = 0.1 V/s, supporting electrolyte 10 mM TBAP).</p>
Full article ">Figure 6
<p>Repeated voltammograms in aqueous solutions containing 1 mM 1-naphthol (<b>a</b>) and 2-naphthol (<b>b</b>) adjusted the pH to 7 with 0.05 M phosphate buffer (black curve: initial scan, red curve: scan after the first immersion and 1 min immersion in DMF, blue curve: scan after the second one and an additional 1 min immersion in DMF).</p>
Full article ">Figure 7
<p>Scanning electronic microscopic images of electrodeposited film from 50 mM solution of phloroglucinol prepared with dimethyl sulfoxide (<b>a</b>), in the presence of 50 mM 1-naphthol (<b>b</b>), and in the presence of 50 mM 2-naphthol (<b>c</b>).</p>
Full article ">Figure 7 Cont.
<p>Scanning electronic microscopic images of electrodeposited film from 50 mM solution of phloroglucinol prepared with dimethyl sulfoxide (<b>a</b>), in the presence of 50 mM 1-naphthol (<b>b</b>), and in the presence of 50 mM 2-naphthol (<b>c</b>).</p>
Full article ">
19 pages, 3151 KiB  
Article
Genetical and Biochemical Basis of Methane Monooxygenases of Methylosinus trichosporium OB3b in Response to Copper
by Dipayan Samanta, Tanvi Govil, Priya Saxena, Lee Krumholz, Venkataramana Gadhamshetty, Kian Mau Goh and Rajesh K. Sani
Methane 2024, 3(1), 103-121; https://doi.org/10.3390/methane3010007 - 20 Feb 2024
Cited by 1 | Viewed by 1208
Abstract
Over the past decade, copper (Cu) has been recognized as a crucial metal in the differential expression of soluble (sMMO) and particulate (pMMO) forms of methane monooxygenase (MMO) through a mechanism referred to as the “Cu switch”. In this study, we used Methylosinus [...] Read more.
Over the past decade, copper (Cu) has been recognized as a crucial metal in the differential expression of soluble (sMMO) and particulate (pMMO) forms of methane monooxygenase (MMO) through a mechanism referred to as the “Cu switch”. In this study, we used Methylosinus trichosporium OB3b as a model bacterium to investigate the range of Cu concentrations that trigger the expression of sMMO to pMMO and its effect on growth and methane oxidation. The Cu switch was found to be regulated within Cu concentrations from 3 to 5 µM, with a strict increase in the methane consumption rates from 3.09 to 3.85 µM occurring on the 6th day. Our findings indicate that there was a decrease in the fold changes in the expression of methanobactin (Mbn) synthesis gene (mbnA) with a higher Cu concentration, whereas the Ton-B siderophore receptor gene (mbnT) showed upregulation at all Cu concentrations. Furthermore, the upregulation of the di-heme enzyme at concentrations above 5 µM Cu may play a crucial role in the copper switch by increasing oxygen consumption; however, the role has yet not been elucidated. We developed a quantitative assay based on the naphthalene–Molisch principle to distinguish between the sMMO- and pMMO-expressing cells, which coincided with the regulation profile of the sMMO and pMMO genes. At 0 and 3 µM Cu, the naphthol concentration was higher (8.1 and 4.2 µM, respectively) and gradually decreased to 0 µM naphthol when pMMO was expressed and acted as the sole methane oxidizer at concentrations above 5 µM Cu. Using physical protein–protein interaction, we identified seven transporters, three cell wall biosynthesis or degradation proteins, Cu resistance operon proteins, and 18 hypothetical proteins that may be involved in Cu toxicity and homeostasis. These findings shed light on the key regulatory genes of the Cu switch that will have potential implications for bioremediation and biotechnology applications. Full article
(This article belongs to the Special Issue Trends in Methane-Based Biotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Growth statistics, (<b>b</b>) methane consumption, (<b>c</b>) oxygen consumption, and (<b>d</b>) methanol concentration of <span class="html-italic">M. trichosporium</span> OB3b using methane as substrate.</p>
Full article ">Figure 2
<p>Changes in the concentration of (<b>a</b>) Cu and (<b>b</b>) Fe in presence of different Cu concentrations.</p>
Full article ">Figure 3
<p>(<b>a</b>) Heat map showing the gene expressions with concentration distance variances and (<b>b</b>) naphthol concentration.</p>
Full article ">Figure 4
<p>The optimization and working principle of naphthalene–Molisch assay.</p>
Full article ">Figure 5
<p>(<b>a</b>) Gene expression profile of Mbn and (<b>b</b>) operon system of methanobactin in <span class="html-italic">M. trichosporium</span> OB3b.</p>
Full article ">Figure 6
<p>Physical protein–protein interactions between the (<b>a</b>) methanobactin (in red), (<b>b</b>) MMOs (in blue), (<b>c</b>) transporters (in green), (<b>d</b>) cell division (in yellow), and (<b>e</b>) hypothetical/domain-containing proteins (in purple). The circles represent the node associated with each gene (marked either with gene name or UniProt accession), and the solid connecting lines represent the edges (network between two genes).</p>
Full article ">
14 pages, 1066 KiB  
Article
Selenium-Containing (Hetero)Aryl Hybrids as Potential Antileishmanial Drug Candidates: In Vitro Screening against L. amazonensis
by Maria Helena Fermiano, Amarith Rodrigues das Neves, Fernanda da Silva, Manuella Salustiano Andrade Barros, Camila Barbosa Vieira, André L. Stein, Tiago Elias Allievi Frizon, Antonio Luiz Braga, Carla Cardozo Pinto de Arruda, Eduardo Benedetti Parisotto, Sumbal Saba, Jamal Rafique and Thalita Bachelli Riul
Biomedicines 2024, 12(1), 213; https://doi.org/10.3390/biomedicines12010213 - 17 Jan 2024
Cited by 1 | Viewed by 1276
Abstract
Leishmaniasis remains a significant global health concern, with current treatments relying on outdated drugs associated with high toxicity, lengthy administration, elevated costs, and drug resistance. Consequently, the urgent need for safer and more effective therapeutic options in leishmaniasis treatment persists. Previous research has [...] Read more.
