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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Search Results (306)

Search Parameters:
Keywords = green chemistry approach

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 1375 KiB  
Article
Green Chemistry Within the Circular Bioeconomy to Harness Chestnut Burr Extract’s Synergistic Antimicrobial Activity Against Helicobacter pylori
by Maria Lucia Schiavone, Roberta Barletta, Alfonso Trezza, Michela Geminiani, Lia Millucci, Natale Figura and Annalisa Santucci
Molecules 2025, 30(2), 324; https://doi.org/10.3390/molecules30020324 - 15 Jan 2025
Viewed by 24
Abstract
Green chemistry principles are pivotal in driving sustainable and innovative solutions to global health challenges. This study explores a hydroalcoholic extract from Castanea sativa (chestnut) burrs, an underutilized natural resource, as a potent source of antimicrobial compounds against Helicobacter pylori (H. pylori [...] Read more.
Green chemistry principles are pivotal in driving sustainable and innovative solutions to global health challenges. This study explores a hydroalcoholic extract from Castanea sativa (chestnut) burrs, an underutilized natural resource, as a potent source of antimicrobial compounds against Helicobacter pylori (H. pylori). The extract demonstrated significant bactericidal activity, synergizing effectively with clarithromycin and showing additive effects with metronidazole. Remarkably, combining the extract with clarithromycin and sub-inhibitory concentrations of pantoprazole reduced clarithromycin’s Minimum Bactericidal Concentration (MBC) to just 1.56% of its original value. Mechanistic studies suggest that the extract’s polyphenolic compounds compromise bacterial membrane integrity, enhancing antibiotic uptake, while pantoprazole disrupts bacterial ATPase activity. This research highlights the critical role of natural product extraction within the framework of green chemistry, offering a sustainable and environmentally friendly alternative to synthetic antimicrobials. By harnessing bioactive compounds from plant sources, this approach addresses the pressing issue of antibiotic resistance while promoting the responsible use of natural resources. The findings underscore the transformative potential of green chemistry in developing effective, eco-conscious antimicrobial therapies that align with global sustainability goals. Full article
(This article belongs to the Special Issue Green Chemistry Approaches to Analysis and Environmental Remediation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>MBC determination of <span class="html-italic">C. sativa</span> shell extract alone. (<b>A</b>) Bacterial agar spot assay; (<b>B</b>) Corresponding microplate layout. Serial two-fold dilutions of <span class="html-italic">C. sativa</span> extract, at concentrations ranging from 25 mg/mL to 0.2 mg/mL, were carried out across the first column. The MBC value was 3.125 mg/mL (well E1, pointed to by a black arrow and circled in black).</p>
Full article ">Figure 2
<p>Determination of the Minimum Bactericidal Concentration (MBC) of <span class="html-italic">C. sativa</span> shell extract, clarithromycin, and their combination. (<b>A</b>) Bacterial agar spot assay; (<b>B</b>) Corresponding microplate layout. Colored wells indicate bacterial growth, whereas clear wells show no growth. The first and second column contain serial two-fold dilutions of the extract and clarithromycin alone, respectively, while the third and fourth columns represent negative controls. Wells A–H of columns 5–8 include both the extract and the antibiotic, with the extract diluted along the columns and the antibiotic diluted across the rows. Well E1, corresponding to the MBC value of the extract alone (3.125 mg/mL), is indicated by a black arrow on the agar plate and circled in black. Well F2 represents the MBC value of clarithromycin alone (0.25 mg/mL), indicated by a blue arrow on the plate and circled in blue, while well F9 denotes the MBC value of the combined extract/antibiotic, with its corresponding spot on the agar plate indicated by a red arrow and circled in red.</p>
Full article ">Figure 3
<p>Synergistic effects through checkerboard analysis, showing a substantial reduction in the MBC of clarithromycin when combined with the extract, indicating improved antimicrobial efficacy. Determination of the synergistic effect between <span class="html-italic">C. sativa</span> extract and the antibiotic using the checkerboard method. The extract was serially diluted across the columns, while the antibiotic was diluted across the rows. Column 1 (black), 2 (blue), 5 to 8 (red) contain the extract alone, the antibiotic alone, and the extract and antibiotic combinations, respectively, while columns 3 and 4 represent negative controls.</p>
Full article ">
27 pages, 4058 KiB  
Review
NMR and MD Simulations of Non-Ionic Surfactants
by Gerd Buntkowsky and Markus Hoffmann
Molecules 2025, 30(2), 309; https://doi.org/10.3390/molecules30020309 - 14 Jan 2025
Viewed by 188
Abstract
Non-ionic surfactants are an important solvent in the field of green chemistry with tremendous application potential. Understanding their phase properties in bulk or in confined environments is of high commercial value. In recent years, the combination of molecular dynamics (MD) simulations with multinuclear [...] Read more.
Non-ionic surfactants are an important solvent in the field of green chemistry with tremendous application potential. Understanding their phase properties in bulk or in confined environments is of high commercial value. In recent years, the combination of molecular dynamics (MD) simulations with multinuclear solid-state NMR spectroscopy and calorimetric techniques has evolved into the most powerful tool for their investigation. Showing recent examples from our groups, the present review demonstrates the power and versatility of this approach, which can handle both small model-surfactants like octanol and large technical surfactants like technical polyethylene glycol (PEG) mixtures and reveals otherwise unobtainable knowledge about their phase behavior and the underlying molecular arrangements. Full article
(This article belongs to the Special Issue Molecular Simulation in Interface and Surfactant—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the common types of surfactants (<b>left</b>) and scheme of micelle formation when the surfactant concentration, c, is greater than the critical micelle concentration, CMC.</p>
Full article ">Figure 2
<p>Chemical structures of Surfactants: (<b>a</b>) PEG, (<b>b</b>) C<sub>m</sub>E<sub>n</sub>, (<b>c</b>) C<sub>m</sub>E<sub>n</sub>P<sub>q</sub>, and (<b>d</b>) Triton X-100. For the surfactants, m indicates the number of carbon atoms in the alkyl chain (C), while n and <span class="html-italic">q</span> are the repetition units of ethylene oxide (E) and propylene glycol (P), respectively.</p>
Full article ">Figure 3
<p>Low-temperature <sup>1</sup>H/<sup>29</sup>Si CP-MAS FSLG HETCOR obtained with 9 ms contact time of an 80:20 mol% mixture 1-octanol:water mixture, depicted by Kumari et al. [<a href="#B52-molecules-30-00309" class="html-bibr">52</a>]. Reprinted with permission from Kumari et al. [<a href="#B52-molecules-30-00309" class="html-bibr">52</a>]. Copyright 2018 American Chemical Society.</p>
Full article ">Figure 4
<p>Selected temperature-dependent <sup>2</sup>H MAS NMR spectra (left) of 1-octanol-d17 confined in mesoporous SBA-15 (illustrated, middle). All spectra are normalized to an equal height. Changes in the line shapes reflect the motional dynamics affected by the liquid-to-solid phase transitions whose distribution of activation energies, g<sub>E</sub> (right panel), become visibly broader under confinement (blue) compared to bulk (red) for the data obtained from MAS conditions (solid line) compared to static conditions (dashed line) (see ref. [<a href="#B100-molecules-30-00309" class="html-bibr">100</a>] for details). Reprinted with permission from Döller et al. [<a href="#B100-molecules-30-00309" class="html-bibr">100</a>] Copyright 2021 American Chemical Society.</p>
Full article ">Figure 5
<p>Hydrogen bonding of octanol and the silica surface orients the octanol molecules (<b>left</b>). By a combination of hydrophilic and hydrophobic interactions, this ordering is transmitted inside the pores (<b>right</b>), creating a cylindrical double-layer structure (for details, see Kumari et al. [<a href="#B52-molecules-30-00309" class="html-bibr">52</a>]). Reprinted with permission from Kumari et al. [<a href="#B52-molecules-30-00309" class="html-bibr">52</a>] and Döller et al. [<a href="#B100-molecules-30-00309" class="html-bibr">100</a>]. Copyright 2018 and 2021 American Chemical Society.</p>
Full article ">Figure 6
<p>Solid-state NMR and MD simulations of water -isobutyric acid mixtures (56 wt% iBA) confined in SBA-15 Upper row: 2D <sup>1</sup>H-<sup>29</sup>Si FSLG HETCOR at 9.4Tesla, 110 K and 8 kHz MAS. The spectrum measured with a long contact time of 3 ms (<b>left</b>) reveals cross-peaks from silica to carboxyl, hydroxyl, and aliphatic protons. At short contact times (<b>right</b>) of 0.5 ms, only cross peaks to aliphatic and hydroxyl-protons are visible. Lower row: density profiles for iBA and water (<b>left</b>) and hydrogen atom density (<b>right</b>). The pore center is at 0 Å, and the pore wall is at 25 Å. Reprinted with permission from Harrach et al. [<a href="#B92-molecules-30-00309" class="html-bibr">92</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 7
<p>Temperature-dependent <sup>1</sup>H MAS spectra [<a href="#B93-molecules-30-00309" class="html-bibr">93</a>] of (<b>a</b>) neat C<sub>10</sub>E<sub>6</sub>P<sub>2</sub> and (<b>b</b>) C<sub>10</sub>E<sub>6</sub>P<sub>2</sub> confined in mesoporous silica material measured at 5 kHz spinning in the range between 106.9 K and 253.6 K.</p>
Full article ">Figure 8
<p>Illustration of (<b>a</b>) SBA-5%, (<b>b</b>) SBA-20% impregnated with a hydrophilic polarizing agent (PA) in C<sub>10</sub>E<sub>6</sub>. The larger APTES coverage in SBA-20% renders the pore surface more nonpolar than in SBA-5%, which supports the formation of a structured bilayer arrangement of C<sub>10</sub>E<sub>6</sub> within the pore. The gray blocks represent the silica pore wall, and the red oval area represents the region around the polarizing agent where nuclei cannot be detected by NMR. The green oval represents the region to which nuclei may receive polarization directly from the polarizing agent. Reprinted with permission from Hoffmann et al. [<a href="#B113-molecules-30-00309" class="html-bibr">113</a>]. Copyright 2020 American Chemical Society.</p>
Full article ">Figure 9
<p>Schematic illustration of (<b>a</b>) the hydrophilic surfactants (E5 and PEG 200) and (<b>b</b>) the amphiphilic surfactants (C<sub>10</sub>E<sub>6</sub> and Triton) oriented in the pores of the mesoporous silica host material. E<sub>n</sub> represents the polyethylene glycol units, C<sub>x</sub>H<sub>y</sub> represents the lipophilic moiety of the amphiphilic surfactants, and PA represents the polarizing agent AMUPol (not to scale). Reprinted with permission from Döller et al. [<a href="#B124-molecules-30-00309" class="html-bibr">124</a>]. Copyright 2023 American Chemical Society.</p>
Full article ">Figure 10
<p>Potential function <span class="html-italic">U</span><sub>total</sub> consisting of the bonded interactions (<b>a</b>–<b>d</b>) and nonbonded interactions (<b>e</b>,<b>f</b>) illustrated schematically by example chemical systems. Bonded interactions describe chemical bond (<b>a</b>) and angle vibrations (<b>b</b>) by harmonic oscillator functions with force constants <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>k</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>k</mi> </mrow> <mrow> <mi>θ</mi> </mrow> </msub> </mrow> </semantics></math> and equilibrium bond lengths/angles <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>b</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math>, respectively. Proper dihedrals (<b>c</b>) may be represented by the Ryckaert–Bellemans potential with torsional energy barrier coefficients <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> <mo>,</mo> <mtext> </mtext> <mn>1</mn> <mo>,</mo> <mtext> </mtext> <mn>2</mn> <mo>,</mo> <mtext> </mtext> <mn>3</mn> <mo>,</mo> <mtext> </mtext> <mn>4</mn> <mo>,</mo> <mtext> </mtext> <mn>5</mn> </mrow> </semantics></math> and torsional angle <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math>. Improper dihedrals (<b>d</b>) may be represented by a periodic harmonic cosine potential, where <math display="inline"><semantics> <mrow> <mi>ξ</mi> </mrow> </semantics></math> is the improper torsional angle at any point in time, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> is the phase shift, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>k</mi> </mrow> <mrow> <mi>ξ</mi> </mrow> </msub> </mrow> </semantics></math> is the amplitude of the potential energy curve, and <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> represents the multiplicity/periodicity of the cosine function. The nonbonded interactions consist of two terms. The Lennard Jones potential represents the London dispersion forces that are present in all chemicals, even noble gases such as argon (<b>e</b>), with contact distance, <span class="html-italic">σ</span>, and well-depth, <span class="html-italic">ε</span>, between two atoms at distance <span class="html-italic">r</span> being the interaction parameters. The Coulomb potential function represents the interactions between charges, <span class="html-italic">q</span>, where <span class="html-italic">ε</span><sub>0</sub> is the permittivity of vacuum (<b>f</b>).</p>
Full article ">Figure 11
<p><b>Left panel</b>: Adjusted ratio of intra-over intermolecular hydrogen bonds between hydroxy hydrogen and hydroxy oxygen for each oligomer with itself in PEG200 obtained from MD simulations at 328 K using the OPLS force field (circles) and modified OPLS force field (squares). The ratio numbers were adjusted by the number of possible intra- and intermolecular hydrogen bonds, the respective oligomer mole fraction, and a scaling factor depending on the number of oligomer components. Reprinted with permission from Hoffmann et al. [<a href="#B158-molecules-30-00309" class="html-bibr">158</a>]. Copyright 2023 American Chemical Society. <b>Right panel</b>: Radial distribution functions obtained with the TIP4P/2005 and OPLS forcefields of water oxygen with hydroxy (black) and with ether oxygen (red) of hexaethylene glycol, each from four different water mass fraction of 0.001, 0.005, 0.010, and 0.020. Aside from different noise level, the respective radial distribution functions are indistinguishable.</p>
Full article ">
20 pages, 1493 KiB  
Article
Green Extraction of Bioactives from Curcuma longa Using Natural Deep Eutectic Solvents: Unlocking Antioxidative, Antimicrobial, Antidiabetic, and Skin Depigmentation Potentials
by Jelena Jovanović, Marko Jović, Jelena Trifković, Katarina Smiljanić, Uroš Gašić, Maja Krstić Ristivojević and Petar Ristivojević
Plants 2025, 14(2), 163; https://doi.org/10.3390/plants14020163 - 8 Jan 2025
Viewed by 441
Abstract
This study evaluates the efficiency of 20 Natural Deep Eutectic Solvents (NADES) formulations for extracting curcuminoids and other bioactive compounds from turmeric and emphasize their ability to preserve and enhance antioxidant, antimicrobial, antidiabetic, and skin depigmentation effects. The NADES formulations, prepared using choline [...] Read more.
This study evaluates the efficiency of 20 Natural Deep Eutectic Solvents (NADES) formulations for extracting curcuminoids and other bioactive compounds from turmeric and emphasize their ability to preserve and enhance antioxidant, antimicrobial, antidiabetic, and skin depigmentation effects. The NADES formulations, prepared using choline chloride (ChCl) combined with sugars, carboxylic acids, glycerol, amino acids, urea, polyols, and betaine, were assessed for their extraction efficiency based on the total phenolic content and curcumin concentration. Fourier transform infrared spectroscopy was employed to characterize the synthesized NADES and confirm their chemical composition. Bioactivity evaluations included antioxidant assays (ABTS and DPPH), antidiabetic tests (α-amylase inhibition), antimicrobial assays, and skin depigmentation (tyrosinase inhibition). The results demonstrated that NADES significantly enhanced the extraction efficiency and bioactive properties of turmeric extracts compared to water as a conventional green solvent. NADES 18 (ChCl/1,2-propanediol/water 1:1:1) and NADES 19 (glycerol/betaine/water 1:1:3) exhibited the highest extraction yields, with curcumin concentrations of 30.73 ± 1.96 mg/g and 31.70 ± 2.02 mg/g, respectively, outperforming water (26.91 ± 1.72 mg/g), while NADES 17 (ChCl/1,2-propanediol/water 0.5:3:0.5:5) and NADES 20 (glycerol/lysine/water 1:1:3) exhibited the most potent antioxidant activity. Furthermore, NADES 14 (ChCl/lactic acid/water 1:2:5) demonstrated the strongest tyrosinase inhibition (98.7%), supporting its potential for skin-brightening applications, including notable α-amylase inhibition exceeding 90%. This study aligns with the principles of green chemistry, as NADES are effective and sustainable solvents for natural product extraction. The presenting benefits of improved extraction efficiency and enhanced bioactivities position NADES as a promising and eco-friendly approach for developing efficient bioactive compound extraction methodologies. Full article
Show Figures

