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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (4,036)

Search Parameters:
Keywords = organic solvents

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 4174 KiB  
Review
Study of Grain Boundary: From Crystallization Engineering to Machine Learning
by Zhengran He, Sheng Bi and Kyeiwaa Asare-Yeboah
Coatings 2025, 15(2), 164; https://doi.org/10.3390/coatings15020164 - 2 Feb 2025
Viewed by 270
Abstract
Grain boundaries play a vital role in determining the structural, functional, mechanical, and electrical properties of semiconductor materials. Recent studies have yielded great advances in understanding and modulating the grain boundaries via semiconductor crystallization engineering and machine learning. In this article, we first [...] Read more.
Grain boundaries play a vital role in determining the structural, functional, mechanical, and electrical properties of semiconductor materials. Recent studies have yielded great advances in understanding and modulating the grain boundaries via semiconductor crystallization engineering and machine learning. In this article, we first provide a review of the miscellaneous methods and approaches that effectively control the nucleation formation, semiconductor crystallization, and grain boundary of organic semiconductors. Using the benchmark small molecular semiconductor 6,13-bis(triisopropylsilylethynyl) pentacene (TIPS pentacene) as a representative example, the crystallization engineering methods include polymer additive mixing, solvent annealing, gas injection, and substrate temperature control. By studying the grain-width-dependent charge transport, we propose a grain boundary model as a fundamental basis to theoretically understand the intrinsic relation between grain boundary engineering and charge carrier mobility. Furthermore, we discuss the various machine learning algorithms and models used to analyze grain boundaries for the various important traits and properties, such as grain boundary crystallography, energy, mobility, and dislocation density. This work highlights the unique advantages of both crystallization engineering and machine learning methods, demonstrates new insights into discovering the presence of grain boundaries and understanding new properties of materials, and sheds light on the great potential of material application in various fields, such as organic electronics. Full article
Show Figures

Figure 1

Figure 1
<p>Polarized optical images showing the effect of the PEO polymer additive on modulating the grain boundary of TIPS pentacene organic crystals, including different weight ratios of (<b>a</b>) 5% and (<b>b</b>) 10%. The triangle in (<b>a</b>,<b>b</b>) marks the uncovered substrate. (<b>c</b>) A plot showing the change in TIPS pentacene crystal grain boundary with different loading ratios of PEO, based on six measurements. The red arrows in (<b>a</b>,<b>b</b>) represent the long axis [210] of TIPS pentacene crystals, which is also shown in (<b>c</b>). Crystal width, or <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>W</mi> </mrow> <mrow> <mi>G</mi> </mrow> </msub> </mrow> </semantics></math>, is measured as the width along the short axis of [1<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mn>2</mn> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>0]. (<b>d</b>) A cartoon showing the improvement of TIPS pentacene (TP) morphology and highlighting the improved crystal alignment and grain boundary enlargement as a result of PEO polymer addition. Reproduced from reference [<a href="#B200-coatings-15-00164" class="html-bibr">200</a>], with permission from Springer.</p>
Full article ">Figure 2
<p>Polarized optical images in subfigures (<b>a</b>–<b>d</b>) showing the enlarged grain width of TIPS pentacene crystals with the blending of poly(butyl acrylate) as a polymer additive. All images have the same scale bar as in (<b>a</b>). The crystal long axis [210] is marked with arrows. Reproduced from reference [<a href="#B201-coatings-15-00164" class="html-bibr">201</a>], with permission from Springer.</p>
Full article ">Figure 3
<p>AFM height images of (<b>a</b>) as-cast TIPS pentacene film, (<b>b</b>) TIPS pentacene film with TMB solvent vapor annealing, and (<b>c</b>) TIPS pentacene film with toluene solvent vapor annealing. Reproduced from reference [<a href="#B187-coatings-15-00164" class="html-bibr">187</a>], with permission from MDPI.</p>
Full article ">Figure 4
<p>(<b>a</b>) An optical micrograph image of TIPS pentacene needles with grain widths larger than 6 µm. (<b>b</b>) An AFM image of rounded and isotropic-shaped TIPS pentacene film with grain widths smaller than 4 µm. The gold contact electrode is shown in the bottom right corner. Reproduced from reference [<a href="#B203-coatings-15-00164" class="html-bibr">203</a>], with permission from AIP Publishing.</p>
Full article ">Figure 5
<p>Thin-film morphology of TIPS pentacene grains depending on different Ar gas flow rates, including (<b>a</b>) at 300 sccm: no clear correlation between alignment and flow rate, (<b>b</b>) at 500 sccm: crystal alignment started to be correlated with flow direction, (<b>c</b>) at 1000 sccm: the elongated crystal grains were aligned along the flow direction. Reproduced from reference [<a href="#B158-coatings-15-00164" class="html-bibr">158</a>], with permission from Elsevier.</p>
Full article ">Figure 6
<p>Optical microscope images showing the morphologies of TIPS pentacene single droplet based on different drying temperatures, including room temperature (RT), 36 °C, 46 °C, and 56 °C, respectively. Reproduced from reference [<a href="#B205-coatings-15-00164" class="html-bibr">205</a>], with permission from Elsevier.</p>
Full article ">Figure 7
<p>Visualization of the [<a href="#B100-coatings-15-00164" class="html-bibr">100</a>] symmetric tilt grain boundaries based on SPRING. The colored nodes represent the grain boundaries dependent on the structural unit content in the grain boundaries, as indicated in the color scale. White nodes indicate an absence of structural unit in the grain boundary. Reproduced from reference [<a href="#B230-coatings-15-00164" class="html-bibr">230</a>], with permission from Elsevier.</p>
Full article ">Figure 8
<p>Illustration of the construction process for the ASR and LER. ASR: after forming a SOAP matrix, an averaged SOAP vector is produced based on the sum of Q columns in the matrix, which represents all grain boundaries. LER: the grouped SOAP vectors are reduced to unique vectors (a set of U) represented by unique LAE based on the SOAP similarity metric. A histogram was generated based on each grain boundary that counts the number of unique vector examples existing in the grain boundary. The LER matrix is a collection of histograms showing LEA unique for the M grain boundaries in the M × U collection. Reproduced from reference [<a href="#B227-coatings-15-00164" class="html-bibr">227</a>], with permission from Springer Nature.</p>
Full article ">
27 pages, 20660 KiB  
Article
Microwave-Assisted In-Situ Synthesis of Polyethersulfone–ZnO Nanocomposite Membranes for Dye Removal: Enhanced Antifouling, Self-Cleaning, and Antibacterial Properties
by Lassaad Gzara, Ibtissem Ounifi, Hussam Organji, Faïçal Khlissa, Iqbal Ahmed Moujdin, Abdulmohsen Omar Alsaiari, Mohamed Abdel Salam and Amor Hafiane
Polymers 2025, 17(3), 398; https://doi.org/10.3390/polym17030398 - 2 Feb 2025
Viewed by 275
Abstract
Microwave-assisted synthesis presents a promising method for enhancing the formation of nanocomposites due to its rapid heating and uniform energy distribution. In this study, we successfully fabricated polyethersulfone–zinc-oxide (PES-ZnO) nanocomposite membranes by exposing PES/ZnCl2/DMF dope solutions to microwave radiation. Before synthesizing [...] Read more.
Microwave-assisted synthesis presents a promising method for enhancing the formation of nanocomposites due to its rapid heating and uniform energy distribution. In this study, we successfully fabricated polyethersulfone–zinc-oxide (PES-ZnO) nanocomposite membranes by exposing PES/ZnCl2/DMF dope solutions to microwave radiation. Before synthesizing the membranes, zinc-oxide nanoparticles (ZnO-NPs) were optimized in an organic phase using microwave radiation to ensure effective nanoparticle formation. The synthesis of ZnO-NPs in DMF solvent was validated through UV–Vis spectroscopy, X-ray diffraction (XRD), and Dynamic Light Scattering (DLS). We examined the surface morphology and roughness of the PES-ZnO membranes through Atomic Force Microscopy (AFM) and Scanning Electron Microscopy (SEM). Moreover, we assessed the membranes’ hydrophilicity, permeability, and physicochemical properties through contact-angle measurements, pure water flux tests, water uptake assessments, and porosity tests. Energy-dispersive X-ray spectroscopy (EDS) and X-ray diffraction (XRD) verified the successful integration of ZnO nanoparticles (ZnO-NPs) into the membrane matrix. The results indicate that including ZnO-NPs significantly improves the membrane’s permeability and hydrophilicity. The nanocomposite membranes exhibited high dye rejection efficiency, with ZnO-NPs facilitating photocatalytic self-cleaning properties. Antibacterial tests also demonstrated a substantial inhibition of common bacteria, suggesting enhanced resistance to biofouling. This research highlights the potential of microwave-assisted PES-ZnO nanocomposite membranes as effective and sustainable solutions for wastewater treatment, offering scalable applications along with added benefits of antifouling, self-cleaning, and antibacterial properties. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p>UV–Vis spectrum of ZnO prepared via microwave method.</p>
Full article ">Figure 2
<p>The sizes of the nanoparticles obtained by DLS.</p>
Full article ">Figure 3
<p>XRD of the ZnO-NPs. The XRD analysis illustrates the structural differences between ZnO-NPs before (Nps-5) and after (Nps-6) the addition of NaOH. The left side (Nps-5) exhibits no characteristic ZnO peaks, indicating an absence of crystalline ZnO formation. In contrast, the right side (Nps-6) shows well-defined ZnO diffraction peaks, confirming successful nanoparticle synthesis facilitated by NaOH addition.</p>
Full article ">Figure 4
<p>SEM images of the membranes.</p>
Full article ">Figure 5
<p>EDS analyses of PES-ZnO nanocomposite membranes.</p>
Full article ">Figure 6
<p>AFM images of the PES-ZnO nanocomposite membranes.</p>
Full article ">Figure 6 Cont.
