[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
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

Search Results (899)

Search Parameters:
Keywords = montmorillonite

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 3195 KiB  
Article
Sodium Alginate–Montmorillonite Composite Film Coatings for Strawberry Preservation
by Xiaoping Yan, Zuolong Yu, Yao Chen, Chao Han, Yunxiao Wei, Fan Yang, Yan Qian and Yong Wang
Coatings 2024, 14(10), 1331; https://doi.org/10.3390/coatings14101331 - 17 Oct 2024
Abstract
In this study, we prepared sodium alginate (SA) and montmorillonite (MMT) composite films for application in coatings for strawberry preservation. SA and MMT were used as the matrix and glycerol was used as a plasticizer. Six types of composite films with different MMT [...] Read more.
In this study, we prepared sodium alginate (SA) and montmorillonite (MMT) composite films for application in coatings for strawberry preservation. SA and MMT were used as the matrix and glycerol was used as a plasticizer. Six types of composite films with different MMT contents were compared by analyzing their mechanical properties, permeability, and preservation effects. The results show that the mechanical properties of the 10 and 20% MMT composite films were superior, with tensile strength and fracture elongation values reaching 63.09 and 48.06 MPa and 5.75 and 6.47%, respectively. Increased MMT content caused the water vapor permeability to decrease, while the effect on oil permeability was the opposite. A comparison of the preservation effect provided by the coatings showed that, on day 12, the weight loss, malondialdehyde content, and respiratory intensity of strawberries treated with the 20% MMT coating liquid decreased by 43.3, 25.8, and 57.1%, respectively, compared with the control. The contents of titratable acid, soluble sugar, total phenols, and soluble solids decreased by 25.8, 37.7, 25.9, and 14.5%, respectively. The results provide data support for the application of these new composite films as edible coatings for fruit preservation. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of montmorillonite (MMT) content on the mechanical properties of the composite films.</p>
Full article ">Figure 2
<p>Effect of MMT content on the permeability of composite films.</p>
Full article ">Figure 3
<p>Infrared spectra of composite films with different MMT contents.</p>
Full article ">Figure 4
<p>Effect of different composite film liquid coatings on strawberry weight loss.</p>
Full article ">Figure 5
<p>Effect of different composite film liquid coatings on the malondialdehyde content of strawberries.</p>
Full article ">Figure 6
<p>Effect of different composite film liquid coatings on the respiratory intensity of strawberries.</p>
Full article ">Figure 7
<p>Effect of different composite film liquid coatings on titratable acid content of strawberries.</p>
Full article ">Figure 8
<p>Effect of different composite film liquid coatings on the soluble sugar content of strawberries.</p>
Full article ">Figure 9
<p>Effect of different composite film liquid coatings on the total phenol content of strawberries.</p>
Full article ">Figure 10
<p>Effect of different composite film liquid coatings on the soluble solid content of strawberries.</p>
Full article ">
12 pages, 5565 KiB  
Article
The Effects of Montmorillonite–Humic Acid Composite Particles on the Photolysis of Tetracycline in Water
by Wenfang Zhou, Zirui Wang, Qingfeng Wu, Qinping Nie and Yi Wang
Crystals 2024, 14(10), 896; https://doi.org/10.3390/cryst14100896 - 16 Oct 2024
Viewed by 343
Abstract
Suspended particulate matter (SPM) is an important component of natural water bodies and can significantly influence the photolytic behavior of water pollutants. A comprehensive understanding of the photochemical behavior of water pollutants in natural waters requires consideration of the presence of SPM. In [...] Read more.
Suspended particulate matter (SPM) is an important component of natural water bodies and can significantly influence the photolytic behavior of water pollutants. A comprehensive understanding of the photochemical behavior of water pollutants in natural waters requires consideration of the presence of SPM. In this study, montmorillonite–humic acid (MMT-HA) composite particles were synthesized to simulate SPM in natural waters and their effects on the photolysis of tetracycline (TC) were investigated. The results demonstrated that the presence of MMT-HA composite particles in water significantly enhanced the photolysis of TC, with the photolytic kinetics following a pseudo-first-order model. Electron spin resonance spectra and free radical quenching experiments indicated that the photoactive components (MMT and humic acids) in the composite particles induced the generation of reactive oxygen species under light exposure, further contributing to the enhanced photolysis of TC. Comparative analysis of the free radical signals and adsorption experiments revealed that the accelerated photolysis of TC was also related to the interfacial interaction between the MMT in the composite particles and the TC molecules. The formation of surface complexes between TC molecules and the negatively charged sites on the MMT surface facilitated light absorption and electron transfer, thereby accelerating the photolysis of TC. Photoproduct analysis indicated that the primary degradation pathways of TC in the composite particle systems included the addition of hydroxyl radicals to the aromatic ring, as well as demethylation, deamination and dehydration in the side chains. This study shows that SPM in water bodies can affect the photochemical behavior of pollutants and should be taken into account when assessing the phototransformation of pollutants in natural waters. Full article
(This article belongs to the Special Issue Advanced Surface Modifications on Materials)
Show Figures