Figure 1

Figure 1
<p>Tyrosinase enzyme test inhibition and HaCaT cells’ survival under the action of 20 NADES and conventional turmeric extracts. (<b>A</b>) percentage of tyrosinase inhibition with Kojic acid as reference standard. (<b>B</b>) Cytotoxicity test reflecting survival of spontaneously immortalized, human keratinocyte cell line (HaCaT). Distinct letters above the columns indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05) as determined by Tukey’s multiple comparisons test. The dashed line marks 100% viability of untreated control cells.</p>
Full article ">Figure 2
<p>Percentages of inhibition for α-amylase. Distinct letters above the columns indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05) as determined by Tukey’s multiple comparisons test.</p>
Full article ">Figure 3
<p>Curcumin content in NADES and water extracts. Distinct letters above the columns indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05) as determined by Tukey’s multiple comparisons test.</p>
Full article ">
39 pages, 14231 KiB  
Review
Asymmetric Transformations of Levulinic Acid to γ-Valerolactone and 5-Methylpyrrolidin-2-one Derivatives: Chiral Compounds with Biological Potential
by Elżbieta Łastawiecka and Katarzyna Szwaczko
Symmetry 2025, 17(1), 82; https://doi.org/10.3390/sym17010082 - 7 Jan 2025
Viewed by 421
Abstract
Levulinic acid is a key platform molecule derived from biomass and readily available from natural sources, making it an attractive starting material for the synthesis of high-value chiral compounds. Among them, γ-valerolactone and 5-methylpyrrolidin-2-one derivatives are notable for their widespread occurrence and [...] Read more.
Levulinic acid is a key platform molecule derived from biomass and readily available from natural sources, making it an attractive starting material for the synthesis of high-value chiral compounds. Among them, γ-valerolactone and 5-methylpyrrolidin-2-one derivatives are notable for their widespread occurrence and biological importance. This review paper highlights the importance of γ-valerolactone and 5-methylpyrrolidin-2-one derivatives as frameworks found in biologically active compounds and pharmaceuticals. It focuses on the asymmetric synthesis of these chiral building blocks from levulinic acid, highlighting recent advances in catalytic transformations that allow for efficient and selective transformations. The potential applications of these chiral molecules in medicine and industry underscore the importance of developing sustainable and scalable processes for their production. This review also examines future directions in the field, given the growing demand for green chemistry approaches and the increasing importance of chiral molecules in drug development. Full article
(This article belongs to the Section Chemistry: Symmetry/Asymmetry)
Show Figures