<p>AFM images of the PES-ZnO nanocomposite membranes.</p>
Full article ">Figure 7
<p>XRD diffractograms of (<b>a</b>) the neat membranes and (<b>b</b>) the PES-ZnO nanohybrid membrane.</p>
Full article ">Figure 8
<p>(<b>a</b>) The contact angle; and (<b>b</b>) water absorption of the fabricated membranes.</p>
Full article ">Figure 9
<p>The porosity of the PES-ZnO membranes.</p>
Full article ">Figure 10
<p>The effect of pressure on pure water flux for the PES-ZnO membranes.</p>
Full article ">Figure 11
<p>Percentages of ZnO’s leakage from the prepared composites.</p>
Full article ">Figure 12
<p>Water fluxes as a function of time throughout three cycles of the BSA UF testing of the prepared PES-ZnO membranes.</p>
Full article ">Figure 13
<p>(<b>a</b>) The FRR percentage results of the membranes throughout the first two cycles. (<b>b</b>) Percentage of the total fouling ratio (Rt); and the irreversible fouling ratio (Rirr).</p>
Full article ">Figure 14
<p>Effect of pressure on dye retention. Left side (Congo Red), Right side (Methylene Blue).</p>
Full article ">Figure 15
<p>Effect of initial dye concentration. Left side (Congo Red), Right side (Methylene Blue).</p>
Full article ">Figure 16
<p>Self-cleaning of the membrane.</p>
Full article ">Figure 17
<p>The antibacterial activity of the synthesized PES-ZnO nanocomposite membranes.</p>
Full article ">Figure 18
<p>Antibacterial efficacy of PES-ZnO nanocomposite membranes: inhibition zones against Gram-negative and Gram-positive bacteria.</p>
Full article ">
25 pages, 891 KiB  
Article
Evaluation of the Pharmacological Activities of a Xylan from Corn Cobs
by Rayssa Lourenna Trigueiro Nobrega, Rony Lucas Silva Viana, Marianna Barros Silva, Luciana Duarte Martins Matta, Giulianna Paiva Viana Andrade Souza, Hugo Alexandre Oliveira Rocha and Raniere Fagundes Melo-Silveira
Polysaccharides 2025, 6(1), 9; https://doi.org/10.3390/polysaccharides6010009 (registering DOI) - 1 Feb 2025
Viewed by 278
Abstract
Xylans, polysaccharides abundantly derived from agricultural byproducts, have shown potential pharmacological properties, making them a subject of increasing research interest. This study aimed to expand the understanding of xylans’ pharmacological properties and relate them to their composition. A method combining ultrasound and alkaline [...] Read more.
Xylans, polysaccharides abundantly derived from agricultural byproducts, have shown potential pharmacological properties, making them a subject of increasing research interest. This study aimed to expand the understanding of xylans’ pharmacological properties and relate them to their composition. A method combining ultrasound and alkaline media for xylan extraction from corn cobs (ERX) was used, resulting in a significant increase in final yield compared to other methodologies. The physicochemical characterization of ERX was carried out, and its antioxidant, cytotoxic, anticoagulant, and immunomodulatory properties were evaluated. ERX demonstrated significant antioxidant activity with metal-chelating properties and induced apoptosis in HeLa tumor cells (p < 0.0001). It also reduced nitric oxide (NO) production by activated macrophages and extended the blood coagulation time, as assessed by the APTT assay (p < 0.0001). Further fractionation of ERX using various organic solvents resulted in multiple xylan subfractions. Among them, the ethanol-derived subfraction E1.4 exhibited remarkable pharmacological activities, including metal-chelation, cytotoxicity against HeLa cells via apoptosis, reduced NO production (p < 0.0001), and prolonged coagulation times (p < 0.0001). E1.4 is heteroxylan with a molecular weight of approximately 100 kDa. These findings suggest that corn cobs could be a promising source of pharmacologically significant molecules, particularly the heteroxylan E1.4. Future studies should focus on the structural characterization of this xylan to understand the relationship between structure and biological activity and explore the therapeutic potential of E1.4 in vivo models. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">
21 pages, 7878 KiB  
Article
Carboxyethylsilanetriol-Functionalized Al-MIL-53-Supported Palladium Catalyst for Enhancing Suzuki–Miyaura Cross-Coupling Reaction
by Yucang Liang, Xin Ning and Yanzhong Zhen
Molecules 2025, 30(3), 656; https://doi.org/10.3390/molecules30030656 - 1 Feb 2025
Viewed by 399
Abstract
The application of metal–organic frameworks (MOFs) has attracted increasing attention in organic synthesis. The modification of MOFs can efficiently tailor the structure and improve the property for meeting ongoing demand in various applications, such as the alteration of gas adsorption and separation, catalytic [...] Read more.
The application of metal–organic frameworks (MOFs) has attracted increasing attention in organic synthesis. The modification of MOFs can efficiently tailor the structure and improve the property for meeting ongoing demand in various applications, such as the alteration of gas adsorption and separation, catalytic activity, stability, and sustainability or reusability. In this study, carboxyethylsilanetriol (CEST) disodium salt was used as a dual-functional ligand for modified Al-MIL-53 to fabricate CEST-functionalized Al-MIL-53 samples through a hydrothermal reaction of aluminum nitrate, terephthalic acid, and CEST disodium salt by varying the molar ratio of CEST to terephthalic acid and keeping a constant molar ratio of Al3+/-COOH of 1:1. The structure, composition, morphology, pore feature, and stability were characterized by XRD, different spectroscopies, electron microscopy, N2 physisorption, and thermogravimetric analysis. With increasing CEST content, CEST-Al-MIL-53 still preserves an Al-MIL-53-like structure, but the microstructure changed compared with pure Al-MIL-53 due to the integration of CEST. Such a CEST-Al-MIL-53 was used as the support to load Pd particles and afford a catalyst Pd/CEST-Al-MIL-53 for Suzuki–Miyaura C-C cross-coupling reaction of aryl halides and phenylboronic acid under basic conditions. The resulting Pd/CEST-Al-MIL-53 showed a high catalytic activity compared with Pd/Al-MIL-53, due to the nanofibrous structure of silicon species-integrated CEST-Al-MIL-53. The nanofiber microstructure undergoes a remarkable transformation into intricate 3D cross-networks during catalytic reaction, which enables the leachable Pd particles to orientally redeposit and inlay into these networks as the monodisperse spheres and thereby effectively preventing Pd particles from aggregation and leaching, therefore demonstrating a high catalytic performance, long-term stability, and enhanced reusability. Obviously, the integration of CEST into MOFs can effectively prevent the leaching of active Pd species and ensure the re-deposition during catalysis. Moreover, catalytic performance strongly depended on catalyst dosage, temperature, time, solvent, and the type of the substituted group on benzene ring. This work further extends the catalytic application of hybrid metal–organic frameworks. Full article
(This article belongs to the Section Inorganic Chemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns of samples CEST-Al-MIL-53 (<b>1</b>–<b>3</b>) with various CEST contents, Al-MIL-53 (<b>4</b>), Pd/CEST-Al-MIL-53 (<b>5</b>), Pd/Al-MIL-53 (<b>6</b>), and simulated as-synthesized Al-MIL-53. The diffraction peaks of crystalline metallic Pd are marked as “#”. (<b>b</b>) Infrared resonance spectra of sample CEST-Al-MIL-53 (<b>2</b>), Al-MIL-53 (<b>4</b>), Pd/CEST-Al-MIL-53 (<b>5</b>), and Pd/Al-MIL-53 (<b>6</b>).</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM and (<b>b</b>) TEM image of sample <b>2</b>. (<b>c</b>,<b>d</b>) SEM images of sample <b>4</b> with different magnifications.</p>
Full article ">Figure 3
<p>(<b>a</b>) <sup>13</sup>C NMR spectra of CEST-Al-MIL-53 (<b>2</b>) and Al-MIL-53 (<b>4</b>); (<b>b</b>) <sup>29</sup>Si NMR spectrum of CEST-Al-MIL-53 (<b>2</b>); (<b>c</b>) TGA curves of samples CEST-Al-MIL-53 (<b>1</b>–<b>3</b>) and Pd/CEST-Al-MIL-53 (<b>5</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>) N<sub>2</sub> physisorption isotherms of CEST-Al-MIL-53 (<b>1</b>–<b>3</b>) and Al-MIL-53 (<b>4</b>) activated at different temperatures. (<b>b</b>) N<sub>2</sub> physisorption isotherms of Pd/CEST-Al-MIL-53 (<b>5</b>) and Pd/Al-MIL-53 (<b>6</b>).</p>
Full article ">Figure 5
<p>(<b>a</b>) SEM and (<b>b</b>) TEM images of sample Pd/CEST-Al-MIL-53 (<b>5</b>). The inset is HRTEM images in (<b>b</b>). (<b>c</b>) Scanning electron microscopy and EDX spectroscopic elemental mappings of C, O, Si, Al, and Pd in sample <b>5</b>.</p>
Full article ">Figure 6
<p>(<b>a</b>) XPS survey scan spectrum and high-resolution XPS spectra of sample Pd/CEST-Al-MIL-53 (<b>5</b>) for (<b>b</b>) Al 2p, (<b>c</b>) C 1s, (<b>d</b>) O 1s, (<b>e</b>) Si 2p, (<b>f</b>) Pd 3d, and (<b>g</b>) Cl 2p.</p>
Full article ">Figure 7
<p>(<b>a</b>) Time-dependent catalytic performance of catalysts Pd/CEST-Al-MIL-53 (<b>5</b>) and Pd/Al-MIL-53 (<b>6</b>) for the C-C cross-coupling reaction of iodobenzene and phenylboronic acid in the presence of K<sub>2</sub>CO<sub>3</sub> in ethanol; (<b>b</b>) the fitted plots of −ln(C<sub>t</sub>/C<sub>0</sub>) versus initial reaction time in the range of 0.5~5 h for catalysts <b>5</b> and <b>6</b>.</p>
Full article ">Figure 8
<p>The catalytic conversion of Suzuki–Miyaura C-C cross-coupling reaction of iodobenzene and phenylboronic acid with number of reuse cycles over Pd/CEST-Al-MIL-53 (<b>5</b>) and Pd/Al-MIL-53 (<b>6</b>).</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>c</b>) SEM and (<b>b</b>,<b>d</b>) TEM images of the sample Pd/CEST-Al-MIL-53 (<b>5</b>) after the fifth recycled run. (<b>e</b>–<b>i</b>) The energy-dispersive X-ray spectroscopic elemental mappings of Al, O, Pd, and Si according to image (<b>c</b>).</p>
Full article ">Figure 10
<p>(<b>a</b>) SEM and (<b>b</b>) TEM image of Pd-Al-MIL-53 after third recycling run. (<b>c</b>) SEM image of Pd/Al-MIL-53 before catalysis.</p>
Full article ">
14 pages, 1414 KiB  
Article
Silica-Nanocoated Membranes with Enhanced Stability and Antifouling Performance for Oil-Water Emulsion Separation
by Mengfan Zhu, Chengqian Huang and Yu Mao
Membranes 2025, 15(2), 41; https://doi.org/10.3390/membranes15020041 - 1 Feb 2025
Viewed by 171
Abstract
Despite the potential of glass fiber (GF) membranes for oil-water emulsion separations, efficient surface modification methods to enhance fouling resistance while preserving membrane performance and stability remain lacking. We report a silica nanocoating method to modify GF membranes through a vapor deposition method. [...] Read more.