Figure 1

Figure 1
<p>The XRD Patterns of MMT, MMT-HA, and MMT-FA particles.</p>
Full article ">Figure 2
<p>The SEM images of MMT (<b>a</b>,<b>b</b>), MMT-FA (<b>c</b>,<b>d</b>), and MMT-HA (<b>e</b>,<b>f</b>) at different magnifications. The adsorption of FA and HA on MMT surface is marked by red circles.</p>
Full article ">Figure 3
<p>The FTIR spectra of MMT before and after the modification with HA (<b>a</b>) and FA (<b>b</b>).</p>
Full article ">Figure 4
<p>The photolysis of TC in both pure water and MMT-HA composite particle systems under simulated solar light. <span class="html-italic">C<sub>t</sub></span>/<span class="html-italic">C</span><sub>0</sub> is the ratio of the concentration of TC at <b><span class="html-italic">t</span></b> time relative to the initial concentration.</p>
Full article ">Figure 5
<p>The ESR signals of <sup>1</sup>O<sub>2</sub> (<b>a</b>), ·OH (<b>b</b>), and O<sub>2</sub><sup>−</sup> (<b>c</b>) generated in the MMT-FA composite particle system under different irradiation times.</p>
Full article ">Figure 6
<p>The ESR signals of <sup>1</sup>O<sub>2</sub> (<b>a</b>), ·OH (<b>b</b>) and O<sub>2</sub><sup>−</sup> (<b>c</b>) generated in the MMT-HA composite particle system under different irradiation times.</p>
Full article ">Figure 7
<p>The effects of radical quenchers on the photolysis of TC in MMT-FA (<b>a</b>) and MMT-HA (<b>b</b>) composite particle systems.</p>
Full article ">Figure 8
<p>Comparison of the ESR signals of <sup>1</sup>O<sub>2</sub> (<b>a</b>), OH (<b>b</b>) and O<sub>2</sub><sup>−</sup> (<b>c</b>) generated in the MMT, MMT-HA, and MMT-FA systems after 10 min irradiation.</p>
Full article ">Figure 9
<p>The photolysis of TC in the MMT, MMT-FA, and MMT-HA particle systems.</p>
Full article ">Figure 10
<p>The possible photolytic pathways of TC in the MMT-FA and MMT-HA composite particle systems.</p>
Full article ">
20 pages, 13856 KiB  
Article
Clay Minerals/TiO2 Composites—Characterization and Application in Photocatalytic Degradation of Water Pollutants
by Bogna D. Napruszewska, Dorota Duraczyńska, Joanna Kryściak-Czerwenka, Paweł Nowak and Ewa M. Serwicka
Molecules 2024, 29(20), 4852; https://doi.org/10.3390/molecules29204852 - 13 Oct 2024
Viewed by 361
Abstract
TiO2 used for photocatalytic water purification is most active in the form of nanoparticles (NP), but their use is fraught with difficulties in separation from solution or/and a tendency to agglomerate. The novel materials designed in this work circumvent these problems by [...] Read more.
TiO2 used for photocatalytic water purification is most active in the form of nanoparticles (NP), but their use is fraught with difficulties in separation from solution or/and a tendency to agglomerate. The novel materials designed in this work circumvent these problems by immobilizing TiO2 NPs on the surface of exfoliated clay minerals. A series of TiO2/clay mineral composites were obtained using five different clay components: the Na-, CTA-, or H-form of montmorillonite (Mt) and Na- or CTA-form of laponite (Lap). The TiO2 component was prepared using the inverse microemulsion method. The composites were characterized with X-ray diffraction, scanning/transmission electron microscopy/energy dispersive X-ray spectroscopy, FTIR spectroscopy, thermal analysis, and N2 adsorption–desorption isotherms. It was shown that upon composite synthesis, the Mt interlayer became filled by a mixture of CTA+ and hydronium ions, regardless of the nature of the parent clay, while the structure of Lap underwent partial destruction. The composites displayed high specific surface area and uniform mesoporosity determined by the size of the TiO2 nanoparticles. The best textural parameters were shown by composites containing clay components whose structure was partially destroyed; for instance, Ti/CTA-Lap had a specific surface area of 420 m2g−1 and a pore volume of 0.653 cm3g−1. The materials were tested in the photodegradation of methyl orange and humic acid upon UV irradiation. The photocatalytic activity could be correlated with the development of textural properties. In both reactions, the performance of the most photoactive composites surpassed that of the reference commercial P25 titania. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of (<b>a</b>) clay minerals used as supports for TiO<sub>2</sub> nanoparticles; (<b>b</b>) composites of TiO<sub>2</sub> and clay minerals.</p>
Full article ">Figure 2
<p>FTIR spectra of laponite-based composites, clay supports, reference nanocrystalline TiO<sub>2,</sub> and sum spectra of individual components.</p>
Full article ">Figure 3
<p>TEM images of (<b>a</b>) nanoparticles present in hydrolyzed Ti-containing inverse microemulsion; (<b>b</b>) Ti/CTA-Mt composite; (<b>c</b>) Ti//H-Mt composite. Sample suspensions deposited on 200 mesh copper grids covered with carbon film. Magnification ×200,000.</p>
Full article ">Figure 4
<p>SEM/EDX compositional analysis of selected areas of synthesized composites: (<b>a</b>) SEM image of Ti/CTA-Mt; (<b>b</b>) EDX mapping of Ti in Ti/CTA-Mt; (<b>c</b>) EDX mapping of Si in Ti/CTA-Mt; (<b>d</b>) SEM image of Ti/Na-Mt; (<b>e</b>) EDX mapping of Ti in Ti/Na-Mt; (<b>f</b>) EDX mapping of Si in Ti/Na-Mt; (<b>g</b>) SEM image of Ti/H-Mt; (<b>h</b>) EDX mapping of Ti in Ti/H-Mt; (<b>i</b>) EDX mapping of Si in Ti/H-Mt; (<b>j</b>) SEM image of Ti/CTA-Lap; (<b>k</b>) EDX mapping of Ti in Ti/CTA-Lap; (<b>l</b>) EDX mapping of Si in Ti/CTA-Lap; (<b>m</b>) SEM image of Ti/Na-Lap; (<b>n</b>) EDX mapping of Ti in Ti/Na-Lap; (<b>o</b>) EDX mapping of Si in Ti/Na-Lap.</p>
Full article ">Figure 4 Cont.
<p>SEM/EDX compositional analysis of selected areas of synthesized composites: (<b>a</b>) SEM image of Ti/CTA-Mt; (<b>b</b>) EDX mapping of Ti in Ti/CTA-Mt; (<b>c</b>) EDX mapping of Si in Ti/CTA-Mt; (<b>d</b>) SEM image of Ti/Na-Mt; (<b>e</b>) EDX mapping of Ti in Ti/Na-Mt; (<b>f</b>) EDX mapping of Si in Ti/Na-Mt; (<b>g</b>) SEM image of Ti/H-Mt; (<b>h</b>) EDX mapping of Ti in Ti/H-Mt; (<b>i</b>) EDX mapping of Si in Ti/H-Mt; (<b>j</b>) SEM image of Ti/CTA-Lap; (<b>k</b>) EDX mapping of Ti in Ti/CTA-Lap; (<b>l</b>) EDX mapping of Si in Ti/CTA-Lap; (<b>m</b>) SEM image of Ti/Na-Lap; (<b>n</b>) EDX mapping of Ti in Ti/Na-Lap; (<b>o</b>) EDX mapping of Si in Ti/Na-Lap.</p>
Full article ">Figure 5
<p>(<b>a</b>) TG traces of clay components; (<b>b</b>) TG traces of Ti/clay composites and of reference TiO<sub>2</sub>; (<b>c</b>) DSC traces of clay components; (<b>d</b>) DSC traces of Ti/clay composites and of reference TiO<sub>2</sub>.</p>
Full article ">Figure 6
<p>N<sub>2</sub> adsorption/desorption isotherms at −196 °C of (<b>a</b>) CTA-Mt and Ti/CTA-Mt; (<b>c</b>) Na-Mt and Ti/Na-Mt; (<b>e</b>) H-Mt and Ti/H-Mt; (<b>g</b>) CTA-Lap and Ti/CTA-Lap; (<b>i</b>) Na-Lap and Ti/Na-Lap. Differential pore size distribution of (<b>b</b>) CTA-Mt and Ti/CTA-Mt; (<b>d</b>) Na-Mt and Ti/Na-Mt; (<b>f</b>) H-Mt and Ti/H-Mt; (<b>h</b>) CTA-Lap and Ti/CTA-Lap; (<b>j</b>) Na-Lap and Ti/Na-Lap.</p>
Full article ">Figure 6 Cont.