Figure 1

Figure 1
<p>Transformations of levulinic acid to fine chemicals.</p>
Full article ">Figure 2
<p>Chiral phosphine ligands for the synthesis of optically active GVL.</p>
Full article ">Figure 3
<p>Represented biologically active <span class="html-italic">γ</span>-valerolactones.</p>
Full article ">Figure 4
<p>Represented biologically active <span class="html-italic">γ</span>-lactams.</p>
Full article ">Scheme 1
<p>Preparation of levulinic acid from biomass.</p>
Full article ">Scheme 2
<p>Conversion of monosaccharides to levulinic acid LA; (<b>a</b>) hexose and (<b>b</b>) pentose routes towards LA.</p>
Full article ">Scheme 3
<p>Routes based on the catalytic hydrogenation of LA to obtain GVL.</p>
Full article ">Scheme 4
<p>Pathways for the synthesis of 5-MPs by reductive amination or amidation of LA and levulinates.</p>
Full article ">Scheme 5
<p>Transformation of levulinic acid and α-angelica lactone to 5-methylpyrrolidin-2-one using a heterogeneous nickel catalyst.</p>
Full article ">Scheme 6
<p>Bio-based click reaction for the synthesis of dihydropyridazinone as a platform for 5-methylpyrrolidin-2-one.</p>
Full article ">Scheme 7
<p>Synthesis of chiral <span class="html-italic">γ</span>-lactones with [Ru-tetraMe-BITIOP](OTf)<sub>2</sub>.</p>
Full article ">Scheme 8
<p>RuCl<sub>3</sub>–BINAP–HCl catalyst in the reduction of alkyl levulinate.</p>
Full article ">Scheme 9
<p>Application of Ir-SIPHOX-type phosphine complex in the reduction of <span class="html-italic">γ</span>-ketoacids.</p>
Full article ">Scheme 10
<p>Optimizing ligand selection to enhance enantioselectivity.</p>
Full article ">Scheme 11
<p>Ru−catalyzed direct reductive amination of levulinic acid and methyl levulinate to the chiral pyrrolidinone.</p>
Full article ">Scheme 12
<p>Possible paths of asymmetric reductive amination of LA.</p>
Full article ">Scheme 13
<p>Biosynthesis of enantioenriched esters and lactones with ADH or LBADH.</p>
Full article ">Scheme 14
<p>Enzymatic method for the synthesis of pure (<span class="html-italic">S</span>)-GVL.</p>
Full article ">Scheme 15
<p>General transformation of <span class="html-italic">γ</span>- and <span class="html-italic">δ</span>-keto esters into optically active lactams mediated by transaminases.</p>
Full article ">Scheme 16
<p>Transaminase-catalyzed transformation of levulinic ester into chiral 5-methylpyrrolidin-2-one; <span class="html-italic"><sup>(</sup></span><sup>a)</sup> isolated yield.</p>
Full article ">Scheme 17
<p>Asymmetric synthesis of enantiomerically pure (<span class="html-italic">R</span>)−4−aminovaleric acid by TtherAmDH.</p>
Full article ">Scheme 18
<p>Asymmetric synthesis of enantiomerically pure (<span class="html-italic">S</span>)-4-aminovaleric acid by HBV-ω-TA.</p>
Full article ">Scheme 19
<p>Synthesis of (<span class="html-italic">S</span>)−4−aminopentanoic acid catalyzed by AmDH−4 with a FDH cofactor regeneration system.</p>
Full article ">Scheme 20
<p>Synthesis steps involved in the catalytic process for producing (<span class="html-italic">R</span>)-5-methylpyrrolidin-2-one.</p>
Full article ">
14 pages, 6588 KiB  
Article
Sustainable Corrosion Inhibitors from Pharmaceutical Wastes: Advancing Energy-Efficient Chemistry with Green Solutions
by Narasimha Raghavendra, Sharanappa Chapi, Murugendrappa M. V., Małgorzata Pawlak and Mohammad Reza Saeb
Energies 2025, 18(2), 224; https://doi.org/10.3390/en18020224 - 7 Jan 2025
Viewed by 437
Abstract
Pharmaceutical waste is a type of bio-waste inevitably generated by the pharmaceutical industry, often due to regulatory changes, product deterioration, or expiration. However, their collection and valorization can be approached from a sustainable perspective, offering potential energy-efficient solutions. In this work, the expired [...] Read more.
Pharmaceutical waste is a type of bio-waste inevitably generated by the pharmaceutical industry, often due to regulatory changes, product deterioration, or expiration. However, their collection and valorization can be approached from a sustainable perspective, offering potential energy-efficient solutions. In this work, the expired Eslicarbazepine acetate drug (ESLD) was utilized as a sustainable anticorrosive agent against mild steel in a 3 M HCl wash solution. Experimental tests combined with theoretical Density Functional Theory (DFT) and Monte Carlo (MC) simulations revealed the corrosion inhibition potential of ESLD. The gasometrical results revealed a high inhibition efficiency rate of 98% upon increases in concentration of expired ESLD from 0.25 to 1.00 mg·L−1, whereas hydrogen gas evolution decreased to 0.7 mL. An impedance investigation evidenced the pivotal role of charge transfer in reducing the disintegration process. As per DFT computations and MC simulation, electron-rich elements in the expired ESLD were key in controlling the dissolution through the adsorption process. Contact angle studies revealed that the increment in the contact angle from 61° to 80° in the presence of expired ESLD validates the chemical, electrochemical, and computational results. This approach not only mitigates pharmaceutical pollution, but also exemplifies the integration of green chemistry principles into corrosion protection, contributing to energy-efficient and sustainable industrial practices. Full article
(This article belongs to the Section B: Energy and Environment)
Show Figures