Despite the potential of glass fiber (GF) membranes for oil-water emulsion separations, efficient surface modification methods to enhance fouling resistance while preserving membrane performance and stability remain lacking. We report a silica nanocoating method to modify GF membranes through a vapor deposition method. The high smoothness (<1 nm r.m.s.) and high conformality of the vapor-deposited silica nanocoatings enabled the preservation of membrane microstructure and permeability, which, combined with the enhanced surface hydrophilicity, led to an oil rejection rate exceeding 99% and more than a 40% improvement in permeate flux in oil-water emulsion separations. Furthermore, the silica nanocoatings provided the membranes with excellent wet strength and stability against organic solvents, strong acids, oxidants, boiling, and sonication. The silica-nanocoated membrane demonstrated enhanced fouling resistance, achieving flux recovery higher than 75% during repeated oil-water emulsion separations and bovine serum albumin and humic acid fouling tests. Full article
(This article belongs to the Special Issue Membrane Separation and Water Treatment: Modeling and Application)
16 pages, 2180 KiB  
Article
A Precipitation-Based Process to Generate a Solid Formulation of a Therapeutic Monoclonal Antibody: An Alternative to Lyophilization
by Athanas A. Koynov, Wei Lin, Jameson R. Bothe, Luke Schenck, Bibek Parajuli, Zhao Li, Richard Ruzanski, Natalie Hoffman, Derek Frank and Zachary VanAernum
J. Pharm. BioTech Ind. 2025, 2(1), 2; https://doi.org/10.3390/jpbi2010002 - 31 Jan 2025
Viewed by 335
Abstract
Lyophilization, or freeze-drying, is the default technique for the manufacture of solid-state formulations of therapeutic proteins. This established method offers several advantages, including improved product stability by minimizing chemical degradation, reduced storage requirements through water removal, and elimination of cold chain dependence. However, [...] Read more.
Lyophilization, or freeze-drying, is the default technique for the manufacture of solid-state formulations of therapeutic proteins. This established method offers several advantages, including improved product stability by minimizing chemical degradation, reduced storage requirements through water removal, and elimination of cold chain dependence. However, the lyophilization process itself presents limitations. It is a lengthy, batch-based operation, potentially leading to product inconsistencies and high manufacturing costs. Additionally, some proteins are susceptible to structural alterations during the freezing step, impacting their biological activity. This paper presents an alternative approach based on the co-precipitation of protein and excipients using an organic solvent. We explore the impact of various processing parameters on the viability of the formulation. We also provide an extensive characterization of proteins reconstituted from precipitated formulations and compare protein stability in solution and in lyophilized and precipitated solid formulations under long-term, accelerated, and stressed storage conditions. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Typical precipitated powders correspond to the three repeat runs for the experimental conditions 1′.</p>
Full article ">Figure 2
<p>Reproducibility of the precipitation process: a comparison of critical quality attributes of three lots of precipitated formulation.</p>
Full article ">Figure 3
<p>Comparison of secondary, tertiary, and higher-order structures of the reconstituted precipitated solid formulation (PPT) and mAb in solution (mAb X): (<b>A</b>) secondary structure via far UV CD; (<b>B</b>) tertiary structure via near UV CD; (<b>C</b>) tertiary structure via Intrinsic Tryptophan Fluorescence; and (<b>D</b>) higher-order structure via Dynamic Light Scattering. (<b>E</b>) The thermal stability of reconstituted precipitated solid formulation (PPT) and mAb in solution (mAb X) were determined using the DSC method. Changes in heat capacity (ΔCp) were plotted against different temperatures.</p>
Full article ">Figure 4
<p>High-level assessment of the results of the precipitations in the process characterization experiments. * Corresponds to a different solvent-to-anti-solvent ration (3×); ′ Corresponds to the initial triplicate run.</p>
Full article ">Figure 5
<p>Comparison of % of basic variants (<b>A</b>), % of HMW species (<b>B</b>), % of acidic variants (<b>C</b>), % of LMW species (<b>D</b>), and % of total impurity (<b>E</b>) of precipitated mAb X (•) and mAb X in a solution formulation (◦) stored at 5 °C (black) and 25 °C (green) for 3 months.</p>
Full article ">Figure 6
<p>Comparison of % of basic variants (<b>A</b>), % of HMW species (<b>B</b>), and % of acidic variants (<b>C</b>) of precipitated mAb X (•) and mAb X in a lyophilized formulation (□) at 5 °C (black), 25 °C (green), and 40 °C (red) for 3 months.</p>
Full article ">
17 pages, 4415 KiB  
Article
High Internal Phase Oil-in-Water Emulsions Stabilised by Cost-Effective Rhamnolipid/Alginate Biocomplexes
by Ilona E. Kłosowska-Chomiczewska, Gabriela Burakowska, Paulina Żmuda-Trzebiatowska, Aleksandra Soukup, Iwona Rok-Czapiewska, Elżbieta Hallmann, Tetiana Pokynbroda, Olena Karpenko, Krystyna Mędrzycka and Adam Macierzanka
Molecules 2025, 30(3), 595; https://doi.org/10.3390/molecules30030595 - 28 Jan 2025
Viewed by 562
Abstract
A novel, cost-effective, partially purified biosurfactant in the form of a rhamnolipid biocomplex (RLBC) was investigated for its emulsifying properties. The RLBC was obtained through the cultivation of Pseudomonas sp. SP-17 on glycerol, followed by acidic precipitation, without the use of organic solvents [...] Read more.