<p>N<sub>2</sub> adsorption/desorption isotherms at −196 °C of (<b>a</b>) CTA-Mt and Ti/CTA-Mt; (<b>c</b>) Na-Mt and Ti/Na-Mt; (<b>e</b>) H-Mt and Ti/H-Mt; (<b>g</b>) CTA-Lap and Ti/CTA-Lap; (<b>i</b>) Na-Lap and Ti/Na-Lap. Differential pore size distribution of (<b>b</b>) CTA-Mt and Ti/CTA-Mt; (<b>d</b>) Na-Mt and Ti/Na-Mt; (<b>f</b>) H-Mt and Ti/H-Mt; (<b>h</b>) CTA-Lap and Ti/CTA-Lap; (<b>j</b>) Na-Lap and Ti/Na-Lap.</p>
Full article ">Figure 7
<p>Results of photocatalytic experiments after 5 h irradiation: (<b>a</b>) composites ordered according to the increasing crystallinity of the titania component; (<b>b</b>) composites ordered according to the increasing pore volume.</p>
Full article ">Figure 8
<p>Scheme of setup for photocatalytic experiments.</p>
Full article ">
23 pages, 3251 KiB  
Article
Regeneration and Single Stage Batch Adsorber Design for Efficient Basic Blue-41 Dye Removal by Porous Clay Heterostructures Prepared from Al13 Montmorillonite and Pillared Derivatives
by Saheed A. Popoola, Hmoud Al Dmour, Rawan Al-Faze, Mohd Gulfam Alam, Souad Rakass, Hicham Oudghiri Hassani and Fethi Kooli
Materials 2024, 17(20), 4948; https://doi.org/10.3390/ma17204948 - 10 Oct 2024
Viewed by 566
Abstract
Porous clay heterostructures are a hybrid precursor between the pillaring process and organoclays. In this study, the organoclay was substituted by an aluminium intercalated species clay or pillared alumina clays. A porous clay heterostructure was successfully achieved from an aluminium intercalated species clay, [...] Read more.
Porous clay heterostructures are a hybrid precursor between the pillaring process and organoclays. In this study, the organoclay was substituted by an aluminium intercalated species clay or pillared alumina clays. A porous clay heterostructure was successfully achieved from an aluminium intercalated species clay, due to the easy exchange of the aluminium species by the cosurfactant and silica species. However, using alumina pillared clays, the porous clay heterostructures were not formed; the alumina species were strongly attached to clay sheets which made difficult their exchange with cosurfactant molecules. In this case, the silica species were polymerized and decorated the surface of the used materials as indicated by different characterization techniques. The specific surface area of the porous clay heterostructure material reached 880 m2/g, and total pore volume of 0.258 cc/g, while the decorated silica alumina pillared clays exhibited lower specific surface area values of 244–440 m2/g and total pore volume of 0.315 to 0.157 cc/g. The potential of the synthesized materials was evaluated as a basic blue-41 dye removal agent. Porous clay heterostructure material has a removal capacity of 279 mg/g; while the other materials exhibited lower removal capacities between 75 mg/g and 165 mg/g. The used regeneration method was related to the acidity of the studied materials. The acidity of the materials possessed an impact on the adopted regeneration procedure in this study, the removal efficiency was maintained at 80% of the original performance after three successive regeneration cycles for the porous clay heterostructure. The Langmuir isotherm characteristics were used to propose a single-stage batch design. Porous clay heterostructures with a higher removal capacity resulted in a decrease in the quantities needed to achieve the target removal percentage of the BB-41 dye from an aqueous solution. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>left</b>) PXRD patterns of raw clay, intercalated with the Al<sub>13</sub> species and calcined at different temperatures; (<b>right</b>) after reaction with C<sub>12</sub>amine and TEOS, then calcined at 550 °C.</p>
Full article ">Figure 2
<p>TEM micrographs of (<b>a</b>) raw Mt, (<b>b</b>) intercalated with the Al13 species (Al-IMt), (<b>c</b>) after calcination at 500 °C, (<b>d</b>) PAl-MtCH, and (<b>e</b>) PAl-Mt500CH.</p>
Full article ">Figure 3
<p>(<b>left</b>) <sup>29</sup>Si MAS NMR and (<b>right</b>) <sup>27</sup>Al MAS NMR of the different materials.</p>
Full article ">Figure 4
<p>TGA (black) and DTG (red) features of the different materials: (<b>a</b>) Mt, (<b>b</b>) Al-IMt, (<b>c</b>) PAl-IMtCH, (<b>d</b>) PAl-Mt(500), and (<b>e</b>) derived PAl-Mt(500)CH.</p>
Full article ">Figure 5
<p>N<sub>2</sub> adsorption-desorption isotherms of different materials: (<b>a</b>) Mt, (<b>b</b>) Al-IMt, (<b>c</b>) PAlMtCH, (<b>d</b>) Al-PMt(500), and (<b>e</b>) PMt(500)CH.</p>
Full article ">Figure 6
<p>Effect of on the removal properties of BB-41 dye, (<b>left</b>) PAl-IMtCH used mass and (<b>right</b>) initial BB-41 pH solution.</p>
Full article ">Figure 7
<p>Effect of the BB-41 initial concentration of the removal properties of the PAl-IMtCH (filled triangles) and PAl-PMt(500)CH (non-filled triangles).</p>
Full article ">Figure 8
<p>Variation of removal percentage (%) after different regeneration cycles.</p>
Full article ">Figure 9
<p>Required masses of PAl-IMtCH (<b>left</b>) and PAl-PMt(500)CH (<b>right</b>) to reduce different volumes (L) of BB-41 solutions (C<sub>i</sub> = 200 mg/L) to different removal percentages.</p>
Full article ">
15 pages, 2463 KiB  
Article
Physical–Chemical and Thermal Properties of Clays from Porto Santo Island, Portugal
by André Valente, Paula C. S. Carvalho and Fernando Rocha
Appl. Sci. 2024, 14(19), 8962; https://doi.org/10.3390/app14198962 - 5 Oct 2024
Viewed by 470
Abstract
The use of clays for thermal treatments and cosmetic purposes continues to be a worldwide practice, whether through the preservation of native cultural traditions, pharmaceutical formulations or integrative health and well-being practices. Special clays, such as bentonites, are very common for healing applications [...] Read more.
The use of clays for thermal treatments and cosmetic purposes continues to be a worldwide practice, whether through the preservation of native cultural traditions, pharmaceutical formulations or integrative health and well-being practices. Special clays, such as bentonites, are very common for healing applications due to their high cation exchange capacity (CEC), high specific surface area (SSA) and alkaline pH values and, therefore, are used in multiple therapeutic and dermocosmetic treatments. Numerous bentonitic deposits occur on Porto Santo Island with different chemical weathering degrees. This research evaluates which residual soils have the most suitable characteristics for pelotherapy. The texture of residual soils varies from silt loam to loamy sand and SSA between 39 and 90 m2/g. The pH is alkaline (8.7 to 9.6), electrical conductivity ranges from 242 to 972 µS/cm, and CEC from 50.4 to 86.8 µS/cm. The residual soils have a siliciclastic composition (41.36 to 54.02% SiO2), between 12.52 and 17.65% Al2O3 and between 52 and 82% smectite content, which are montmorillonite and nontronite. Specific heat capacity (0.5–0.9 J/g°C) and cooling kinetics (14.5–19 min) show that one residual soil has the potential to be suitable for pelotherapy according to the literature. Moreover, the residual soils have As, Cd, Co, Cr, Hg, Mn, Ni, Pb, Sb and V concentrations higher than the limits of guidelines for cosmetics and pharmaceutical products. Full article
(This article belongs to the Section Earth Sciences)
Show Figures