Leishmaniasis remains a significant global health concern, with current treatments relying on outdated drugs associated with high toxicity, lengthy administration, elevated costs, and drug resistance. Consequently, the urgent need for safer and more effective therapeutic options in leishmaniasis treatment persists. Previous research has highlighted selenium compounds as promising candidates for innovative leishmaniasis therapy. In light of this, a library of 10 selenium-containing diverse compounds was designed and evaluated in this study. These compounds included selenium-substituted indole, coumarin, chromone, oxadiazole, imidazo[1,2-a]pyridine, Imidazo[2,1-b]thiazole, and oxazole, among others. These compounds were screened against Leishmania amazonensis promastigotes and intracellular amastigotes, and their cytotoxicity was assessed in peritoneal macrophages, NIH/3T3, and J774A.1 cells. Among the tested compounds, MRK-106 and MRK-108 displayed the highest potency against L. amazonensis promastigotes with reduced cytotoxicity. Notably, MRK-106 and MRK-108 exhibited IC50 values of 3.97 µM and 4.23 µM, respectively, and most of the tested compounds showed low cytotoxicity in host cells (CC50 > 200 µM). Also, compounds MRK-107 and MRK-113 showed activity against intracellular amastigotes (IC50 18.31 and 15.93 µM and SI 12.55 and 10.92, respectively). In conclusion, the identified selenium-containing compounds hold potential structures as antileishmanial drug candidates to be further explored in subsequent studies. These findings represent a significant step toward the development of safer and more effective therapies for leishmaniasis, addressing the pressing need for novel and improved treatments. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of selenium-substituted (hetero)aryl hybrids [<a href="#B34-biomedicines-12-00213" class="html-bibr">34</a>,<a href="#B35-biomedicines-12-00213" class="html-bibr">35</a>,<a href="#B36-biomedicines-12-00213" class="html-bibr">36</a>,<a href="#B37-biomedicines-12-00213" class="html-bibr">37</a>,<a href="#B38-biomedicines-12-00213" class="html-bibr">38</a>,<a href="#B39-biomedicines-12-00213" class="html-bibr">39</a>,<a href="#B40-biomedicines-12-00213" class="html-bibr">40</a>] used in this study.</p>
Full article ">Figure 2
<p>Antileishmanial activity of synthesized selenium compounds <b>MRK-105</b>, <b>MRK-106</b>, <b>MRK-107</b>, <b>MRK-108</b>, and <b>MRK-113</b> against promastigotes (<b>A</b>–<b>E</b>, upper graphs) and intracellular amastigotes (<b>F</b>–<b>J</b>, lower graphs) of <span class="html-italic">L. amazonensis</span>. * and *** mean <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.001, respectively (ANOVA and Tukey’s post-test).</p>
Full article ">Figure 3
<p>Cytotoxicity of synthesized selenium compounds <b>MRK-105</b>, <b>MRK-106</b>, <b>MRK-107</b>, <b>MRK-108</b>, and <b>MRK-113</b> against NIH/3T3 fibroblasts (<b>A</b>–<b>E</b>, upper graphs), J774A.1 macrophages (<b>F</b>–<b>J</b>, middle graphs) and murine peritoneal macrophages (<b>K</b>–<b>O</b>, lower graphs). *** mean <span class="html-italic">p</span> &lt; 0.001 (ANOVA and Tukey´s post-test).</p>
Full article ">
Back to TopTop