A novel, cost-effective, partially purified biosurfactant in the form of a rhamnolipid biocomplex (RLBC) was investigated for its emulsifying properties. The RLBC was obtained through the cultivation of Pseudomonas sp. SP-17 on glycerol, followed by acidic precipitation, without the use of organic solvents for isolation or purification. Composed of rhamnolipids (RLs) and the exopolysaccharide alginate, RLBC exhibited emulsifying properties towards rapeseed oil comparable to those of purified RLs at concentrations as low as 0.15% (w/w), sufficient for the effective stabilisation of oil-in-water (o/w) high internal phase emulsions (HIPEs, 80% oil). Dynamic light scattering analysis revealed similar droplet sizes (9.54 ± 0.96 µm for RLBC vs. 8.93 ± 0.58 µm for RLs), while multiple light scattering confirmed high emulsion stability over 120 days. The emulsions displayed shear-thinning behaviour, with yield stresses of approximately 11.5 Pa and 7.7 Pa for systems prepared with RLBC and RLs, respectively, after seven days of pre-storage. Although increasing the RLBC concentration from 0.15% to 1% (w/w) slightly improved the degree of emulsion dispersion, it did not substantially impact the long-term stability observed at the lowest concentration. Biodegradation tests demonstrated that the RLBC preparations are environmentally friendly alternatives to synthetic surfactants, achieving 60% biodegradation within 2.5 days and complete biodegradation within 14 days, which outperformed synthetic emulsifiers. The RLBC offers both environmental and economic advantages over purified RLs, including reduced production costs and the elimination of organic solvents. Our findings highlight the potential of RLBC for stabilising HIPEs in applications requiring sustainable and biodegradable formulations, such as cosmetics, lubricants, and industrial fluids widely manufactured and utilised today. Full article
(This article belongs to the Collection Advances in Food Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Surface tension isotherms for various RL preparations: RLs alone and their mixtures with alginate at 1.3:1 and 1:1 <span class="html-italic">w</span>/<span class="html-italic">w</span> ratios (model RLBCs). Isotherms are plotted against the concentration of the RL preparation as a whole (<b>A</b>) or recalculated based solely on the RL content within the RL preparation (<b>B</b>).</p>
Full article ">Figure 2
<p>Backscattering (BS) analysis of emulsions stabilised by 0.15% (<span class="html-italic">w</span>/<span class="html-italic">w</span>) of RLBC and containing 10% (<b>A</b>), 70% (<b>B</b>), 80% (<b>C</b>), or 90% (<b>D</b>) of oil phase. Visible droplet migration instability (such as creaming) of emulsions containing up to 70% oil phase (e.g., BS peak formation in the uppermost location of the sample shown in (<b>A</b>) and phase separation in the emulsion with 90% oil (BS decrease over time and accumulation of remaining emulsion in the lower part of the sample shown in (<b>D</b>) can be seen)).</p>
Full article ">Figure 3
<p>Microscopic structure of HIPEs (80% oil) stabilised with 0.15% (<span class="html-italic">w</span>/<span class="html-italic">w</span>) of purified RLs (<b>A</b>) or the RLBC biosurfactant at different concentrations (<span class="html-italic">w</span>/<span class="html-italic">w</span>): 0.15% (<b>B</b>), 0.25% (<b>C</b>), and 1.0% (<b>D</b>).</p>
Full article ">Figure 4
<p>The time-dependent evolution of mean backscattering (BS<sub>5-40</sub>) for emulsions (80% oil phase) stabilised by 0.15% <span class="html-italic">w</span>/<span class="html-italic">w</span> of RLBC or purified RLs (<b>A</b>), or by the RLBC at different concentrations (<b>B</b>).</p>
Full article ">Figure 5
<p>The relative change in oil droplet diameter for emulsions (80% oil phase) stabilised by 0.15% <span class="html-italic">w</span>/<span class="html-italic">w</span> RLBC or purified RLs (<b>A</b>), or by RLBC at different concentrations (<b>B</b>).</p>
Full article ">Figure 6
<p>Storage time-dependent rheological properties of rapeseed oil emulsions (80%) stabilised with 0.15% RLBC (<b>A</b>,<b>B</b>) or purified RLs (<b>C</b>,<b>D</b>): viscosity (<b>A</b>,<b>C</b>) and flow curves (<b>B</b>,<b>D</b>) at ascending (solid lines) and descending (dashed lines) shear rates.</p>
Full article ">Figure 7
<p>Time-dependent biodegradation rate for RLBC, RLs, and alginate.</p>
Full article ">
20 pages, 4016 KiB  
Article
Optimization of Green Ultrasound-Assisted Extraction of Carotenoids and Tocopherol from Tomato Waste Using NADESs
by Georgiana Ileana Badea, Florentina Gatea, Simona Carmen Litescu-Filipescu, Andreia Alecu, Ana Chira, Celina Maria Damian and Gabriel Lucian Radu
Molecules 2025, 30(3), 591; https://doi.org/10.3390/molecules30030591 - 28 Jan 2025
Viewed by 464
Abstract
The purpose of this study was to extract the lipophilic fraction from one of the largest source of waste in the industrial sector, namely, the tomato residue from processing the fruit. In order to make this process more environmentally sustainable, this study used [...] Read more.
The purpose of this study was to extract the lipophilic fraction from one of the largest source of waste in the industrial sector, namely, the tomato residue from processing the fruit. In order to make this process more environmentally sustainable, this study used a green extraction protocol employing natural deep eutectic solvents (NADESs) combined with a less energy-consuming technology, the ultrasound-assisted extraction (UAE) method, to simultaneously recover carotenoids and tocopherol from dried powder tomato waste. Two NADESs, one hydrophilic and one hydrophobic, were prepared and compared to support high extraction efficiency and increase the stability of the extracted compounds. The optimal extraction parameters were identified as choline chloride:1,3-butanediol (1:5)-based NADES, a solid-to-liquid ratio of 1:20 (w/v), time of extraction 12 min, temperature 65 °C, radiation frequency 37 Hz, and an ultrasound power level of 70%. The extraction process was intensified and resulted in extracts rich in lycopene (215.13 ± 4.31 μg/g DW), β-carotene (206.95 ± 3.27 μg/g DW), and tocopherol (130.86 ± 8.97 μg/g DW) content, with the highest antioxidant capacity 93.84 ± 0.18 mM Trolox equivalent. Incorporating NADESs for the extraction of bioactive compounds offers numerous benefits, such as improved sustainability, enhanced extraction efficiency, better protection of sensitive compounds, and reduced environmental impact. These advantages make NADESs a promising alternative to traditional organic solvents, especially in industries that require natural, green, and efficient extraction processes for valuable bioactive molecules. Full article
(This article belongs to the Section Green Chemistry)
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of pure choline chloride (black spectrum), 1,3-butanediol (red spectrum), and the prepared NADES-1 (green spectrum).</p>
Full article ">Figure 2
<p>FTIR spectra of pure menthol (black spectrum), oleic acid (red spectrum), and the prepared NADES-2 (green spectrum).</p>
Full article ">Figure 3
<p>Thermogram of NADES 1,3-BD–ChCl (1:5). The x-axis shows the increase in temperature (°C), while the y-axis shows the loss in weight (%).</p>
Full article ">Figure 4
<p>HPLC profile of the major carotenoids (1-lutein, 2-lycopene, 3-β-carotene, at 458 nm) and tocopherol (4-inset, at 293 nm) in tomato waste extracts using NADES-1 (<b>A</b>) and NADES-2 (<b>B</b>) as extraction solvents: 30 min time of extraction, temperature 55 °C, ultrasound power 100 W, radiation frequency 37 Hz, and S/L 1:15 (<span class="html-italic">w</span>/<span class="html-italic">v</span>).</p>
Full article ">Figure 5
<p>Comparison of HPLC chromatograms for different extraction solvents—sunflower oil (<b>A</b>) andNADES-2 (<b>B</b>)—with the chromatogram of a mixture of standards (<b>C</b>) at 458 nm (1-Lutein, 2-Lycopene, 3-β-carotene).</p>
Full article ">Figure 6
<p>Optimization of carotenoid extraction from tomato waste using different parameters (NADES component molar ratio, solid-to-solvent ratio, time of extraction, extraction temperature, and sonication power). The extraction conditions are given in <a href="#molecules-30-00591-t002" class="html-table">Table 2</a>, <a href="#molecules-30-00591-t003" class="html-table">Table 3</a>, <a href="#molecules-30-00591-t004" class="html-table">Table 4</a>, <a href="#molecules-30-00591-t005" class="html-table">Table 5</a> and <a href="#molecules-30-00591-t006" class="html-table">Table 6</a>.</p>
Full article ">Figure 7
<p>Dried tomato powder used for UAE extraction with NADESs.</p>
Full article ">Figure 8
<p>The investigated mixtures at room temperature (compositions defined in <a href="#molecules-30-00591-t010" class="html-table">Table 10</a>, NADES-1 left, NADES-2 right).</p>
Full article ">
36 pages, 6468 KiB  
Review
Sustainable Extraction of Critical Minerals from Waste Batteries: A Green Solvent Approach in Resource Recovery
by Afzal Ahmed Dar, Zhi Chen, Gaixia Zhang, Jinguang Hu, Karim Zaghib, Sixu Deng, Xiaolei Wang, Fariborz Haghighat, Catherine N. Mulligan, Chunjiang An, Antonio Avalos Ramirez and Shuhui Sun
Batteries 2025, 11(2), 51; https://doi.org/10.3390/batteries11020051 - 28 Jan 2025
Viewed by 617
Abstract
This strategic review examines the pivotal role of sustainable methodologies in battery recycling and the recovery of critical minerals from waste batteries, emphasizing the need to address existing technical and environmental challenges. Through a systematic analysis, it explores the application of green organic [...] Read more.