Figure 1

Figure 1
<p>The geographic settings of Porto Santo Island and the Madeira archipelago (<b>a</b>,<b>b</b>) and the geological setting of Porto Santo Island with the sampling distribution (<b>c</b>).</p>
Full article ">Figure 2
<p>SEM images (ampliation 150× and resolution 15 keV) of Sample 14 (<b>a</b>), with different particles sizes and shapes, and Sample 36A (<b>b</b>), with a homogeneous morphology.</p>
Full article ">Figure 3
<p>XRD graphs of samples: (<b>a</b>) 14A and (<b>b</b>) 36A.</p>
Full article ">Figure 4
<p>Photography of a fine particle (&lt;2 μm) of montmorillonite (ampliation 150× and resolution 15 keV).</p>
Full article ">
17 pages, 5080 KiB  
Article
Study on Rheological Properties of Waste Cooking Oil and Organic Montmorillonite Composite Recycled Asphalt
by Cheng Xie, Qunshan Ye, Lingyi Fan, Anqi Weng and Haobin Liu
Buildings 2024, 14(10), 3149; https://doi.org/10.3390/buildings14103149 - 2 Oct 2024
Viewed by 504
Abstract
Pre-treated waste cooking oil (WCO) and organic montmorillonite (OMMT) were employed for the recycling of aged asphalt, which resulted in the improvement of the design of WCO asphalt rejuvenators and the enhancement of high-temperature performance of WCO-recycled asphalt. The effect of the rejuvenator [...] Read more.
Pre-treated waste cooking oil (WCO) and organic montmorillonite (OMMT) were employed for the recycling of aged asphalt, which resulted in the improvement of the design of WCO asphalt rejuvenators and the enhancement of high-temperature performance of WCO-recycled asphalt. The effect of the rejuvenator and the properties of recycled asphalt were evaluated by viscosity, dynamic shear rheometer (DSR), bending beam rheometer (BBR) and scanning electron microscopy (SEM), Fourier-transform infrared spectroscopy (FTIR), and gel permeation chromatography (GPC) tests. The results indicated that aged asphalt could be obviously softened and restored to the level of original asphalt by adding 6% WCO. However, the high-temperature properties of recycled asphalt would be declined by adding too large a dose of WCO rejuvenator. The high-temperature performance of recycled asphalt was significantly improved by the WCO-OMMT complex rejuvenator, and the viscosity and rutting factor of recycled asphalt were increased. Light components of aged asphalt could be supplemented by WCO of the complex rejuvenator. The volatilization of small molecules could be slowed down by the peel structure formed by OMMT and small molecules of the asphalt, which resulted in the proportion of small molecular substances (SMS) being increased by 4% and improvement of the colloidal structure of aged asphalt. The high-temperature and low-temperature performance of recycled asphalt can be improved concurrently by the combination of 6% WCO and 1% OMMT, and this was evidenced by the fact that the high-temperature and low-temperature PG were all upgraded by one level. Full article
(This article belongs to the Special Issue Research on Advanced Materials in Road Engineering)
Show Figures

Figure 1

Figure 1
<p>WCO.</p>
Full article ">Figure 2
<p>OMMT.</p>
Full article ">Figure 3
<p>Different Asphalts’ Brookfield Viscosities at 135 °C.</p>
Full article ">Figure 4
<p>Different Asphalts’ viscous activation energy.</p>
Full article ">Figure 5
<p>(<b>a</b>) Rutting factor of WCO-recycled asphalt; (<b>b</b>) rutting factor of composite recycled asphalt.</p>
Full article ">Figure 6
<p>(<b>a</b>) The R-value of WCO-recycled asphalt; (<b>b</b>) the R-value of composite recycled asphalt.</p>
Full article ">Figure 7
<p>(<b>a</b>) The S-value of WCO-recycled asphalt; (<b>b</b>) the m-value of WCO-recycled asphalt; (<b>c</b>) the S-value of composite recycled asphalt; (<b>d</b>) the m-value of composite recycled asphalt.</p>
Full article ">Figure 8
<p>(<b>a</b>) FTIR spectra of different functional groups; (<b>b</b>) FTIR spectra of WOA with varying degrees of aging.</p>
Full article ">Figure 9
<p>(<b>a</b>) The SEM image of OA; (<b>b</b>) The SEM image of RA; (<b>c</b>) The SEM image of WOA.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>) The SEM image of OA; (<b>b</b>) The SEM image of RA; (<b>c</b>) The SEM image of WOA.</p>
Full article ">Figure 10
<p>The molecular weight distribution of different asphalt.</p>
Full article ">Figure 11
<p>(<b>a</b>) Average molecular weight of different asphalts with diverse degrees of aging; (<b>b</b>) PDIs of different asphalt with diverse degrees of aging.</p>
Full article ">
19 pages, 4920 KiB  
Article
Novel, Fluorine-Free Membranes Based on Sulfonated Polyvinyl Alcohol and Poly(ether-block-amide) with Sulfonated Montmorillonite Nanofiller for PEMFC Applications
by Manhal H. Ibrahim Al-Mashhadani, Gábor Pál Szijjártó, Zoltán Sebestyén, Zoltán Károly, Judith Mihály and András Tompos
Membranes 2024, 14(10), 211; https://doi.org/10.3390/membranes14100211 - 1 Oct 2024
Viewed by 490
Abstract
Novel blend membranes containing S-PVA and PEBAX 1657 with a blend ratio of 8:2 (referred to as SPP) were prepared using a solution-casting technique. In the manufacturing process, sulfonated montmorillonite (S-MMT) in ratios of 0%, 3%, 5%, and 7% was used as a [...] Read more.
Novel blend membranes containing S-PVA and PEBAX 1657 with a blend ratio of 8:2 (referred to as SPP) were prepared using a solution-casting technique. In the manufacturing process, sulfonated montmorillonite (S-MMT) in ratios of 0%, 3%, 5%, and 7% was used as a filler. The crystallinity of composite membranes has been investigated by X-ray diffraction (XRD), while the interaction between the components was evaluated using Fourier-transform infrared spectroscopy (FT-IR). With increasing filler content, good compatibility between the components due to hydrogen bonds was established, which ultimately resulted in improved tensile strength and chemical stability. In addition, due to the sulfonated moieties of S-MMT, the highest ion exchange capacity (0.46 meq/g) and water uptake (51.61%) can be achieved at the highest filler content with an acceptable swelling degree of 22.65%. The composite membrane with 7% S-MMT appears to be suitable for application in proton exchange membrane fuel cells (PEMFCs). Amongst the membranes studied, this membrane achieved the highest current density and power density in fuel cell tests, which were 149.5 mA/cm2 and 49.51 mW/cm2. Our fluorine-free composite membranes can become a promising new membrane family in PEMFC applications, offering an alternative to Nafion membranes. Full article
(This article belongs to the Special Issue Recent Advances in Fluorine-Free Membranes)
Show Figures