Figure 1

Figure 1
<p>Interactive chemical structure model (Ball and Stick) and chemical structure depiction of ESLD.</p>
Full article ">Figure 2
<p>Tafel plots with and without an inhibitor for MS in 3 M HCl solution.</p>
Full article ">Figure 3
<p>Nyquist curves for the MS exposed to the blank solution and acidic electrolytes with four distinct expired ESLD concentrations and equivalent circuits used in the present investigation (<a href="#energies-18-00224-f003" class="html-fig">Figure 3</a>).</p>
Full article ">Figure 4
<p>The relationship between the concentration of inhibitor and protection efficiency at different exposure periods.</p>
Full article ">Figure 5
<p>SEM images of (<b>a</b>) without and (<b>b</b>) with 1 mg/L of expired ESL drug at room temperature.</p>
Full article ">Figure 6
<p>Water contact angle measurement. (<b>a</b>) Bare and (<b>b</b>) inhibited system.</p>
Full article ">Figure 7
<p>DFT results: (<b>a</b>) HOMO, (<b>b</b>) LUMO, and (<b>c</b>) density mapping of an expired drug.</p>
Full article ">Figure 8
<p>Images of the different views of the adsorption of expired ESLD molecules on the Fe surface in an aqueous environment.</p>
Full article ">Figure 8 Cont.
<p>Images of the different views of the adsorption of expired ESLD molecules on the Fe surface in an aqueous environment.</p>
Full article ">Figure 9
<p>Diagram interpretation of the adsorption of expired ESLD on MS in acidic environments.</p>
Full article ">
8 pages, 908 KiB  
Communication
Efficient Synthesis of cis,cis-Muconic Acid by Catechol Oxidation of Ozone in the Presence of a Base
by Kohtaro Katayama, Hiroki Hotta and Yoshio Tsujino
Molecules 2025, 30(1), 201; https://doi.org/10.3390/molecules30010201 - 6 Jan 2025
Viewed by 518
Abstract
Muconic acid, a crucial precursor in synthesizing materials like PET bottles and nylon, is pivotal for the anticipated growth in the textiles and plastics industries. This study presents a novel chemical synthesis route for cis,cis-muconic acid (ccMA) using catechol. Biochemical [...] Read more.
Muconic acid, a crucial precursor in synthesizing materials like PET bottles and nylon, is pivotal for the anticipated growth in the textiles and plastics industries. This study presents a novel chemical synthesis route for cis,cis-muconic acid (ccMA) using catechol. Biochemical methods face scale-up challenges due to microorganism sensitivity and complex extraction processes, while chemical methods involve environmentally harmful substances and have low yields. Our research introduces a method that enhances ccMA yield to 56% by employing ozonation in the presence of an alkali, significantly simplifying the synthesis process. This one-step synthesis reduces reagent use and labor, aligns with green chemistry principles, and avoids using toxic chemicals. The methodology, involving the low-temperature ozonation of catechol with base addition, reduces ccMA degradation and improves yield, as confirmed by an HPLC analysis and replicated experiments. This promising approach could lead to sustainable industrial synthesis of muconic acid derivatives. Further investigations will focus on refining this method for larger-scale applications and testing its economic viability, aiming to optimize conditions for maximum efficiency and yield. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Scheme 1
<p>(<b>a</b>) Synthesis of <span class="html-italic">cis</span>,<span class="html-italic">cis</span>-muconic acid (ccMA) by oxidative cleavage of catechol with peracids [<a href="#B15-molecules-30-00201" class="html-bibr">15</a>], (<b>b</b>) synthesis of ccMA from catechol oxidation by ozone [<a href="#B23-molecules-30-00201" class="html-bibr">23</a>], and (<b>c</b>) synthesis of ccMA from catechol oxidation by ozone in the presence of an alkali.</p>
Full article ">Scheme 2
<p>Reaction of ccMA with ozone in alcohol.</p>
Full article ">Scheme 3
<p>Reaction of muconic acid in alcohol with ozone.</p>
Full article ">Scheme 4
<p>Preparation of ccMA.</p>
Full article ">
21 pages, 3141 KiB  
Article
Biorefining Brazilian Green Propolis: An Eco-Friendly Approach Based on a Sequential High-Pressure Extraction for Recovering High-Added-Value Compounds
by Guilherme Dallarmi Sorita, Wilson Daniel Caicedo Chacon, Monique Martins Strieder, Camilo Rodriguez-García, Alcilene Monteiro Fritz, Silvani Verruck, Germán Ayala Valencia and José A. Mendiola
Molecules 2025, 30(1), 189; https://doi.org/10.3390/molecules30010189 - 6 Jan 2025
Viewed by 470
Abstract
Propolis is a valuable natural resource for extracting various beneficial compounds. This study explores a sustainable extraction approach for Brazilian green propolis. First, supercritical fluid extraction (SFE) process parameters were optimized (co-solvent: 21.11% v/v CPME, and temperature: 60 °C) to maximize [...] Read more.
Propolis is a valuable natural resource for extracting various beneficial compounds. This study explores a sustainable extraction approach for Brazilian green propolis. First, supercritical fluid extraction (SFE) process parameters were optimized (co-solvent: 21.11% v/v CPME, and temperature: 60 °C) to maximize yield, total phenolic content (TPC), antioxidant capacity, and LOX (lipoxygenase) inhibitory activity. GC–MS analysis identified 40 metabolites in SFE extracts, including fatty acids, terpenoids, phenolics, and sterols. After selecting the optimum SFE process parameters, a sequential high-pressure extraction (HPE) approach was developed, comprising SFE, pressurized liquid extraction (PLE) with EtOH/H2O, and subcritical water extraction (SWE). This process was compared to a similar sequential extraction using low-pressure extractions (LPE) with a Soxhlet extractor. The HPE process achieved a significantly higher overall yield (80.86%) than LPE (71.43%). SFE showed higher selectivity, resulting in a lower carbohydrate content in the non-polar fraction, and PLE extracted nearly twice the protein amount of LPE–2. Despite the HPE selectivity, LPE extracts exhibited better acetylcholinesterase (AChE), butyrylcholinesterase (BChE), and LOX inhibition, demonstrating that the neuroprotective and anti-inflammatory activity of the extracts may be associated with a symbiosis of a set of compounds. Finally, a comprehensive greenness assessment revealed that the HPE process proved more sustainable and aligned with green chemistry principles than the LPE method. Full article
(This article belongs to the Special Issue Bioactive Molecules in Foods: From Sources to Functional Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Response surfaces of SFE of bioactive compounds from propolis. Studied factors: temperature and co-solvent percentage on the (<b>A</b>) extraction yield, (<b>B</b>) TPC, (<b>C</b>) ABTS, and (<b>D</b>) DPPH. Extraction pressure 20 MPa and flow 4 mL min<sup>−1</sup> constant in all experiments.</p>
Full article ">Figure 2
<p>Relative abundance (%) of phenolic compounds, fatty acids, terpenoids, and sterols of SFE extracts, optimum point, and SOX-Hex determined by GC–MS. Legend: Fatty acids: sum of the relative abundance (%) of hexadecanoic acid–palmitic acid, octadecanoic acid–stearic acid, 9,12-Octadecadienoic acid–linoleic acid, alpha-linolenic acid, alpha-linolenic acid isomer, alpha-linolenic acid isomer 2, 9,12-Octadecadienoic acid isomer 2, 11-Eicosenoic acid–omega 9, pentanoic acid, pentanoic acid–valeric acid isomer, hexadecanoic acid–monopalmitin, 5-Eicosene, (E)-, octadecanoic acid–Stearic acid, tetracosanoic acid–lignoceric acid, tetracosanoic acid isomer, cis-5,8,11-eicosatrienoic acid, and decanoic acid. Terpenoids: sum of the relative abundance (%) of trans-Caryophyllene, trans-Caryophyllene isomer, phenol, 2,6-di-tert-butylphenol, alpha-Copaene, nerolidol isomer, (+) spathulenol, (−)-Caryophyllene oxide, trans–trans-farnesol, pseduosarsasapogenin-5,20-dien, trans, trans-farnesol isomer, pentitol, farnesol, myrtenol, tetrahydro linalool, and globulol. Sterols: relative abundance (%) of ergost-25-ene-3,5,6,12-tetrol). Phenolic compounds: sum of the relative abundance (%) of benzenepropanoic acid, hydrocinnamic acid, cinnamic acid, 1,3-Benzenedicarboxylic acid, and trans-Aconitic acid. OP: optimized SFE extract. Means followed by the different letters in the bars indicate a significant statistical difference by Tukey’s test (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 3
<p>Fractions (<b>A</b>) yield (%), (<b>B</b>) total carbohydrate and protein content, (<b>C</b>) total phenolic content (TPC) and total flavonoid content (TFC), (<b>D</b>) antioxidant capacity (ABTS and DPPH methods), (<b>E</b>) AChE and BChE inhibitory activities, and (<b>F</b>) LOX inhibitory activity. Fractions obtained from green propolis by high-pressure extraction (HPE) and low-pressure extraction (LPE) using the biorefinery approach. See extraction conditions in <a href="#molecules-30-00189-f004" class="html-fig">Figure 4</a>. Means followed by the different letters in the bars indicate a significant statistical difference by Tukey’s test (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 4
<p>Sequential extraction approach proposed to recover high-added-value compounds from Brazilian green propolis.</p>
Full article ">Figure 5
<p>Comparative environmental impact assessment of (<b>a</b>) LPE and (<b>b</b>) HPE biomass processing: analysis of biomass (I), transport (II), pre-treatment (III), solvent (IV), scale-up (V), purification (VI), yield (VII), post-treatment (VIII), energy (IX), application (X), repurposing (XI), and waste (XII). Indications of a good score (green), areas requiring attention (yellow), and poor score (red).</p>
Full article ">
22 pages, 1658 KiB  
Article
Optimization of a Pressurized Extraction Process Based on a Ternary Solvent System for the Recovery of Neuroprotective Compounds from Eucalyptus marginata Leaves
by Soumaya Hasni, Hajer Riguene, Jose A. Mendiola, Elena Ibáñez, Lidia Montero, Gloria Domínguez-Rodríguez, Hanene Ghazghazi, Ghayth Rigane and Ridha Ben Salem
Int. J. Mol. Sci. 2025, 26(1), 94; https://doi.org/10.3390/ijms26010094 - 26 Dec 2024
Viewed by 319
Abstract
Green chemistry focuses on reducing the environmental impacts of chemicals through sustainable practices. Traditional methods for extracting bioactive compounds from Eucalyptus marginata leaves, such as hydro-distillation and organic solvent extraction, have limitations, including long extraction times, high energy consumption, and potential toxic solvent [...] Read more.
Green chemistry focuses on reducing the environmental impacts of chemicals through sustainable practices. Traditional methods for extracting bioactive compounds from Eucalyptus marginata leaves, such as hydro-distillation and organic solvent extraction, have limitations, including long extraction times, high energy consumption, and potential toxic solvent residues. This study explored the use of supercritical fluid extraction (SFE), pressurized liquid extraction (PLE), and gas-expanded liquid (GXL) processes to improve efficiency and selectivity. These techniques were combined in a single mixture design, where CO2 was used in the experiments carried out under SFE, while water and ethanol were used for the PLE and GXL experiments by varying the concentration of the solvents to cover all the extraction possibilities. The neuroprotective activity of the extracts was evaluated by measuring their antioxidant, anti-inflammatory, and acetylcholinesterase inhibition properties. The optimization resulted in a novel GXL extraction with an optimal ternary mixture of 27% CO2, 55% ethanol, and 18% water, with a high degree of desirability (R2 = 88.59%). Chromatographic analysis carried out by GC-MS and HPLC-ESI-MS/MS identified over 49 metabolites. The designed sustainable extraction process offers a promising approach for producing phenolic-rich plant extracts in industrial applications. Full article
(This article belongs to the Special Issue Neuroprotective Effects of Food Ingredients)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Cumulative yield curves for various solvent compositions. (<b>b</b>) Total extraction yields represented in a ternary phase diagram.</p>
Full article ">Figure 2
<p>GC-MS chromatogram of the optimal GXL extract from <span class="html-italic">Eucalyptus marginata</span> leaves, obtained with CO<sub>2</sub>–ethanol–water (27:55:18, <span class="html-italic">v</span>/<span class="html-italic">v</span>/<span class="html-italic">v</span>) (100 bar, T = 50 °C).</p>
Full article ">Figure 3
<p>HPLC-ESI-MS chromatogram of the optimal GXL extract from <span class="html-italic">Eucalyptus marginata</span> leaves, obtained with CO<sub>2</sub>–ethanol–water (27:55:18, <span class="html-italic">v</span>/<span class="html-italic">v</span>/<span class="html-italic">v</span>) (100 bar, 50 °C).</p>
Full article ">Figure 4
<p>Suggested schematic and chemical structure for the fragmentation of quercetin-7-O-rutinoside (peak 12) in the HPLC-ESI-MS analysis.</p>
Full article ">Figure 5
<p>Suggested schematic and chemical structure for the fragmentation of kaempferol-rutinoside in the HPLC-ESI-MS analysis.</p>
Full article ">
13 pages, 1901 KiB  
Article
A Novel Pot-Economy Approach to the Synthesis of Triantennary GalNAc-Oligonucleotide
by Artem Evgenievich Gusev, Vladimir Nikolaevich Ivanov, Nikolai Andreevich Dmitriev, Aleksandr Viktorovich Kholstov, Vladislav Aleksandrovich Vasilichin, Ilya Andreevich Kofiadi and Musa Rakhimovich Khaitov
Molecules 2024, 29(24), 5959; https://doi.org/10.3390/molecules29245959 - 17 Dec 2024
Viewed by 426
Abstract
N-Acetylgalactosamine (GalNAc) is an efficient and multifunctional delivery tool in the development and synthesis of chemically modified oligonucleotide therapeutics (conjugates). Such therapeutics demonstrate improved potency in vivo due to the selective and efficient delivery to hepatocytes in the liver via receptor-mediated endocytosis, which [...] Read more.
N-Acetylgalactosamine (GalNAc) is an efficient and multifunctional delivery tool in the development and synthesis of chemically modified oligonucleotide therapeutics (conjugates). Such therapeutics demonstrate improved potency in vivo due to the selective and efficient delivery to hepatocytes in the liver via receptor-mediated endocytosis, which is what drives the high interest in this molecule. The ways to synthesize such structures are relatively new and have not been optimized in terms of the yields and stages both in lab and large-scale synthesis. Another significant criterion, especially in large-scale synthesis, is to match ecological norms and perform the synthesis in accordance with the Green Chemistry approach, i.e., to control and minimize the amounts of reagents and resources consumed and the waste generated. Here, we provide a robust and resource effective pot-economy method for the synthesis of triantennary GalNAc and GalNAc phosphoramidite/CPG optimized for laboratory scales. Full article
Show Figures