This strategic review examines the pivotal role of sustainable methodologies in battery recycling and the recovery of critical minerals from waste batteries, emphasizing the need to address existing technical and environmental challenges. Through a systematic analysis, it explores the application of green organic solvents in mineral processing, advocating for establishing eco-friendly techniques aimed at clipping waste and boosting resource utilization. The escalating demand for and shortage of essential minerals including copper, cobalt, lithium, and nickel are comprehensively analyzed and forecasted for 2023, 2030, and 2040. Traditional extraction techniques, including hydrometallurgical, pyrometallurgical, and bio-metallurgical processes, are efficient but pose substantial environmental hazards and contribute to resource scarcity. The concept of green extraction arises as a crucial step towards ecological conservation, integrating sustainable practices to lessen the environmental footprint of mineral extraction. The advancement of green organic solvents, notably ionic liquids and deep eutectic solvents, is examined, highlighting their attributes of minimal toxicity, biodegradability, and superior efficacy, thus presenting great potential in transforming the sector. The emergence of organic solvents such as palm oil, 1-octanol, and Span 80 is recognized, with advantageous low solubility and adaptability to varying temperatures. Kinetic (mainly temperature) data of different deep eutectic solvents are extracted from previous studies and computed with machine learning techniques. The coefficient of determination and mean squared error reveal the accuracy of experimental and computed data. In essence, this study seeks to inspire ongoing efforts to navigate impediments, embrace technological advancements including artificial intelligence, and foster an ethos of environmental stewardship in the sustainable extraction and recycling of critical metals from waste batteries. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Rechargeable batteries by market size, 2022–2023 (billion USD), and (<b>B</b>) rechargeable batteries by battery type, 2022.</p>
Full article ">Figure 2
<p>Impact of mineral extraction and environmental deterioration, emission of gases during mining and battery production, and ubiquitous consumption of electronic appliances.</p>
Full article ">Figure 3
<p>(<b>a</b>) Critical minerals in different aspects including demand, mining, and refining minerals in 2023, 2030, and 2040, (<b>b</b>) shortage of minerals, which is calculated by demand and mining parameters, and (<b>c</b>) refinery addition of minerals to fulfill the demand.</p>
Full article ">Figure 4
<p>Hydrometallurgical process comprising the pretreatment of negative-electrode and positive-electrode materials followed by potential leaching and mineral extraction.</p>
Full article ">Figure 5
<p>Pyrometallurgical recycling layout to recover the minerals.</p>
Full article ">Figure 6
<p>Innovations in ionic liquid: (<b>a</b>) common polymeric, acidic, ionic liquids, (<b>b</b>) polymeric, acidic, ionic liquid synthesis process [<a href="#B156-batteries-11-00051" class="html-bibr">156</a>], and (<b>c</b>) sulfonic-acid-functionalized acidic, ionic liquid synthesis by grafting [<a href="#B157-batteries-11-00051" class="html-bibr">157</a>,<a href="#B158-batteries-11-00051" class="html-bibr">158</a>]. Reprinted with permission from reference [<a href="#B158-batteries-11-00051" class="html-bibr">158</a>]. Copyright 2016, American Chemical Society.</p>
Full article ">Figure 7
<p>Deep eutectic solvents with their five different types.</p>
Full article ">Figure 8
<p>Mineral extraction with DES-based leaching, initiating with choline chloride as the hydrogen bond acceptor and urea as the hydrogen bond donor.</p>
Full article ">Figure 9
<p>EDS analysis and SEM images from before (<b>A</b>) and after (<b>B</b>–<b>D</b>) the leaching process: (<b>A</b>) ChCl + urea, (<b>B</b>) ChCl + EG, (<b>C</b>) ChCl + urea + EG, (<b>D</b>) DES: (ChCl:urea 1:2, ChCl:EG 1:2, and ChCl:urea:EG 1:2:1 at 100 °C) [<a href="#B188-batteries-11-00051" class="html-bibr">188</a>]. Reprinted with permission from reference [<a href="#B188-batteries-11-00051" class="html-bibr">188</a>], Copyright 2022, Elsevier.</p>
Full article ">Figure 10
<p>Effect of leaching temperature on mineral efficiency of three different types of DESs—ChCl, urea, and EG—and combination of urea and EG. (<b>A</b>) Percentage Li recovery, (<b>B</b>) percentage Co recovery, (<b>C</b>) percentage Ni recovery, and (<b>D</b>) percentage manganese recovery. Source: [<a href="#B99-batteries-11-00051" class="html-bibr">99</a>,<a href="#B102-batteries-11-00051" class="html-bibr">102</a>,<a href="#B103-batteries-11-00051" class="html-bibr">103</a>].</p>
Full article ">Figure 11
<p>Percentage concentrations of respective metal combinations of DESs with fluctuating temperature (50, 75, and 100 °C), where black, blue, and red represent actual, linear fit, and quadratic fit values, respectively.</p>
Full article ">
14 pages, 3379 KiB  
Article
Recovery and Reuse of Acetone from Pharmaceutical Industry Waste by Solar Distillation
by Eva Carina Tarango Brito, Carlos Eduardo Barrera Díaz, Liliana Ivette Ávila Córdoba, Bernardo Antonio Frontana Uribe and Dora Alicia Solís Casados
Processes 2025, 13(2), 361; https://doi.org/10.3390/pr13020361 - 28 Jan 2025
Viewed by 516
Abstract
Solvents are particularly hazardous among the mixture of pollutants found in the air, as their low vapor pressure allows them to reach the atmosphere, causing damage to ecosystems, and producing secondary deleterious effects on living organisms through a wide variety of possible reactions. [...] Read more.
Solvents are particularly hazardous among the mixture of pollutants found in the air, as their low vapor pressure allows them to reach the atmosphere, causing damage to ecosystems, and producing secondary deleterious effects on living organisms through a wide variety of possible reactions. In response, innovative, sustainable, and ecological methods are being developed to recover solvents from industrial wastewater, which is typically contaminated with other organic compounds. This study describes the procedure for recovering acetone from a residue from the pharmaceutical industry. This compound contains a high amount of solid organic compounds, which are generated during the manufacture of medicines. The treatment consisted of performing a simple solar distillation using a single-slope glass solar still, which separated the acetone from the mother solution. Under ideal circumstances, the use of solar radiation allowed an efficiency rate of 80% using solar concentration by means of mirrors to increase the temperature and 85% without the use of mirrors in the production of distilled acetone, which was characterized to evaluate its quality using instrumental analytical techniques: NMR, IR, and GC. The results obtained indicate that the acetone recovered by this procedure has a good quality of 84%; however, due to this percentage obtained, its reuse is limited for certain applications where a high degree of purity is required, such as its reuse for pharmaceutical use; for this reason, it was proposed to use said compound to eliminate the organic impurities contained in the catalyst waste granules used in a Mexican oil refinery. The resulting material was examined by SEM and EDS, revealing a high initial carbon content that decreased by 29% after treatment. Likewise, as an additional study, a study was carried out to evaluate the characteristics of the residues obtained at the end of the distillation where rubidium, silicon, carbon, nitrogen, oxygen, and chlorine contents were observed. Full article
(This article belongs to the Section Chemical Processes and Systems)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Solar still: (<b>a</b>) without mirror adaptation, and (<b>b</b>) with mirrors as solar concentrators.</p>
Full article ">Figure 2
<p>The determination of the temperature in the solar still without mirror adaptation (˗) and with mirrors (- - -) as solar concentrators.</p>
Full article ">Figure 3
<p>Nuclear magnetic resonance analysis of distilled acetone: (<b>a</b>) <sup>1</sup>H NMR, 300 MHz; (<b>b</b>) <sup>13</sup>C NMR, 75 MHz.</p>
Full article ">Figure 4
<p>IR absorption spectrum obtained for distilled acetone.</p>
Full article ">Figure 5
<p>Gas chromatography of distilled acetone.</p>
Full article ">Figure 6
<p>(<b>a1</b>,<b>a2</b>) The morphological analysis of the catalyst before and after the impurity removal process using distilled acetone by SEM. (<b>b1</b>,<b>b2</b>) The EDS spectrum of the catalyst before and after the impurity removal process using distilled acetone.</p>
Full article ">Figure 7
<p>The wide spectrum of the solid from distillation.</p>
Full article ">Figure 8
<p>(<b>a</b>) Deconvoluted rubidium, (<b>b</b>) silicon, (<b>c</b>) carbon, and (<b>d</b>) oxygen region spectrums. Note: Different color lines were used in order to get differences between the several binding energies deconvoluted in each spectrum.</p>
Full article ">Figure 8 Cont.
<p>(<b>a</b>) Deconvoluted rubidium, (<b>b</b>) silicon, (<b>c</b>) carbon, and (<b>d</b>) oxygen region spectrums. Note: Different color lines were used in order to get differences between the several binding energies deconvoluted in each spectrum.</p>
Full article ">Figure 9
<p>The EDS spectrum of the solid from distillation.</p>
Full article ">Figure 10
<p>The XRD analysis of the solid from distillation.</p>
Full article ">
11 pages, 490 KiB  
Article
Environmentally Friendly Green O-Alkylation Reaction for Ethenzamide Synthesis
by Paulina Niedziejko-Ćwiertnia, Anna Karolina Drabczyk, Damian Kułaga, Patrycja Podobińska, Wojciech Bachowski, Kamila Zeńczak-Tomera, Piotr Michorczyk, Ruilong Sheng and Jolanta Jaśkowska
Appl. Sci. 2025, 15(3), 1342; https://doi.org/10.3390/app15031342 - 27 Jan 2025
Viewed by 449
Abstract
Ethenzamide (2-ethoxybenzamide), besides acetylsalicylic acid, is one of the mostly used salicylic acid derivatives in pharmaceuticals. It has analgesic and anti-inflammatory effects that originate from the inhibition of cyclooxygenase (COX-1) activity, thus blocking prostaglandin synthesis. In this work, efficient and eco-friendly methods were [...] Read more.