Figure 1

Figure 1
<p>Manufacturing procedure of blend membranes with different S-MMT content.</p>
Full article ">Figure 2
<p>Water uptake and swelling ratio of SPP blend membranes and recast Nafion.</p>
Full article ">Figure 3
<p>Ion exchange capacity for SPP blend membranes and recast Nafion.</p>
Full article ">Figure 4
<p>(<b>A</b>): (<b>a</b>) Thermogravimetric (TG) and (<b>b</b>) derivate thermogravimetric (DTG) curves of S-PVA, PEBAX, and S-MMT; (<b>B</b>): (<b>a</b>) TG curves of the SPP blends membranes and (<b>b</b>) DTG curve and evolution profiles of some characteristic mass spectrometric ions of the SPP 7% S-MMT sample, <span class="html-italic">m</span>/<span class="html-italic">z</span>: 18 (water), <span class="html-italic">m</span>/<span class="html-italic">z</span>: 43 and 58 (acetone), <span class="html-italic">m</span>/<span class="html-italic">z</span>: 44 (carbon dioxide), <span class="html-italic">m</span>/<span class="html-italic">z</span>: 48 and 64 (sulfur dioxide).</p>
Full article ">Figure 5
<p>In-plane and cross-sectional SEM images (<b>I</b> and <b>II</b>) for blend membranes ((<b>A</b>–<b>D</b>) refer to SPP 0% S-MMT, SPP 3% S-MMT, SPP 5% S-MMT, and SPP 7% S-MMT, respectively).</p>
Full article ">Figure 6
<p>FTIR spectra of (<b>A</b>) SPP blend membranes and (<b>B</b>) the different constituents, such as parent and sulfonated montmorillonite (MMT and S-MMT), PVA, and S-PVA, as well as the pure PEBAX 1657.</p>
Full article ">Figure 7
<p>X-ray diffraction patterns of SPP blend membranes.</p>
Full article ">Figure 8
<p>Mechanical stability for SPP blend membranes.</p>
Full article ">Figure 9
<p>Chemical stability by Fenton’s test for SPP blend membranes and recast Nafion.</p>
Full article ">Figure 10
<p>Polarization curves and power density curves for SPP blend membranes. Solid and dashed lines correspond to voltage and power density.</p>
Full article ">
17 pages, 3762 KiB  
Article
Electrochemical Performance of Ti Gr. 2 as Electrodes in Contact with Saline Suspension of Clays during the Electroflotation Process
by Alvaro Soliz, Felipe M. Galleguillos-Madrid, José Ángel Cobos-Murcia, Sebastian Angulo, Sebastian Salazar-Avalos, Bernabé Alonso-Fariñas and Alexis Guzmán
Appl. Sci. 2024, 14(19), 8825; https://doi.org/10.3390/app14198825 - 1 Oct 2024
Viewed by 740
Abstract
The presence of clays in copper minerals has a significant negative impact during their processing, leading to low recoveries during the flotation process. In saline environments, the presence of these clays promotes operational problems associated with salinity, leading to decreases in the copper [...] Read more.
The presence of clays in copper minerals has a significant negative impact during their processing, leading to low recoveries during the flotation process. In saline environments, the presence of these clays promotes operational problems associated with salinity, leading to decreases in the copper concentrate grade, alterations in the rheology of the mineral pulp, reduction in the selectivity of copper during the flotation process, declines in the quality of clarified water, and excessive corrosion of metallic components. This study explores the electroflotation of kaolinite and montmorillonite clays in NaCl solutions using a modified Hallimond tube coupled with Ti Gr. 2 electrodes for bubble generation via water electrolysis and the corrosion analysis of these electrodes applying the superposition model. The electroflotation results show recovery of clays close to 72.68% for kaolinite, 88.44% for montmorillonite, and 67.36% for a mixture of both clays. The presence of clays helps reduce the corrosive effects of Ti Gr. 2 from 0.069 A/m2 in NaCl to 0.0073 A/m2 in NaCl with montmorillonite clay. Full article
Show Figures

Figure 1

Figure 1
<p>Diagram of the modified Hallimond tube used for the electroflotation tests.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>,<b>c</b>) montmorillonite and (<b>b</b>,<b>d</b>) kaolinite clays.</p>
Full article ">Figure 3
<p>Elemental mapping by EDS analysis of (<b>a</b>) kaolinite and (<b>b</b>) montmorillonite clays.</p>
Full article ">Figure 4
<p>XRD patterns of (<b>a</b>) montmorillonite and (<b>b</b>) kaolinite clays.</p>
Full article ">Figure 5
<p>Modified Hallimond tube, (<b>a</b>) before and (<b>b</b>) after the electroflotation process of clays.</p>
Full article ">Figure 6
<p>Polarization (<b>a</b>) and Tafel (<b>b</b>) curves for Ti Gr. 2 in contact with 0.5 M NaCl and slurry of clays. Experimental curve (dotted line) and fitted curve (line) obtained from the superposition model.</p>
Full article ">Figure 7
<p>First derivative curve of current density-potential data for Ti Gr. 2 in 0.5 M NaCl and solution kaolinite slurry.</p>
Full article ">Figure 8
<p>Pitting corrosion (<b>a</b>,<b>b</b>), and cracks on the oxide layer (<b>c</b>,<b>d</b>) detected over the Ti anode at 10 V operating for 1 h in 0.5 M NaCl.</p>
Full article ">
17 pages, 4439 KiB  
Article
The Use of Organoclays as Excipient for Metformin Delivery: Experimental and Computational Study
by Sondes Omrani, Safa Gamoudi, César Viseras, Younes Moussaoui and C. Ignacio Sainz-Díaz
Molecules 2024, 29(19), 4612; https://doi.org/10.3390/molecules29194612 - 28 Sep 2024
Viewed by 371
Abstract
This work combines experimental and computational modeling studies for the preparation of a composite of metformin and an organoclay, examining the advantages of a Tunisian clay used for drug delivery applications. The clay mineral studied is a montmorillonite-like smectite (Sm-Na), and the organoclay [...] Read more.
This work combines experimental and computational modeling studies for the preparation of a composite of metformin and an organoclay, examining the advantages of a Tunisian clay used for drug delivery applications. The clay mineral studied is a montmorillonite-like smectite (Sm-Na), and the organoclay derivative (HDTMA-Sm) was used as a drug carrier for metformin hydrochloride (MET). In order to assess the MET loading into the clays, these materials were characterized by means of cation exchange capacity assessment, specific surface area measurement, and with the techniques of X-ray diffraction (XRD), differential scanning calorimetry, X-ray fluorescence spectroscopy, and Fourier-transformed infrared spectroscopy. Computational molecular modeling studies showed the surface adsorption process, identifying the clay–drug interactions through hydrogen bonds, and assessing electrostatic interactions for the hybrid MET/Sm-Na and hydrophobic interactions and cation exchange for the hybrid MET/HDTMA-Sm. The results show that the clays (Sm-Na and HDTMA-Sm) are capable of adsorbing MET, reaching a maximum load of 12.42 and 21.97 %, respectively. The adsorption isotherms were fitted by the Freundlich model, indicating heterogeneous adsorption of the studied adsorbate–adsorbent system, and they followed pseudo-second-order kinetics. The calculations of ΔGº indicate the spontaneous and reversible nature of the adsorption. The calculation of ΔH° indicates physical adsorption for the purified clay (Sm-Na) and chemical adsorption for the modified clay (HDTMA-Sm). The release of intercalated MET was studied in media simulating gastric and intestinal fluids, revealing that the purified clay (Sm-Na) and the modified organoclay (HDTMA-Sm) can be used as carriers in controlled drug delivery in future clinical applications. The molecular modeling studies confirmed the experimental phenomena, showing that the main adsorption mechanism is the cation exchange process between proton and MET cations into the interlayer space. Full article
(This article belongs to the Special Issue Advanced Functional Nanomaterials in Medicine and Health Care)
Show Figures