Figure 1

Figure 1
<p>Standard approach to synthesize GalNAc-L96.</p>
Full article ">Figure 2
<p>Standard approach to synthesize GalNAc-L96.</p>
Full article ">Figure 3
<p>Standard approach to synthesize GalNAc-L96.</p>
Full article ">Figure 4
<p>New approach to synthesize GalNAc-L96. First pot.</p>
Full article ">Figure 5
<p>New approach to synthesize GalNAc-L96. Second pot.</p>
Full article ">Figure 6
<p>Synthesis of L-96 GalNAc phosphoramidite and L-96 GalNAc CPG.</p>
Full article ">Figure 7
<p>Part of synthesis in first pot.</p>
Full article ">Figure 8
<p>Part of synthesis in second pot.</p>
Full article ">Figure 9
<p>Synthesis of L-96 GalNAc phosphoramidite.</p>
Full article ">Figure 10
<p>Synthesis of L-96 GalNAc CPG.</p>
Full article ">
31 pages, 15017 KiB  
Article
Green Synthesized Composite AB-Polybenzimidazole/TiO2 Membranes with Photocatalytic and Antibacterial Activity
by Hristo Penchev, Katerina Zaharieva, Silvia Dimova, Ivelina Tsacheva, Rumyana Eneva, Stephan Engibarov, Irina Lazarkevich, Tsvetelina Paunova-Krasteva, Maria Shipochka, Ralitsa Mladenova, Ognian Dimitrov, Daniela Stoyanova and Irina Stambolova
Crystals 2024, 14(12), 1081; https://doi.org/10.3390/cryst14121081 - 16 Dec 2024
Viewed by 783
Abstract
Novel AB-Polybenzimidazole (AB-PBI)/TiO2 nanocomposite membranes have been prepared using a synthetic green chemistry approach. Modified Eaton’s reagent (methansulfonic acid/P2O5) was used as both reaction media for microwave-assisted synthesis of AB-PBI and as an efficient dispersant of partially agglomerated [...] Read more.
Novel AB-Polybenzimidazole (AB-PBI)/TiO2 nanocomposite membranes have been prepared using a synthetic green chemistry approach. Modified Eaton’s reagent (methansulfonic acid/P2O5) was used as both reaction media for microwave-assisted synthesis of AB-PBI and as an efficient dispersant of partially agglomerated titanium dioxide powders. Composite membranes of 80 µm thickness have been prepared by a film casting approach involving subsequent anti-solvent inversion in order to obtain porous composite membranes possessing high sorption capacity. The maximal TiO2 filler content achieved was 20 wt.% TiO2 nanoparticles (NPs). Titania particles were green synthesized (using a different content of Mentha Spicata (MS) aqueous extract) by hydrothermal activation (150 °C), followed by thermal treatment at 400 °C. The various methods such as powder X-ray diffraction and Thermogravimetric analyses, X-ray photoelectron spectroscopy, Fourier transform infrared spectroscopy and Energy-dispersive X-ray spectroscopy, Electronic paramagnetic resonance, Scanning Electron Microscopy and Transmission Electron Microscopy have been used to study the phase and surface composition, structure, morphology, and thermal behavior of the synthesized nanocomposite membranes. The photocatalytic ability of the so-prepared AB-Polybenzimidazole/bio-TiO2 membranes was studied for decolorization of Reactive Black 5 (RB5) as a model azo dye pollutant under UV light illumination. The polymer membrane in basic form, containing TiO2 particles, was obtained with a 40 mL quantity of the MS extract, exhibiting the highest decolorization rate (96%) after 180 min of UV irradiation. The so-prepared AB-Polybenzimidazole/TiO2 samples have a powerful antibacterial effect on E. coli when irradiated by UV light. Full article
Show Figures