Ethenzamide (2-ethoxybenzamide), besides acetylsalicylic acid, is one of the mostly used salicylic acid derivatives in pharmaceuticals. It has analgesic and anti-inflammatory effects that originate from the inhibition of cyclooxygenase (COX-1) activity, thus blocking prostaglandin synthesis. In this work, efficient and eco-friendly methods were developed for the synthesis of ethenzamide via the O-alkylation reaction of salicylamide. The reactions were carried out under conventional conditions in a solvent-free system using variant solvents and different phase transfer catalysts (PTC) in the presence of microwave radiation or ultrasonic conditions. It was shown that in solvent-free conditions using TBAB as a catalyst, ethenzamide can be obtained within 15 min at 80 °C with 79% yield. Meanwhile, using microwave radiation under the same conditions, the reaction time can be shortened to 90 s with 92% yield. Notably, high yields can be achieved under PTC in water (or organic solvent-free) conditions using microwave radiation (2 min, 94%) or ultrasound (10 min, 95% efficiency). The studies prove that the PTC synthesis process of ethenzamide can be conducted under mild conditions, with a shorter reaction time and remarkably lower energy consumption in comparison to conventional processes, thus actualizing “green chemistry” for practical ethenzamide preparation. Full article
(This article belongs to the Special Issue Advances in Organic Synthetic Chemistry)
27 pages, 7929 KiB  
Review
Recent Progress of Chemical Reactions Induced by Contact Electrification
by Xinyi Huo, Shaoxin Li, Bing Sun, Zhong Lin Wang and Di Wei
Molecules 2025, 30(3), 584; https://doi.org/10.3390/molecules30030584 - 27 Jan 2025
Viewed by 686
Abstract
Contact electrification (CE) spans from atomic to macroscopic scales, facilitating charge transfer between materials upon contact. This interfacial charge exchange, occurring in solid–solid (S–S) or solid–liquid (S–L) systems, initiates radical generation and chemical reactions, collectively termed contact-electro-chemistry (CE-Chemistry). As an emerging platform for [...] Read more.
Contact electrification (CE) spans from atomic to macroscopic scales, facilitating charge transfer between materials upon contact. This interfacial charge exchange, occurring in solid–solid (S–S) or solid–liquid (S–L) systems, initiates radical generation and chemical reactions, collectively termed contact-electro-chemistry (CE-Chemistry). As an emerging platform for green chemistry, CE-Chemistry facilitates redox, luminescent, synthetic, and catalytic reactions without the need for external power sources as in traditional electrochemistry with noble metal catalysts, significantly reducing energy consumption and environmental impact. Despite its broad applicability, the mechanistic understanding of CE-Chemistry remains incomplete. In S–S systems, CE-Chemistry is primarily driven by surface charges, whether electrons, ions, or radicals, on charged solid interfaces. However, a comprehensive theoretical framework is yet to be established. While S–S CE offers a promising platform for exploring the interplay between chemical reactions and triboelectric charge via surface charge modulation, it faces significant challenges in achieving scalability and optimizing chemical efficiency. In contrast, S–L CE-Chemistry focuses on interfacial electron transfer as a critical step in radical generation and subsequent reactions. This approach is notably versatile, enabling bulk-phase reactions in solutions and offering the flexibility to choose various solvents and/or dielectrics to optimize reaction pathways, such as the degradation of organic pollutants and polymerization, etc. The formation of an interfacial electrical double layer (EDL), driven by surface ion adsorption following electron transfer, plays a pivotal role in CE-Chemical processes within aqueous S–L systems. However, the EDL can exert a screening effect on further electron transfer, thereby inhibiting reaction progress. A comprehensive understanding and optimization of charge transfer mechanisms are pivotal for elucidating reaction pathways and enabling precise control over CE-Chemical processes. As the foundation of CE-Chemistry, charge transfer underpins the development of energy-efficient and environmentally sustainable methodologies, holding transformative potential for advancing green innovation. This review consolidates recent advancements, systematically classifying progress based on interfacial configurations in S–S and S–L systems and the underlying charge transfer dynamics. To unlock the full potential of CE-Chemistry, future research should prioritize the strategic tuning of material electronegativity, the engineering of sophisticated surface architectures, and the enhancement of charge transport mechanisms, paving the way for sustainable chemical innovations. Full article
Show Figures

Figure 1

Figure 1
<p>CE-Chemistry in S–S and S–L CE.</p>
Full article ">Figure 2
<p>(<b>a</b>) Energy bands between a metal and a dielectric, between two different dielectrics, and between identical dielectrics during contact–separation process (reprinted with permission from Ref. [<a href="#B62-molecules-30-00584" class="html-bibr">62</a>]. 2024, Springer). (<b>b</b>) Schematic of the electron cloud and potential energy profile of two atoms belonging to two materials A and B during CE (reprinted with permission from Ref. [<a href="#B63-molecules-30-00584" class="html-bibr">63</a>]. 2022, John Wiley and Sons). (<b>c</b>) The traditional view (uniformly positively or negatively) and the mosaic picture model (positively charged and negatively charged regions on the nanoscale) during contact–separation process, respectively (reprinted with permission from Ref. [<a href="#B45-molecules-30-00584" class="html-bibr">45</a>]. 2011, American Association for the Advancement of Science).</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic diagram of self-catalysis tribo-electrochemistry reaction process (reprinted with permission from Ref. [<a href="#B47-molecules-30-00584" class="html-bibr">47</a>]. 2024, Springer). (<b>b</b>) The charge signal acquired by the water in a Teflon beaker and in the copper vessel, respectively. Photograph of the Teflon beaker containing 100 mL water with a magnetic pellet covered by a copper foil and the copper vessel containing 50 mL water with a magnetic Teflon pellet. (<b>c</b>) Photographs of the reaction product of Teflon and gold in the presence of glucose under visible light and UV light (mercury vapor lamp) (reprinted with permission from Ref. [<a href="#B70-molecules-30-00584" class="html-bibr">70</a>]. 2019, American Chemical Society).</p>
Full article ">Figure 4
<p>(<b>a</b>) EDS of Cu deposited on charged Teflon. (<b>b</b>) Optical absorbance of 1 mM CuSO<sub>4</sub> solution before and after contact with charged Teflon. (<b>c</b>) Fe(CN)<sub>6</sub><sup>3−</sup> was reduced by charged Teflon to Fe(CN)<sub>6</sub><sup>4−</sup> (reprinted with permission from Ref. [<a href="#B71-molecules-30-00584" class="html-bibr">71</a>]. 2008, Springer Nature). (<b>d</b>) Relative intensity of chemiluminescence as a function of time when a PMMA-rubbed PE rod dipped into a MeCN/H<sub>2</sub>O (1:1, <span class="html-italic">v</span>/<span class="html-italic">v</span>) mixture containing 2.5 mM S<sub>2</sub>O<sub>8</sub><sup>2−</sup> and 0.25 mM Ru(bpy)<sub>3</sub><sup>2+</sup>. Insert shows the reaction mechanism (reprinted with permission from Ref. [<a href="#B74-molecules-30-00584" class="html-bibr">74</a>]. 2009, Elsevier).</p>
Full article ">Figure 5
<p>(<b>a</b>) Polymer pieces are charged by rubbing or pressing against one another and are then immersed in an aqueous reagent solution. (<b>b</b>) SEM image zooming on the Au NPs deposited on the surface of (+) contact-charged PDMS (scale bar = 1 μm). (<b>c</b>) UV−vis spectrum featuring a surface plasmon resonance (SPR) peak centered at 520 nm and characteristic of Au nanoparticles (reprinted with permission from Ref. [<a href="#B75-molecules-30-00584" class="html-bibr">75</a>]. 2012, American Chemical Society). (<b>d</b>) Schematic for the electrodeposition of silver nanoparticles on tribocharged undoped amorphous silicon (a-Si) samples. (<b>e</b>) SEM images of patterns of silver particles generated by reducing silver ions on a-Si samples tribocharged by rolling PVC and Teflon spheres (reprinted with permission from Ref. [<a href="#B33-molecules-30-00584" class="html-bibr">33</a>]. 2021, American Chemical Society).</p>
Full article ">Figure 6
<p>(<b>a</b>) Analysis of silver nanoparticles electrochemically grown on electrical insulators (tribocharged PDMS samples immersed in aqueous AgNO<sub>3</sub> solutions). (<b>b</b>) Schematic depiction of a “mosaic” ensemble of triboelectric charges on two samples of equal charge excess (reprinted with permission from Ref. [<a href="#B76-molecules-30-00584" class="html-bibr">76</a>]. 2019, American Chemical Society).</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic diagram of a unified mechanical model of contact charge charges (reprinted with permission from Ref. [<a href="#B86-molecules-30-00584" class="html-bibr">86</a>]. 2020, John Wiley and Sons). (<b>b</b>) Lin’s hybrid EDL model and the “two-step” process on its formation. In the first step, the molecules and ions in the liquid impact the solid surface due to the thermal motion and the pressure from the liquid, which leads to electron transfer between them; meanwhile, ions may also attach to the solid surface. In the second step, free ions in the liquid would be attracted to the electrified surface due to electrostatic interactions, forming an EDL (reprinted with permission from Ref. [<a href="#B87-molecules-30-00584" class="html-bibr">87</a>]. 2022, American Chemical Society).</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic image of fluorescence microscopy set up for imaging the microfluidic chip. (<b>b</b>) The relationship between fluorescence intensity and reaction time of sample resting in microfluidic chip. (<b>c</b>) The fluorescent profiles of the spontaneous generation of H<sub>2</sub>O<sub>2</sub> along the normal direction of the substrate (top, the optical microscopy image of a typical straight channel; middle, the corresponding fluorescence image; bottom, the corresponding fluorescence intensity) (reprinted with permission from Ref. [<a href="#B90-molecules-30-00584" class="html-bibr">90</a>]. 2022, American Chemical Society). (<b>d</b>) When a drop of luminol solution flowed through the Teflon surface, the electron transfer between luminol and solid occurred, and then the statically charged luminol was dropped into the cuvette containing the catalyst, leading to light blue emission. (<b>e</b>) Schematic depiction of the electron transfer between a liquid droplet and the Teflon interface and the ions adsorbed due to the Coulombic attraction. After separation, the positively charged luminol droplet attracted anions at the interface. (<b>f</b>) Schematic depiction of the electrons from the negatively charged luminol droplet competing with luminol for Fe<sup>3+</sup> (reprinted with permission from Ref. [<a href="#B91-molecules-30-00584" class="html-bibr">91</a>]. 2022, American Chemical Society).</p>