Figure 1

Figure 1
<p>Structure of metformin hydrochloride.</p>
Full article ">Figure 2
<p>X-ray diffraction patterns of Sm-Na (<b>a</b>) and HDTMA-Sm (<b>b</b>).</p>
Full article ">Figure 3
<p>FTIR spectra of HDTMA-Sm (blue), HDTMA-Br (red), and Sm-Na (black).</p>
Full article ">Figure 4
<p>DSC profile of smectite before (up) and after (down) the formation of the HDTMA-Sm complex.</p>
Full article ">Figure 5
<p>FTIR spectra of MET (blue), MET/HDTMA-Sm (black), and HDTMA-Sm (red).</p>
Full article ">Figure 6
<p>Effect of pH on the adsorption of MET in the smectite forms.</p>
Full article ">Figure 7
<p>(<b>a</b>) MET adsorption isotherms on Sm-Na and HDTMA-Sm at 298 K. (<b>b</b>) Comparison of experimental data (Sm-Na in solid symbols, and HDTMA-Sm in hollow symbols) and the results of the Freundlich model (red line for Sm-Na, and black line for HDTMA-Sm).</p>
Full article ">Figure 8
<p>Kinetics of the MET absorption in Sm-Na and HDTMA-Sm at 298 K.</p>
Full article ">Figure 9
<p>Influence of temperature on the adsorption of MET by clays. Black line for Sm-Na and red line for HDTMA_Sm.</p>
Full article ">Figure 10
<p>Release profiles in SGF (pH = 1.2) and SIF (pH = 7.4) of MET/Sm-Na (<b>a</b>) and MET/HDTMA-Sm (<b>b</b>).</p>
Full article ">Figure 11
<p>Relationship between the number of water molecules per 3x2x1 supercell, and the increase in hydration energy (kcal/mol) with respect to the model with 12 water molecules per supercell.</p>
Full article ">Figure 12
<p>Optimized structures of the smectites intercalated with HDTMA in vertical (<b>a</b>) and horizontal (<b>b</b>) orientations with respect to the interlayer mineral surface. The H, O, N, C, Si, Al and Mg atoms are represented in white, red, blue, grey, yellow, pink, and green colors, respectively. This criterion is applied for the rest of the figures in this work.</p>
Full article ">Figure 13
<p>Optimized structures of MET intercalated in the HDTMA organoclay by cation exchange with a water proton (<b>a</b>) and by MET hydrochloride molecular adsorption (<b>b</b>). The atoms of the MET molecule are highlighted in balls.</p>
Full article ">
23 pages, 14535 KiB  
Article
The Synthesis of Green Palladium Catalysts Stabilized by Chitosan for Hydrogenation
by Farida Bukharbayeva, Alima Zharmagambetova, Eldar Talgatov, Assemgul Auyezkhanova, Sandugash Akhmetova, Aigul Jumekeyeva, Akzhol Naizabayev, Alima Kenzheyeva and Denis Danilov
Molecules 2024, 29(19), 4584; https://doi.org/10.3390/molecules29194584 - 26 Sep 2024
Viewed by 471
Abstract
The proposed paper describes a simple and environmentally friendly method for the synthesis of three-component polymer–inorganic composites, which includes the modification of zinc oxide or montmorillonite (MMT) with chitosan (CS), followed by the immobilization of palladium on the resulting two-component composites. The structures [...] Read more.
The proposed paper describes a simple and environmentally friendly method for the synthesis of three-component polymer–inorganic composites, which includes the modification of zinc oxide or montmorillonite (MMT) with chitosan (CS), followed by the immobilization of palladium on the resulting two-component composites. The structures and properties of the obtained composites were characterized by physicochemical methods (IRS, TEM, XPS, SEM, EDX, XRD, BET). Pd–CS species covered the surface of inorganic materials through two different mechanisms. The interaction of chitosan polyelectrolyte with zinc oxide led to the deprotonation of its amino groups and deposition on the surface of ZnO. The immobilization of Pd on CS/ZnO occurred by the hydrolysis of [PdCl4]2−, followed by forming PdO particles by interacting with amino groups of chitosan. In the case of CS/MMT, protonated amino groups of CS interacted with negative sites of MMT, forming a positively charged CS/MMT composite. Furthermore, [PdCl4]2− interacted with the –NH3+ sites of CS/MMT through electrostatic force. According to TEM studies of 1%Pd–CS/ZnO, the presence of Pd nanoclusters composed of smaller Pd nanoparticles of 3–4 nm in size were observed on different sites of CS/ZnO. For 1%Pd–CS/MMT, Pd nanoparticles with sizes of 2 nm were evenly distributed on the support surface. The prepared three-component CS–inorganic composites were tested through the hydrogenation of 2-propen-1-ol and acetylene compounds (phenylacetylene, 2-hexyn-1-ol) under mild conditions (T—40 °C, PH2—1 atm). It was shown that the efficiency of 1%Pd–CS/MMT is higher than that of 1%Pd–CS/ZnO, which can be explained by the formation of smaller Pd particles that are evenly distributed on the support surface. The mechanism of 2-hexyn-1-ol hydrogenation over an optimal 1%Pd–CS/MMT catalyst was proposed. Full article
Show Figures