Figure 1

Figure 1
<p>TEM micrographs and particle size distribution of M0 (<b>a</b>), M1 (<b>b</b>), and M2 (<b>c</b>) particles.</p>
Full article ">Figure 1 Cont.
<p>TEM micrographs and particle size distribution of M0 (<b>a</b>), M1 (<b>b</b>), and M2 (<b>c</b>) particles.</p>
Full article ">Figure 2
<p>Diffuse-reflection spectra of green synthesized TiO<sub>2</sub> particles.</p>
Full article ">Figure 3
<p>Absorption spectra of the green synthesized TiO<sub>2</sub> particles with Kubelka–Munk conversion.</p>
Full article ">Figure 4
<p>Tauc’s plots of green synthesized TiO<sub>2</sub> particles.</p>
Full article ">Figure 5
<p>EPR spectra (right) of green synthesized TiO<sub>2</sub> particles.</p>
Full article ">Figure 6
<p>General preparation scheme for the synthesis of composite AB-PBI/TiO<sub>2</sub> membranes and reaction parameter comparison of the conventional (<b>left</b>) and microwave-assisted approaches (<b>right</b>).</p>
Full article ">Figure 7
<p>PXRD patterns of (<b>a</b>) green synthesized TiO<sub>2</sub> and (<b>b</b>) AB-PBI/bio-TiO<sub>2</sub> membranes and (<b>c</b>) pristine AB-PBI.</p>
Full article ">Figure 8
<p>Deconvolution of C1s, O1s, and N1s core level spectra of the AB-PBI-TiO<sub>2</sub>, M1 membranes (neutralized and acid-doped forms).</p>
Full article ">Figure 9
<p>EPR spectra of 1—polybenzimidazole; 2—PBI/bio-TiO<sub>2</sub>, M1 before UV irradiation; 3—PBI/bio-TiO<sub>2</sub>, M1 after UV irradiation recorded at room temperature; 4—PBI/bio-TiO<sub>2</sub>, M1 after UV irradiation recorded at 123 K.</p>
Full article ">Figure 10
<p>SEM images of (<b>A</b>) AB-PBI/bio-TiO<sub>2</sub>, M1 (basic form) and (<b>B</b>) AB-PBI/bio-TiO<sub>2</sub>, M1 (acid-doped form) membranes.</p>
Full article ">Figure 11
<p>EDS mapping of (<b>A</b>) AB-PBI/bio-TiO<sub>2</sub>, M1 (neutralized form) and (<b>B</b>) AB-PBI/bio-TiO<sub>2</sub>, M1 (acid-doped form) membranes.</p>
Full article ">Figure 11 Cont.
<p>EDS mapping of (<b>A</b>) AB-PBI/bio-TiO<sub>2</sub>, M1 (neutralized form) and (<b>B</b>) AB-PBI/bio-TiO<sub>2</sub>, M1 (acid-doped form) membranes.</p>
Full article ">Figure 12
<p>FTIR spectra of green synthesized TiO<sub>2</sub> and AB-PBI/bio-TiO<sub>2</sub> membranes.</p>
Full article ">Figure 13
<p>Thermogravimetric curves of AB-PBI/TiO<sub>2</sub> membranes.</p>
Full article ">Figure 14
<p>Kinetic curves of UV decolorization of Reactive Black 5 dye using AB-PBI/bio-TiO<sub>2</sub> membranes as photocatalysts.</p>
Full article ">Figure 15
<p>Degree of decolorization of RB 5 dye during UV irradiation time period using (<b>a</b>,<b>b</b>) AB-PBI/bio-TiO<sub>2</sub>, M0 (neutralized and acid-doped forms); (<b>c</b>,<b>d</b>) AB-PBI/bio-TiO<sub>2</sub>, M1 (neutralized and acid-doped forms); (<b>e</b>,<b>f</b>) AB-PBI/bio-TiO<sub>2</sub>, M2 (neutralized and acid-doped forms) membranes as photocatalysts.</p>
Full article ">Figure 16
<p>UV–Vis absorption spectra of RB 5 dye during irradiation time period using AB-PBI/bio-TiO<sub>2</sub>, M1 (basic form) as the photocatalyst.</p>
Full article ">Figure 17
<p>The adsorption capacities (Q) (mg/g) of (1) and (2) AB-PBI/bio-TiO<sub>2</sub>, M2 (acid-doped and neutralized forms); (3) and (4) AB-PBI/bio-TiO<sub>2</sub>, M1 (neutralized and acid-doped forms); (5) and (6) AB-PBI/bio-TiO<sub>2</sub>, M0 (acid-doped and neutralized forms) membranes after a 30 min dark period.</p>
Full article ">Figure 18
<p>Degree of decolorization of RB 5 dye after 180 min under UV irradiation using basic form membranes in three photocatalytic runs. (<b>a</b>) AB-PBI/bio-TiO<sub>2</sub>, M0; (<b>b</b>) AB-PBI/bio-TiO<sub>2</sub>, M1; and (<b>c</b>) AB-PBI/bio-TiO<sub>2</sub>, M2.</p>
Full article ">Figure 19
<p>(<b>A</b>) Antimicrobial effect of the UV-irradiated (violet columns) suspensions of M0, M1, and M2 (0.5 mg/mL) with <span class="html-italic">E. coli</span> compared with their equivalents kept in the dark (gray columns) expressed as CFU/mL. (<b>B</b>) The decrease of CFU under UV light is well visible in the petri dishes.</p>
Full article ">Figure 20
<p>Antimicrobial effect of the composite membranes AB-PBI/bio-TiO<sub>2</sub>, M0, AB-PBI/bio-TiO<sub>2</sub>, M1, and AB-PBI/bio-TiO<sub>2</sub>, M2 on <span class="html-italic">E. coli</span> in the dark tested by the ASTM Standard Test Method E 2149–10. Control samples contain only bacterial suspension.</p>
Full article ">Figure 21
<p>Effect of AB-PBI-TiO<sub>2</sub> composites on standard <span class="html-italic">E. coli</span> suspension under UV irradiation for 10 min. Control samples contain only bacterial suspension.</p>
Full article ">Figure 22
<p>Representative SEM micrographs revealing the surface morphology and adhesion of <span class="html-italic">E. coli</span> 25922 during cultivation with composite membranes AB-PBI/bio-TiO<sub>2</sub>, M0 ((<b>a</b>)—treated with UV, 10 min. (<b>b</b>)—untreated), AB-PBI/bio-TiO<sub>2</sub>, M1 ((<b>c</b>)—treated with UV, 10 min. (<b>d</b>)—untreated) and AB-PBI/bio-TiO<sub>2</sub>, M2 ((<b>e</b>)—treated with UV, 10 min, (<b>f</b>)—untreated). Designations: White arrows—blebs or invaginations; white triangle—ruptured cells; white stars—amorphous substance. Zoom images highlight some of the damage in the bacterial cells. Bars = 5 μm.</p>
Full article ">
27 pages, 4876 KiB  
Article
Halogenated Cobalt Bis-Dicarbollide Strong Acids as Reusable Homogeneous Catalysts for Fatty Acid Esterification with Methanol or Ethanol
by Pavel Kaule, Václav Šícha, Jan Macháček, Yelizaveta Naumkina and Jan Čejka
Int. J. Mol. Sci. 2024, 25(24), 13263; https://doi.org/10.3390/ijms252413263 - 10 Dec 2024
Viewed by 671
Abstract
The most commonly used homogeneous catalyst for fatty acid esterification is a corrosive sulphuric acid. However, this requires costly investment in non-corrosive equipment, presents a safety risk, is time consuming, and increases effluent generation. In this study, inorganic 3D heteroborane cluster strong acids [...] Read more.
The most commonly used homogeneous catalyst for fatty acid esterification is a corrosive sulphuric acid. However, this requires costly investment in non-corrosive equipment, presents a safety risk, is time consuming, and increases effluent generation. In this study, inorganic 3D heteroborane cluster strong acids are employed for the first time as homogeneous catalysts. Three novel isomeric tetrachlorido and tetrabromido derivatives of 3,3′-commo-bis[undecahydrido-closo-1,2-dicarba-3-cobaltadodecaborate](1−) [1] were synthesised and fully characterised using a range of analytical techniques, including NMR, TLC, HPLC, MS, UV-Vis, melting point (MP), CHN analyses, and XRD. Ultimately, H3O[8,8′-Cl2-1] was identified as the most efficient, reusable, and non-corrosive homogeneous catalyst for the esterification of four fatty acids. The reactions are conducted in an excess of alcohol at reflux. The effective absorption of water vapour provided by the molecular sieves maximises acid conversion. The hydrophobic dye Sudan black B was employed as an acid-base indicator to facilitate a comparison of the H0 acidity function of sulphuric acid and halogenated heteroboranoic acids when dissolved together in methanol. The 23Na NMR analysis demonstrated that the application of dry methanol resulted in the displacement of Na+ ions from zeolite, which subsequently exchanged the H3O+ ions of the acid. This process led to a gradual reduction in the efficiency of the catalysts, particularly with repeated use. The solution to this issue is to regenerate the catalyst on the ion exchanger following each reaction. In contrast to the published methods, our new approach meets 10 of 12 green chemistry principles. Full article
Show Figures