Full article ">Figure 9
<p>(<b>a</b>) Charge-induced peptide oxidation by S–L contact. (<b>b</b>) Photographs of the experimental setup (reprinted with permission from Ref. [<a href="#B93-molecules-30-00584" class="html-bibr">93</a>]. 2023, American Chemical Society). (<b>c</b>) Mechanism of free radical generation by CE process (reprinted with permission from Ref. [<a href="#B94-molecules-30-00584" class="html-bibr">94</a>]. 2023, Elsevier).</p>
Full article ">Figure 10
<p>(<b>a</b>) Schematic representation of the experimental setup and overall reaction. (<b>b</b>) Expected large application of S–L CE in the field of efficient production of H<sub>2</sub>O<sub>2</sub>. (<b>c</b>) Proposed reaction mechanism of CEC for H<sub>2</sub>O<sub>2</sub> generation (reprinted with permission from Ref. [<a href="#B71-molecules-30-00584" class="html-bibr">71</a>]. 2023, John Wiley and Sons). (<b>d</b>) Schematic illustration of generating H<sub>2</sub>O<sub>2</sub> from water and H<sub>2</sub>O by ultrasonication in the presence of FEP. (<b>e</b>) Illustration of the experimental setup comprising a thermostatic circulator that regulates the temperature of the reactor and an ultrasonic bath (40 kHz, 110 W). (<b>f</b>) Illustration of the proposed general principle of the mechanism of radicals’ generation by ultrasonication-induced CEC in aqueous solution (reprinted with permission from Ref. [<a href="#B73-molecules-30-00584" class="html-bibr">73</a>]. 2023, John Wiley and Sons).</p>
Full article ">Figure 11
<p>(<b>a</b>) Three-dimensional schematic of the experimental setup and protocol. (<b>b</b>) Proposed mechanism for the degradation of MO by CEC-generated radicals. (<b>c</b>) Graphical description of the experimental set-up and observations. (<b>d</b>) Schematic description of the reduction of various metal ions in aqueous solution by ultrasonically driven CEC in presence of FEP (reprinted with permission from Ref. [<a href="#B54-molecules-30-00584" class="html-bibr">54</a>]. 2024, Elsevier). (<b>e</b>) Copper plated on untreated PMMA surface. (Upper left) An optical image of copper film deposited on a PMMA tube inner surface. No Cu was seen on the outer surface that had been previously contacted with Teflon. (Bottom) Enlarged image of the copper film (0.8 mm × 0.6 mm) where features such as lines reflect the surface structure of the tube instead of scratches. (Upper right) Images of two glass test tubes containing copper plating solution before and after the plating. The blue color of the solution disappeared following the deposition (reprinted with permission from Ref. [<a href="#B95-molecules-30-00584" class="html-bibr">95</a>]. 2009, American Chemical Society).</p>
Full article ">Figure 12
<p>A comprehensive paradigm of CE-Chemistry reactions. (<b>a</b>) Schematic diagram of Fe<sup>2+</sup> being oxidized to Fe<sup>3+</sup>. (<b>b</b>) Schematic diagram of the structure of aniline polymerization. (<b>c</b>) UV–vis spectra of solution samples at different times. (<b>d</b>) Working mechanism of CE oxidation reaction of Fe(CN)<sub>6</sub><sup>4−</sup> to Fe(CN)<sub>6</sub><sup>3−</sup> (reprinted with permission from Ref. [<a href="#B55-molecules-30-00584" class="html-bibr">55</a>]. 2024, Elsevier). (<b>e</b>) Guidance that unified the concept of work functions, triboelectric series, and standard electrode potentials by electron transfer capability. (<b>f</b>) Complete schematic of luminol CEL. The processes (<b>i</b>–<b>iv</b>) upper right corner of the figure represent the formation, growth, and collapse of cavitation bubbles provoked by the propagation of ultrasonic waves in solution (reprinted with permission from Ref. [<a href="#B31-molecules-30-00584" class="html-bibr">31</a>]. 2024, Elsevier). (<b>g</b>) Schematic diagram of the chemical reaction for the degradation of phenol in DMSO by CE (reprinted with permission from Ref. [<a href="#B97-molecules-30-00584" class="html-bibr">97</a>]. 2024, American Chemical Society).</p>
Full article ">Figure 13
<p>Other ways in which exposure electrocatalysis induces chemical reactions. (<b>a</b>) Schematic illustration of a ball mill process using triboelectric materials (reprinted with permission from Ref. [<a href="#B101-molecules-30-00584" class="html-bibr">101</a>]. 2024, Springer). (<b>b</b>) The designed catalysis unit including stator and rotator. The production process of ·OH and ·O<sub>2</sub> and final degradation products of CV molecules. Optical photographs of 20 mg L<sup>−1</sup> CV solution samples from 0 s to 90 s (reprinted with permission from Ref. [<a href="#B102-molecules-30-00584" class="html-bibr">102</a>]. 2023, Elsevier). (<b>c</b>) CE-Chemistry for K<sub>4</sub>[Fe(CN)<sub>6</sub>] solution and FEP tubes (reprinted with permission from Ref. [<a href="#B31-molecules-30-00584" class="html-bibr">31</a>]. 2024, Elsevier).</p>
Full article ">Figure 14
<p>Characteristics of CE and future development direction.</p>
Full article ">
11 pages, 1502 KiB  
Article
Rapid and Efficient Synthesis of Succinated Thiol Compounds via Maleic Anhydride Derivatization
by Hiroshi Yamaguchi, Hikari Sugawa, Himeno Takahashi and Ryoji Nagai
Molecules 2025, 30(3), 571; https://doi.org/10.3390/molecules30030571 - 27 Jan 2025
Viewed by 307
Abstract
Succination is a non-enzymatic post-translational modification of cysteine (Cys) residues, resulting in the formation of S-(2-succino)cysteine (2SC). While hundreds of 2SC-modified proteins have been identified and are associated with the dysfunction of proteins, the underlying molecular mechanisms remain poorly understood. Conventional methods [...] Read more.
Succination is a non-enzymatic post-translational modification of cysteine (Cys) residues, resulting in the formation of S-(2-succino)cysteine (2SC). While hundreds of 2SC-modified proteins have been identified and are associated with the dysfunction of proteins, the underlying molecular mechanisms remain poorly understood. Conventional methods for synthesizing succinated compounds, such as 2SC, often require prolonged reaction times and/or HCl hydrolysis. In this study, we present a rapid and efficient synthesis method for succinated compounds using maleic anhydride, enabling more effective in vivo studies of succination mechanisms. This method was tested on thiol compounds with varying molecular weights, including Cys derivatives, Cys-containing peptides, and reduced bovine serum albumin. By incubating these compounds in an aqueous buffer with maleic anhydride dissolved in an organic solvent like diethyl ether, we achieved significantly improved succination efficiency compared to conventional methods. The succination efficiency using maleic anhydride surpassed that of fumaric acid or maleic acid. Notably, this approach facilitated the succination of amino acids, peptides, and proteins within minutes at 25 °C, without requiring acid hydrolysis. Our findings provide a straightforward, time-efficient strategy for synthesizing succinated thiol compounds, offering a valuable tool to enhance the understanding of succination’s molecular mechanisms and its role in protein function and dysfunction. Full article
(This article belongs to the Special Issue 10th Anniversary of the Bioorganic Chemistry Section of Molecules)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the succination reaction. When the nucleophilic thiol group attacks the electrophilic β-carbon atom of maleic anhydride, a thioether bond is formed through Michael addition. Subsequently, the anhydride portion undergoes hydrolysis, and a succinated compound is formed.</p>
Full article ">Figure 2
<p>Succination of Ac-Cys. (<b>a</b>) HPLC profile of Ac-Cys and its succinated product (Ac-2SC) with maleic anhydride in different solvents: control (no reaction); diethyl ether; ethyl acetate; water. (<b>b</b>) HPLC profile of the succination reaction using maleic anhydride, maleic acid, or fumaric acid dissolved in diethyl ether. All succination reactions were carried out at 25 °C for 30 min in 100 mM phosphate buffer (pH 7.0). The concentrations of Ac-Cys and carboxyl compounds were 0.2 μmol and 4 μmol, respectively.</p>
Full article ">Figure 3
<p>Characterization of succination using maleic anhydride. (<b>a</b>) Influence of the Ac-Cys to maleic anhydride ratio in 100 mM phosphate buffer (pH 7.0) at 25 °C for 30 min. (<b>b</b>) Time course of the succination of Ac-Cys in 100 mM phosphate buffer (pH 7.0) at 25 °C. (<b>c</b>) Effect of pH on the succination of Ac-Cys. The reaction was conducted at 25 °C for 30 min. In all experiments, maleic anhydride was dissolved in diethyl ether and used in the succination reaction. The concentrations of Ac-Cys and maleic anhydride were 0.2 μmol and 4 μmol, respectively. The graphs show the mean ± standard deviation of at least three experiments.</p>
Full article ">Figure 4
<p>2SC synthesis. The succination of Boc-Cys was carried out at 25 °C for 30 min in 200 mM phosphate buffer (pH 7.0). Maleic anhydride, dissolved in diethyl ether, was added in a 20-fold excess. The de-Boc reaction was performed in a 90% TFA aqueous solution.</p>
Full article ">
15 pages, 3228 KiB  
Article
“One Pot” Enzymatic Synthesis of Caffeic Acid Phenethyl Ester in Deep Eutectic Solvent
by Maria Roberta Tripon, Camelia Tulcan, Simona Marc, Dorin-Dumitru Camen and Cristina Paul
Biomolecules 2025, 15(2), 181; https://doi.org/10.3390/biom15020181 - 27 Jan 2025
Viewed by 502
Abstract
Caffeic acid phenethyl ester (CAPE) represents a valuable ester of caffeic acid, which, over time, has demonstrated remarkable pharmacological properties. In general, the ester is obtained in organic solvents, especially by the esterification reaction of caffeic acid (CA) and 2-phenylethanol (PE). In this [...] Read more.