Figure 1

Figure 1
<p>Diffractograms of zinc oxide (<b>a</b>), chitosan (<b>b</b>), and the 5% CS/ZnO (<b>c</b>) composite.</p>
Full article ">Figure 2
<p>Diffractograms of MMT (<b>a</b>), 2% CS/MMT (<b>b</b>), and 4.6% CS/MMT (<b>c</b>) composites.</p>
Full article ">Figure 3
<p>Pore size distributions of MMT (<b>a</b>), 2.0%CS/MMT (<b>b</b>), ZnO (<b>c</b>), and 2.0%CS/ZnO (<b>d</b>).</p>
Full article ">Figure 4
<p>SEM images of zinc oxide (<b>a</b>), CS/ZnO (<b>b</b>), MMT (<b>c</b>) and CS/MMT (<b>d</b>).</p>
Full article ">Figure 5
<p>SEM/EDX elemental mapping images of 1%Pd–CS/ZnO (<b>a</b>) and 1%Pd–CS/MMT (<b>b</b>) catalysts.</p>
Full article ">Figure 6
<p>Survey XPS spectrum of 1%Pd-CS/ZnO (<b>a</b>) and 1%Pd-CS/MMT (<b>b</b>) catalysts.</p>
Full article ">Figure 7
<p>The Pd3<span class="html-italic">d</span> (<b>a</b>–<b>d</b>) and N<span class="html-italic">1s</span> (<b>e</b>,<b>f</b>) XPS spectra of the 1%Pd–CS/ZnO (<b>a</b>,<b>c</b>,<b>e</b>) and 1%Pd–CS/MMT (<b>b</b>,<b>d</b>,<b>f</b>) catalysts before (<b>a</b>,<b>b</b>) and after (<b>c</b>–<b>f</b>) treatment with H<sub>2</sub>.</p>
Full article ">Figure 8
<p>TEM microphotographs of 1%Pd–CS/ZnO (<b>a</b>,<b>b</b>) and 1%Pd–CS/MMT (<b>c</b>,<b>d</b>) catalysts at different magnifications.</p>
Full article ">Figure 9
<p>Proposed scheme of the formation of the CS/MMT (<b>a</b>), 1%Pd–CS/MMT (<b>b</b>), and 1%Pd–CS/ZnO (<b>c</b>) composites.</p>
Full article ">Figure 10
<p>Kinetics of the hydrogen uptake (<b>a</b>) and the change in reaction rate (<b>b</b>) on 1%Pd−CS/ZnO and 1%Pd−CS/MMT catalysts at the hydrogenation of phenylacetylene. Reaction conditions: T—40 °C, P<sub>H2</sub>—1 atm, m<sub>cat</sub>—0.05 g, solvent C<sub>2</sub>H<sub>5</sub>OH—0.25 mL, and C<sub>sub</sub>—0.09 mol/L.</p>
Full article ">Figure 11
<p>Changes in the composition of the reaction mixture during the hydrogenation of phenylacetylene in the presence of 1%Pd−CS/MMT (<b>a</b>) and 1%Pd−CS/ZnO (<b>b</b>). Reaction conditions: T = 40 °C, P<sub>H2</sub>—1 atm, m<sub>cat</sub>—0.05 g, solvent C<sub>2</sub>H<sub>5</sub>OH—0.25 mL, and C<sub>sub</sub>—0.09 mol/L.</p>
Full article ">Figure 12
<p>Changes in the composition of the reaction mixture during the hydrogenation of 2-hexyne-1-ol on 1%Pd−CS/MMT (<b>a</b>) and 1%Pd−CS/ZnO (<b>b</b>). Reaction conditions: T—40 °C, P<sub>H2</sub>—1 atm, m<sub>cat</sub>—0.05 g, solvent C<sub>2</sub>H<sub>5</sub>OH—0.25 mL, and C<sub>sub</sub>—0.09 mol/L.</p>
Full article ">Figure 13
<p>Dependence of the conversion of 2-hexyn-1-ol on selectivity to cis-2-hexen-1-ol in the presence of 1%Pd−CS/MMT and 1%Pd−CS/ZnO composites.</p>
Full article ">Figure 14
<p>Plausible pathways of the hydrogenation of 2-propen-1-ol (<b>a</b>), phenylacetylene (<b>b</b>), and 2-hexyn-1-ol (<b>c</b>).</p>
Full article ">Figure 15
<p>Effect of the variation of reaction parameters on the activity of the 1%Pd-CS/MMT catalyst in the 2-hexyn-1ol hydrogenation: catalyst dosage (<b>a</b>); 2-hexyn-1-ol amount (<b>b</b>); concentration of hydrogen in hydrogen–helium mixture (<b>c</b>); temperature (<b>d</b>). Reaction conditions: 40 °C, 0.1 MPa, 100% H<sub>2</sub>, catalyst—25–100 mg, 2-hexyn-1-ol—0.25 mL, ethanol—25 mL (<b>a</b>); 40 °C, 0.1 MPa, 100% H<sub>2</sub>, catalyst—50 mg, 2-hexyn-1ol—0.10–1.00 mL, ethanol—25 mL (<b>b</b>); hydrogen percentage in H<sub>2</sub>:He gas mixture—30–100 vol%, catalyst—50 mg, 2-hexyn-1ol—0.25 mL, ethanol—25 mL (<b>c</b>); 25–40 °C, 0.1 MPa, 100% H<sub>2</sub>, catalyst—50 mg, 2-hexyn-1ol—0.25 mL, ethanol—25 mL (<b>d</b>).</p>
Full article ">
23 pages, 2408 KiB  
Review
Chitosan–Clay Mineral Nanocomposites with Antibacterial Activity for Biomedical Application: Advantages and Future Perspectives
by Danina Krajišnik, Snežana Uskoković-Marković and Aleksandra Daković
Int. J. Mol. Sci. 2024, 25(19), 10377; https://doi.org/10.3390/ijms251910377 - 26 Sep 2024
Viewed by 721
Abstract
Polymers of natural origin, such as representatives of various polysaccharides (e.g., cellulose, dextran, hyaluronic acid, gellan gum, etc.), and their derivatives, have a long tradition in biomedical applications. Among them, the use of chitosan as a safe, biocompatible, and environmentally friendly heteropolysaccharide has [...] Read more.
Polymers of natural origin, such as representatives of various polysaccharides (e.g., cellulose, dextran, hyaluronic acid, gellan gum, etc.), and their derivatives, have a long tradition in biomedical applications. Among them, the use of chitosan as a safe, biocompatible, and environmentally friendly heteropolysaccharide has been particularly intensively researched over the last two decades. The potential of using chitosan for medical purposes is reflected in its unique cationic nature, viscosity-increasing and gel-forming ability, non-toxicity in living cells, antimicrobial activity, mucoadhesiveness, biodegradability, as well as the possibility of chemical modification. The intuitive use of clay minerals in the treatment of superficial wounds has been known in traditional medicine for thousands of years. To improve efficacy and overcome the ubiquitous bacterial resistance, the beneficial properties of chitosan have been utilized for the preparation of chitosan–clay mineral bionanocomposites. The focus of this review is on composites containing chitosan with montmorillonite and halloysite as representatives of clay minerals. This review highlights the antibacterial efficacy of chitosan–clay mineral bionanocomposites in drug delivery and in the treatment of topical skin infections and wound healing. Finally, an overview of the preparation, characterization, and possible future perspectives related to the use of these advancing composites for biomedical applications is presented. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of chitosan.</p>
Full article ">Figure 2
<p>Schematic representation of the biomedical applications of chitosan-based bio-nanomaterials.</p>
Full article ">Figure 3
<p>Schematic representation of montmorillonite (<b>a</b>) and halloysite (<b>b</b>).</p>
Full article ">Figure 4
<p>Polymer–clay composite structures formed by the interaction between polymers and lamellar clays.</p>
Full article ">Figure 5
<p>Key features of chitosan–clay nanocomposites relevant to their biomedical applications.</p>
Full article ">Figure 6
<p>Release profiles of CLX formulations in different pH media (reprinted from Onnainty, R., Onida, B., Páez, P., Longhi, M., Barresi, A., &amp; Granero, G. (2016). Targeted chitosan-based bionanocomposites for controlled oral mucosal delivery of chlorhexidine. <span class="html-italic">International Journal of Pharmaceutics</span>, 509(1–2), 408–418 [<a href="#B90-ijms-25-10377" class="html-bibr">90</a>], with permission from Elsevier).</p>
Full article ">Figure 7
<p>In vivo lesion reduction vs. time profile evaluated for the following samples: NC—0.05 chitosan oligosaccharide/HTNs nanocomposite (HNT concentration of 300 μg/mL and chitosan oligosaccharide concentration of 4 μg/mL); HNTs (concentration of 300 μg/mL); chitosan oligosaccharide (concentration of 4 μg/mL); saline solution—negative control (mean values ± sd; <span class="html-italic">n</span> = 8) (reprinted from Sandri, G., Aguzzi, C., Rossi, S., Bonferoni, M. C., Bruni, G., Boselli, C., Cornaglia, A. I., Riva, F., Viseras, C., Caramella, C., &amp; Ferrari, F. (2017). Halloysite and chitosan oligosaccharide nanocomposite for wound healing. Acta Biomaterialia, 57, 216–224 [<a href="#B92-ijms-25-10377" class="html-bibr">92</a>], with permission from Elsevier).</p>