Figure 1

Figure 1
<p>The structural formula of COSAN anion is 3,3′-<span class="html-italic">commo</span>-bis[undecahydrido-<span class="html-italic">closo</span>-1,2-dicarba-3-cobaltadodecaborate](1−) [<b>1<sup>−</sup></b>]. The numbered vertices in the polyhedron represent the boron atoms. Each C or B atom is substituted terminally with one hydrogen atom. However, for clarity, cage terminal H atoms attached to numbered B and C vertices have been omitted [<a href="#B62-ijms-25-13263" class="html-bibr">62</a>]. The asterisk * has been used to clearly and visibly distinguish the cluster boron and carbon atoms in the lower cage. The positions of B8 (upper cage) and B8* (lower cage) are different.</p>
Full article ">Figure 2
<p>Reaction scheme of mono-, di-, tetrachlorido-, and tetrabromido derivatives of [<b>1<sup>−</sup></b>] preparation. For clarity, cage terminal H atoms attached to numbered B and C vertices have been omitted. The asterisk * has been used to clearly and visibly distinguish the cluster boron and carbon atoms in the lower cage. The positions of B8 (upper cage) and B8* (lower cage) are different.</p>
Full article ">Figure 3
<p>An ORTEP structure of the isomer Cs[8,8′, 9, 9′-Cl<sub>4</sub>-<b>1<sup>−</sup></b>]·2 C<sub>6</sub>H<sub>6</sub> determined using XRD of the single crystal.</p>
Full article ">Figure 4
<p>An ORTEP structure of the isomer Cs[8,8′,12,12′-Cl<sub>4</sub>-<b>1<sup>−</sup></b>]·2 C<sub>6</sub>H<sub>6</sub> determined using XRD of the single crystal.</p>
Full article ">Figure 5
<p>The structural formulas of commercial Sudan Black B dye components.</p>
Full article ">Figure 6
<p>The overlapped UV-Vis spectra of all used acids (5 × 10<sup>−3</sup> mol) after reaction with the Sudan Black B (III) indicator (7.2 × 10<sup>−5</sup> mol) were measured in pure methanol at 25 °C. The other collected UV-Vis spectra have been included in the <a href="#app1-ijms-25-13263" class="html-app">Supplementary Information (Figures S13 and S14)</a>.</p>
Full article ">Figure 7
<p>The conversion of fatty acid methyl esters (FAMEs) to their corresponding acid using methanol as the solvent and 1 mol% of the catalyst under desiccated conditions was investigated.</p>
Full article ">Figure 8
<p>The conversion of acid to FAEEs with ethanol and a 1 mol% catalyst under desiccation conditions was investigated.</p>
Full article ">Figure 9
<p>Acids conversions of FAMEs with methanol and various mol. % of the H<sub>3</sub>O[Cl<sub>2</sub>-<b>1<sup>−</sup></b>] under desiccation.</p>
Full article ">Figure 10
<p>Acids conversions of FAEEs with ethanol and various mol. % of the H<sub>3</sub>O[Cl<sub>2</sub>-<b>1<sup>−</sup></b>] under desiccation.</p>
Full article ">Figure 11
<p>Acids conversions of FAMEs and FAEEs with 1 mol. % of the catalyst under desiccation.</p>
Full article ">Figure 12
<p>Repeatability of fatty acid conversions into FAMEs after 3 reactions with 1 mol.% of H<sub>3</sub>O[Cl<sub>2</sub>-<b>1<sup>−</sup></b>] without reactivation treatment.</p>
Full article ">Figure 13
<p>Repeatability of fatty acid conversions into FAEEs after 3 reactions with 1 mol.% of H<sub>3</sub>O[Cl<sub>2</sub>-<b>1<sup>−</sup></b>] without reactivation treatment.</p>
Full article ">Figure 14
<p>Repeatability of fatty acid conversions with methanol and 1 mol. % of H<sub>3</sub>O[Cl<sub>2</sub>-<b>1<sup>−</sup></b>] after 3 reactions after reactivation of the catalyst on Amberlyst15.</p>
Full article ">Figure 15
<p>Repeatability of fatty acid conversions with ethanol and 1 mol. % of H<sub>3</sub>O[Cl<sub>2</sub>-<b>1<sup>−</sup></b>] after 3 reactions after reactivation of the catalyst on Amberlyst15.</p>
Full article ">Figure 16
<p>Differences in fatty acid conversions with 1 mol. % of H<sub>3</sub>O[Cl<sub>2</sub>-<b>1<sup>−</sup></b>] in methanol.</p>
Full article ">Figure 17
<p>Differences in fatty acid conversions with 1 mol. % of H<sub>3</sub>O[Cl<sub>2</sub>-<b>1<sup>−</sup></b>] in ethanol.</p>
Full article ">Figure 18
<p>Differences in fatty acid conversions with 1 mol. % of H<sub>3</sub>O.HSO<sub>4</sub> in methanol.</p>
Full article ">Figure 19
<p>Differences in fatty acid conversions with 1 mol. % of H<sub>3</sub>O.HSO<sub>4</sub> in ethanol.</p>
Full article ">
12 pages, 2295 KiB  
Review
Phosphite as a Sustainable and Versatile Alternative for Biostimulation, Biocontrol, and Weed Management in Modern Agriculture
by Libia Iris Trejo-Téllez, Víctor Hugo Carbajal-Vázquez, Jazmín Lavín-Castañeda and Fernando Carlos Gómez-Merino
Processes 2024, 12(12), 2764; https://doi.org/10.3390/pr12122764 - 5 Dec 2024
Viewed by 574
Abstract
Phosphite (Phi), an analog of phosphate (Pi), is an anion widely used in phytosanitary management and agricultural biostimulation schemes. Given that, unlike some species of bacteria, plants do not naturally have the mechanisms to metabolize Phi once they have absorbed it, Phi must [...] Read more.
Phosphite (Phi), an analog of phosphate (Pi), is an anion widely used in phytosanitary management and agricultural biostimulation schemes. Given that, unlike some species of bacteria, plants do not naturally have the mechanisms to metabolize Phi once they have absorbed it, Phi must be used in perfect coordination with adequate nutritional management of Pi in the crop since an excessive level of Phi combined with a deficient supply of Pi causes a disruption in ionic balances that can result in serious toxicity or even the death of the plant. In addition to the adequate Phi/Pi balance, high doses of Phi by themselves cause alterations in the mechanisms of perception and response to phosphorus deficiency leading to toxicity in plants. Hence, in various plant species, it has been proven that Phi can be used with herbicidal effects. Genes that encode enzymes involved in the metabolization of Phi have been isolated from bacterial genomes, and they have been transferred by genetic engineering to plant genomes, allowing the development of dual fertilization and weed control systems. This review provides background on the novel uses of Phi in agriculture and breaks down its potential use as an alternative herbicide in sustainable agriculture approaches supported by green chemistry. Full article
(This article belongs to the Special Issue Feature Review Papers in Section "Environmental and Green Processes")
Show Figures

Figure 1

Figure 1
<p>Structure of the phosphate (Pi; PO<sub>4</sub><sup>3−</sup>) and phosphite (Phi; HPO<sub>3</sub><sup>2−</sup>) anions, showing the replacement of an oxygen atom (O) in Pi by a hydrogen one (H) in the tetrahedral configuration in Phi.</p>
Full article ">Figure 2
<p>Mobility of phosphate (Pi) in the soil towards the roots, following the concentration gradient. Phosphate, when applied to the soil as a solid fertilizer, tends to fix, and its mobility is negatively affected. Additionally, the diffusion coefficient of Pi is low in soil. The Pi depletion zone favors its access to the root.</p>
Full article ">Figure 3
<p>Mobility of phosphite (Phi) in the plant. Liquid applications of phosphite to the plant have great mobility in the conductive tissues (xylem and phloem), so its potential hormetic effects are more readily evident.</p>
Full article ">
17 pages, 10370 KiB  
Article
Green Synthesis of Zinc Oxide Nanoparticles for Tetracycline Adsorption: Experimental Insights and DFT Study
by Solhe F. Alshahateet, Salah A. Al-Trawneh, Mohammed Er-rajy, Mohammed Zerrouk, Khalil Azzaoui, Waad M. Al-Tawarh, Belkheir Hammouti, Rachid Salghi, Rachid Sabbahi, Mohammed M. Alanazi and Larbi Rhazi
Plants 2024, 13(23), 3386; https://doi.org/10.3390/plants13233386 - 2 Dec 2024
Viewed by 822
Abstract
An eco-friendly approach was used to fabricate zinc oxide nanoparticles (ZnO NPs) using thyme, Thymus vulgaris L., leaf extract. The produced ZnO nanoparticles were characterized by XRD and SEM analysis. The ZnO NPs showed remarkable adsorption efficiency for tetracycline (TC) from water systems, [...] Read more.
An eco-friendly approach was used to fabricate zinc oxide nanoparticles (ZnO NPs) using thyme, Thymus vulgaris L., leaf extract. The produced ZnO nanoparticles were characterized by XRD and SEM analysis. The ZnO NPs showed remarkable adsorption efficiency for tetracycline (TC) from water systems, with a maximum removal rate of 95% under optimal conditions (10 ppm, 0.10 g of ZnO NPs, pH 8.5, and 30 min at 25 °C). The adsorption kinetics followed the pseudo-2nd-order model, and the adsorption process fitted the Temkin isotherm model. The process was spontaneous, endothermic, and primarily chemisorptive. Quantum chemistry calculations, utilizing electrostatic potential maps and HOMO-LUMO gap analysis, have confirmed the stability of the TC clusters. This study suggests that green synthesis using plant extracts presents an opportunity to generate nanoparticles with properties suitable for real-world applications. Full article
Show Figures