Caffeic acid phenethyl ester (CAPE) represents a valuable ester of caffeic acid, which, over time, has demonstrated remarkable pharmacological properties. In general, the ester is obtained in organic solvents, especially by the esterification reaction of caffeic acid (CA) and 2-phenylethanol (PE). In this context, the purpose of this study was the use of the “one pot” system to synthesize CAPE through biocatalysis with various lipases in a choline-chloride-based DES system, employing the “2-in-1” concept, where one of the substrates functions as both reactant and solvent. The synthesis process of CAPE is contingent on the molar ratio between CA and PE; thus, this factor was the primary subject of investigation, with different molar ratios of CA and PE being studied. Furthermore, the impact of temperature, time, the nature of the biocatalyst, and the water loading of the DES system was also examined. This ‘green’ synthesis method, which has demonstrated encouraging reaction yields (%), could secure and maintain the therapeutic potential of CAPE, mainly due to the non-toxic character of the reaction medium. Full article
(This article belongs to the Section Molecular Medicine)
Show Figures

Figure 1

Figure 1
<p>Graphical workflow of CAPE synthesis by lipase biocatalysis in organic solvents.</p>
Full article ">Figure 2
<p>Graphical workflow of CAPE synthesis by lipase biocatalysis in deep eutectic solvent.</p>
Full article ">Figure 3
<p>Enzymatic synthesis of CAPE by lipase biocatalysis in different organic solvents.</p>
Full article ">Figure 4
<p>Ester yields (%) and biocatalyst efficiency (µmol/h/g) obtained for SR<sub>6</sub> and SR7 (in isooctane as solvent).</p>
Full article ">Figure 5
<p>Enzymatic synthesis of CAPE by lipase biocatalysis in deep eutectic solvent.</p>
Full article ">Figure 6
<p>Ester yields (%) and catalytic efficiency (µmoles/h/g) obtained for reactions in DES at 80 °C and different molar ratio of CA:PE, catalyzed by immobilized <span class="html-italic">A. niger</span> lipase (AnL-IM).</p>
Full article ">Figure 7
<p>Ester yields (%) and catalytic efficiency (µmoles/h/g) obtained for reactions in DES at 70 °C with the use of different biocatalysts (CaLB-n: <span class="html-italic">C. antarctica</span> lipase B native; CaLB-IM: <span class="html-italic">C. antarctica</span> lipase B, immobilized; PsL-n: <span class="html-italic">P. stutzeri</span> lipase native; PsL-SG2: <span class="html-italic">P. stutzeri</span> lipase sol–gel, immobilized).</p>
Full article ">Figure 8
<p>Ester yields (%) and catalytic efficiency (µmoles/h/g) obtained for reactions in DES at 80 °C with water addition (1%, 2.5%, and 5%), no water added (NW), and <span class="html-italic">A. niger</span> lipase, immobilized (AnL-IM), as biocatalyst.</p>
Full article ">
17 pages, 4405 KiB  
Article
Chemical Characterization of Bioactive Compounds in Extracts and Fractions from Litopenaeus vannamei Muscle
by Sandra Carolina De La Reé-Rodríguez, María Jesús González, Ingrid Fernández, José Luis Garrido, Erika Silva-Campa, Norma Violeta Parra-Vergara, Carmen María López-Saiz and Isabel Medina
Mar. Drugs 2025, 23(2), 59; https://doi.org/10.3390/md23020059 - 27 Jan 2025
Viewed by 496
Abstract
Marine organisms are a vital source of biologically active compounds. Organic extracts from the muscle of the Pacific white shrimp (L. vannamei) have shown antiproliferative effects on tumor cells, including breast adenocarcinoma. This study aimed to analyze these extracts’ composition and [...] Read more.
Marine organisms are a vital source of biologically active compounds. Organic extracts from the muscle of the Pacific white shrimp (L. vannamei) have shown antiproliferative effects on tumor cells, including breast adenocarcinoma. This study aimed to analyze these extracts’ composition and confirm their specificity for breast adenocarcinoma cells without harming normal cells. An organic chloroform extract from L. vannamei muscle was divided using a solvent partition procedure with methanol and hexane. The methanolic partition was fractionated through an open preparative liquid chromatography column to isolate compounds with biological activity, that were later tested on MDA-MB-231 (breast adenocarcinoma), and recently tested on MCF10-A (non-cancerous breast epithelial cells). Cells incubated with these fractions were assessed for viability and morphological changes using fluorescence confocal microscopy. Fractions F#13 and F#14 reduced MDA-MB-231 cancer cell viability at 100 µg/mL without affecting non-cancerous MCF-10A cells, inducing apoptosis-related changes in cancer cells. These fractions contained EPA and DHA free fatty acids, specifically F#13 contained free and esterified astaxanthin as well. The high levels of free linoleic acid 18:2 ω-6, EPA, and DHA (in a 2:1 ratio, EPA:DHA), along with free and esterified astaxanthin in F#13, significantly reduced breast adenocarcinoma cell viability, nearly to that achieved by cisplatin, a chemotherapy drug. Full article
Show Figures

Figure 1

Figure 1
<p>Neutral lipids TLC; R1–R4 standard calibration graph as described in the experimental section; Triacyl glycerides (TG), free fatty acids (FFA), cholesterol (CHL), diacylglycerides (DG), monoacylglycerides (MG), and polar lipids (PL); Chloroform extract (CHCl<sub>3</sub>), hexane partition (Hex), methanol partition (MeOH).</p>
Full article ">Figure 2
<p>Polar lipids TLC; R1–R4 standard calibration graph as described in the experimental section; Phosphatidylethanolamine (PE) and phosphatidylcholine (PC). Chloroform extract (CHCl<sub>3</sub>), hexane partition (Hex), methanol partition (MeOH).</p>
Full article ">Figure 3
<p>TLC of the 22 eluates obtained from the chromatographic column. 1: The eluates within the circle are the fractions of interest, 2: Polar lipids present in the eluate 21.</p>
Full article ">Figure 4
<p>Viability percentage of MDA-MB-231 cells treated with F13 and F14 at different concentrations, as controls eicosapentaenoic acid (EPA), doxorubicin (DX) and, cisplatin (Cisp). Values are mean ± standard error from three independent experiments. Values with different number of asterisks are significantly different (<span class="html-italic">p</span> ≤ 0.05). *** significantly lower viability, ** significantly low viability, * significantly different viability.</p>
Full article ">Figure 5
<p>Viability percentage of MCF-10A cells treated with F13 and F14 at different concentrations, as controls eicosapentaenoic acid (EPA), doxorubicin (DX), and cisplatin (Cisp). Values are mean ± standard error from three independent experiments. Values with different number of asterisks are significantly different (<span class="html-italic">p</span> ≤ 0.05). ** significantly lower viability and * significantly different viability.</p>
Full article ">Figure 6
<p>MCF-10A cells observed at 20×. Morphology after 48h treatment at 100 μg/mL.</p>
Full article ">Figure 7
<p>MDA-MB-231 cells observed at 60× after 24 h of treatment with (1) control, (2) F13 (IC<sub>50</sub>: 102 μg/mL), (3) F14 (IC<sub>50</sub> 139 μg/mL), (4) EPA (IC<sub>50</sub>: 27 μg/mL), and (5) Cisplatin (IC<sub>50</sub>: 8 μg/mL). The cytoskeleton and DNA were dyed with phalloidin and DAPI, respectively.</p>
Full article ">
Back to TopTop