Figure 1

Figure 1
<p>XRD pattern of ZnO NPs.</p>
Full article ">Figure 2
<p>SEM image of synthesized ZnO NPs using thyme, <span class="html-italic">Thymus vulgaris</span> L., leaf extract, (<b>a</b>) ×20,000, (<b>b</b>) ×40,000.</p>
Full article ">Figure 3
<p>External calibration curve for the determination of tetracycline concentrations.</p>
Full article ">Figure 4
<p>Impact of ZnO NPs dosage on removal efficiency and adsorption capacity. Experimental conditions: contact time, 30 min; starting tetracycline (TC) concentration, 10 ppm; temperature, 25 °C.</p>
Full article ">Figure 5
<p>Impact of starting tetracycline (TC) concentration on adsorption by ZnO NPs. Experimental conditions: nanosorbent dosage, 0.10 g; contact time, 30 min; pH, 7.6; temperature, 25 °C.</p>
Full article ">Figure 6
<p>Impact of contact duration on tetracycline (TC) adsorption by ZnO NPs. Experimental conditions: nanosorbent dosage, 0.10 g; pH, 7.6; starting TC concentration, 10 ppm; temperature, 25 °C.</p>
Full article ">Figure 7
<p>Effect of pH on tetracycline (TC) adsorption by ZnO NPs. Experimental conditions: nanosorbent dosage, 0.10 g; contact time, 30 min; starting TC concentration, 10 ppm; temperature, 25 °C.</p>
Full article ">Figure 8
<p>Impact of temperature on tetracycline (TC) adsorption by ZnO NPs. Experimental conditions: nanosorbent dosage, 0.10 g; pH, 8.5; contact time, 30 min; starting TC concentration, 10 ppm.</p>
Full article ">Figure 9
<p>Kinetic plots of (<b>a</b>) pseudo-1st order and (<b>b</b>) pseudo-2nd order for tetracycline adsorption onto ZnO NPs.</p>
Full article ">Figure 10
<p>Isotherm plots of (<b>a</b>) Langmuir isotherm, (<b>b</b>) Freundlich isotherm, and (<b>c</b>) Temkin isotherm of tetracycline adsorption onto ZnO NPs.</p>
Full article ">Figure 11
<p>Optimized geometry of tetracycline, showing atom positions (red = oxygen, blue = nitrogen, grey = carbon).</p>
Full article ">Figure 12
<p>Frontier molecular orbitals of tetracycline and zinc oxide.</p>
Full article ">Figure 13
<p>Molecular electrostatic potential maps for tetracycline and zinc oxide.</p>
Full article ">Figure 14
<p>Non-covalent interaction (NCI) and reduced density gradient (RDG) diagrams for tetracycline.</p>
Full article ">Figure 15
<p>Biosynthesis steps for zinc oxide nanoparticles and their potential applications.</p>
Full article ">
19 pages, 7579 KiB  
Article
Self-Assembled Hydrogel Based on (Bio)polyelectrolyte Complex of Chitosan–Gelatin: Effect of Composition on Physicochemical Properties
by Kashurin Aleksandr, Litvinov Mikhail and Podshivalov Aleksandr
Gels 2024, 10(12), 786; https://doi.org/10.3390/gels10120786 - 1 Dec 2024
Viewed by 549
Abstract
Taking into account the trends in the field of green chemistry and the desire to use natural materials in biomedical applications, (bio)polyelectrolyte complexes ((bio)PECs) based on a mixture of chitosan and gelatin seem to be relevant systems. Using the approach of self-assembly from [...] Read more.
Taking into account the trends in the field of green chemistry and the desire to use natural materials in biomedical applications, (bio)polyelectrolyte complexes ((bio)PECs) based on a mixture of chitosan and gelatin seem to be relevant systems. Using the approach of self-assembly from the dispersion of the coacervate phase of a (bio)PEC at different ratios of ionized functional groups of chitosan and gelatin (z), hydrogels with increased resistance to mechanical deformations and resorption in liquid media were obtained in this work in comparison to a hydrogel from gelatin. It was found that at z ≥ 1 a four-fold increase in the elastic modulus of the hydrogel occurred in comparison to a hydrogel based on gelatin. It was shown that hydrogels at z ≈ 1 had an increased sorption capacity and water sorption rate, as well as increased resistance to the in vitro model environment of phosphate-buffered saline (PBS) solution containing lysozyme at 37 °C. It was also shown that in PBS and simulated gastric fluid (SGF) solutions, the effect of the polyelectrolyte swelling of the hydrogels was significantly suppressed; however, at z ≥ 1, the (bio)PEC hydrogels had increased stability compared to the samples at z < 1 and based on gelatin. Full article
(This article belongs to the Section Gel Chemistry and Physics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Mass yield of Chit–Gel (bio)PEC hydrogels versus <span class="html-italic">z</span> value.</p>
Full article ">Figure 2
<p>Appearance of self-assembled Chit–Gel (bio)PEC hydrogels at shown <span class="html-italic">z</span> values.</p>
Full article ">Figure 3
<p>SEM microphotographs of Chit–Gel (bio)PEC hydrogels at (<b>a</b>) <span class="html-italic">z</span> = 0.58, (<b>b</b>) 0.89, (<b>c</b>) 1.15, (<b>d</b>) 1.44, and (<b>e</b>) 1.73.</p>
Full article ">Figure 4
<p>ATR-FTIR spectra of Gelatin, Chitosan, and Chit–Gel (bio)PEC hydrogels at different <span class="html-italic">z</span> values.</p>
Full article ">Figure 5
<p><span class="html-italic">G</span>′ versus <span class="html-italic">γ</span> profiles for Gel and Chit–Gel (bio)PEC hydrogels at different <span class="html-italic">z</span> values at temperature of 25.0 ± 0.2 °C.</p>
Full article ">Figure 6
<p>Sorption curves of Gelatin and Chit–Gel (bio)PEC hydrogel lyophilizates in (<b>a</b>) distilled water (pH = 6) at 25 °C, (<b>b</b>) <span class="html-italic">PBS</span> solution (pH = 7.4) at 25 °C, (<b>c</b>) <span class="html-italic">SGF</span> solution (pH = 1.2) at 25 °C, and (<b>d</b>) <span class="html-italic">PBS</span> with lysozyme solution (pH = 7.4) at 37 °C.</p>
Full article ">Figure 7
<p>Kinetic dependences of liquid sorption process of Chit–Gel (bio)PEC hydrogel lyophilisates in (<b>a</b>,<b>b</b>) distilled water and (<b>c</b>,<b>d</b>) <span class="html-italic">PBS</span> and (<b>e</b>,<b>f</b>) <span class="html-italic">SGF</span> solutions and their analytical fitting with Equations (5) and (6), respectively.</p>
Full article ">
12 pages, 3497 KiB  
Article
Selenium Disulfide from Sustainable Resources: An Example of “Redneck” Chemistry with a Pinch of Salt
by Eduard Tiganescu, Shahrzad Safinazlou, Ahmad Yaman Abdin, Rainer Lilischkis, Karl-Herbert Schäfer, Claudia Fink-Straube, Muhammad Jawad Nasim and Claus Jacob
Materials 2024, 17(23), 5733; https://doi.org/10.3390/ma17235733 - 23 Nov 2024
Viewed by 796
Abstract
Selenium disulfide (often referred to as SeS2) encompasses a family of mixed selenium-sulfide eight-membered rings, traditionally used as an anti-dandruff agent in shampoos. SeS2 can be produced by reacting hydrogen sulfide (H2S) with selenite (SeO32−) [...] Read more.
Selenium disulfide (often referred to as SeS2) encompasses a family of mixed selenium-sulfide eight-membered rings, traditionally used as an anti-dandruff agent in shampoos. SeS2 can be produced by reacting hydrogen sulfide (H2S) with selenite (SeO32−) under acidic conditions. This chemistry is also possible with natural spring waters that are rich in H2S, thus providing an avenue for the more sustainable, green production of high-quality SeS2 particles from an abundant natural source. The orange material obtained this way consists of small globules with a diameter in the range of 1.1 to 1.2 µm composed of various SexS8−x chalcogen rings. It shows the usual composition and characteristics of a Se-S interchalcogen compound in EDX and Raman spectroscopy. Since the mineral water from Bad Nenndorf is also rich in salts, the leftover brine has been evaporated to yield a selenium-enriched salt mixture similar to table salt. As the water from Bad Nenndorf—in comparison to other bodies of water around the world—is still rather modest in terms of its H2S content, especially when compared with volcanic waters, this approach may be refined further to become economically and ecologically viable, especially as a regional business model for small and medium-sized enterprises. Full article
Show Figures

Figure 1

Figure 1
<p>The H<sub>2</sub>S concentration present in the water samples gradually decreases as affirmed using the MB assay.</p>
Full article ">Figure 2
<p>The water was collected from the underground spring in a field near Bad Nenndorf in northern Germany. A preliminary reaction was carried out “redneck style” at the source of origin, and an immediate change in color confirmed the feasibility of the synthesis. The figure also represents the chemistry carried out in the laboratory and shows a photograph of the orange material obtained.</p>
Full article ">Figure 3
<p>Raman spectroscopy confirmed the presence of S-S, S-Se, and Se-Se bonds.</p>
Full article ">Figure 4
<p>SeS<sub>2</sub> was analyzed to determine chemical composition using EDX coupled to SEM. EDX confirmed the presence of selenium and sulfur at a ratio of around 1:2 (Panel (<b>a</b>)), while the SEM image showed the presence of (aggregated) globular material (Panel (<b>b</b>)).</p>
Full article ">Figure 5
<p>The filtrate was evaporated at 50 °C to obtain salt (Panel (<b>a</b>)), which was analyzed by SEM (Panel (<b>b</b>)) coupled with EDX (Panel (<b>c</b>)) to quantify the elements present in the salt. EDX confirmed the presence of selenium at about 0.40% <span class="html-italic">w</span>/<span class="html-italic">w</span> (dry weight), as compared to the overall salt composition.</p>
Full article ">Figure 6
<p>H<sub>2</sub>S springs may serve as sources for the production of value-added products, such as SeS<sub>2,</sub> and avoid wasting this natural resource as sewage. This strategy not only opens up the door for boosting local economies but also, as a true “hat trick”, decreases the environmental burden posed by the chemical treatment of H<sub>2</sub>S-rich water.</p>
Full article ">
Back to TopTop