Full article ">
13 pages, 7982 KiB  
Article
Thermoplastic-Based Ballistic Helmets: Processing, Ballistic Resistance and Damage Characterization
by Rafael R. Dias, Natalin M. Meliande, Hector G. Kotik, César G. Camerini and Iaci M. Pereira
J. Compos. Sci. 2024, 8(10), 385; https://doi.org/10.3390/jcs8100385 - 24 Sep 2024
Viewed by 591
Abstract
Ballistic helmets are individual pieces of armor equipment designed to protect a soldier’s head from projectiles and fragments. Although very common, these helmets are responsible for several casualties due to their significant back face deformation and low ballistic resistance to projectiles. Therefore, to [...] Read more.
Ballistic helmets are individual pieces of armor equipment designed to protect a soldier’s head from projectiles and fragments. Although very common, these helmets are responsible for several casualties due to their significant back face deformation and low ballistic resistance to projectiles. Therefore, to enhance helmet performance, studies have focused on the development of new materials and new ballistic protection solutions. The purpose of this study was to develop and evaluate a new ballistic solution using thermoplastic-based matrices. The first matrix was based on high-density polyethylene (HDPE). The second matrix was based on HDPE modified with exfoliated montmorillonite (MMT). The main manufacturing processes of a thermoplastic-based ballistic helmet are presented, along with its ballistic performance, according to the National Institute of Justice (NIJ) standard 0106.01 and an investigation of its failure mechanisms via a non-destructive technique. All the helmets resulted in level III-A ballistic protection. The postimpact helmets were scanned to evaluate the back face deformation dimensions, which revealed that the global cone deformation was deeper in the HDPE than in the HDPE/MMT helmet. The failure analysis revealed an overall larger deformation area in the HDPE and HDPE/MMT helmet delamination zones in the regions with a large radius of curvature than in the zones with the lowest radius, which is in accordance with previous simulations reported in the literature. Full article
(This article belongs to the Section Composites Modelling and Characterization)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Ballistic test setup.</p>
Full article ">Figure 2
<p>Displacement measurement.</p>
Full article ">Figure 3
<p>(<b>a</b>) Seibt ES-35 single-screw extruder and (<b>b</b>) blown film.</p>
Full article ">Figure 4
<p>Ballistic helmet manufacturing process flowchart.</p>
Full article ">Figure 5
<p>Ballistic helmet pressing: (<b>a</b>) ACH mold; (<b>b</b>) layers of preform in place; and (<b>c</b>) pressed composite.</p>
Full article ">Figure 6
<p>Temperature–force vs. time applied in processing.</p>
Full article ">Figure 7
<p>Impact points: (<b>a</b>) front, (<b>b</b>) right, (<b>c</b>) back and (<b>d</b>) left.</p>
Full article ">Figure 8
<p>Ballistic tested helmets: (<b>a</b>) HDPE/MMT helmet and (<b>b</b>) 3D reconstructed images.</p>
Full article ">Figure 9
<p>Color maps of the (<b>a</b>) HDPE front impact outer and inner shell, (<b>b</b>) HDPE/MMT front impact outer and inner shell, (<b>c</b>) HDPE right impact outer and inner shell, and (<b>d</b>) HDPE/MMT right impact outer and inner shell.</p>
Full article ">Figure 10
<p>Results obtained with the 3D scanning metrology method: (<b>a</b>) displacement and (<b>b</b>) area.</p>
Full article ">Figure 11
<p>X-ray tomography images of postprocessed helmet shells: (<b>a</b>) HDPE and (<b>b</b>) HDPE/MMT (The yellow circles highlight areas where voids have been caused by manufacturing).</p>
Full article ">Figure 12
<p>X-ray CT images of the HDPE helmet after ballistic tests: (<b>a</b>) global visualization of the piece; slides showing the delaminations of (<b>b</b>) left-side impact, (<b>c</b>) right-side impact, (<b>d</b>) horizontal (xy) section plane—top view showing the impact position, (<b>e</b>) frontal impact, and (<b>f</b>) back-side impact.</p>
Full article ">Figure 12 Cont.
<p>X-ray CT images of the HDPE helmet after ballistic tests: (<b>a</b>) global visualization of the piece; slides showing the delaminations of (<b>b</b>) left-side impact, (<b>c</b>) right-side impact, (<b>d</b>) horizontal (xy) section plane—top view showing the impact position, (<b>e</b>) frontal impact, and (<b>f</b>) back-side impact.</p>
Full article ">Figure 13
<p>X-ray CT images of the HDPE/MMT helmet after ballistic tests: (<b>a</b>) global visualization of the piece; slides showing the delaminations of (<b>b</b>) left-side impact, (<b>c</b>) right-side impact, (<b>d</b>) horizontal (xy) section plane—top view showing the impact position, (<b>e</b>) frontal impact, and (<b>f</b>) back-side impact.</p>
Full article ">
19 pages, 13812 KiB  
Article
Structural and Thermal Characterization of Some Thermoplastic Starch Mixtures
by Maria Daniela Stelescu, Ovidiu-Cristian Oprea, Maria Sonmez, Anton Ficai, Ludmila Motelica, Denisa Ficai, Mihai Georgescu and Dana Florentina Gurau
Polysaccharides 2024, 5(4), 504-522; https://doi.org/10.3390/polysaccharides5040032 - 24 Sep 2024
Viewed by 481
Abstract
The paper presents the production of thermoplastic starch (TPS) mixtures using potato starch and two types of plasticizers: glycerol and sorbitol. The effects of plasticizers, citric acid, organically modified montmorillonite clay nanofiller (OMMT) and an additive based on ultrahigh molecular weight siloxane polymer [...] Read more.
The paper presents the production of thermoplastic starch (TPS) mixtures using potato starch and two types of plasticizers: glycerol and sorbitol. The effects of plasticizers, citric acid, organically modified montmorillonite clay nanofiller (OMMT) and an additive based on ultrahigh molecular weight siloxane polymer on the structure and physical–mechanical and thermal properties of TPS samples were analysed. Starch mixtures plasticized with glycerol were obtained, where the starch/glycerol mass ratio was 70:30, as well as starch mixtures plasticized with glycerol and sorbitol, with a starch/glycerol/sorbitol mass ratio of 60:20:20. The starch gelatinization process to obtain TPS was carried out in a Brabender Plasti-Corder internal mixer at 120 °C, with a mixing speed of 30–80 rpm, for 10 min. The obtained results indicate that by adding 2% (weight percentage) of citric acid to the TPS mixtures, there is an improvement in the physical–mechanical properties, as well as structural changes that can indicate both cross-linking reactions by esterification in stages and depolymerisation reactions. The sample of TPS plasticized with glycerol, which contains OMMT, shows an increase in tensile strength by 34.4%, compared to the control sample. Full article
Show Figures

Figure 1

Figure 1
<p>Diagrams showing the variations in the torque and the temperature versus time, when obtaining the mixtures using the Brabender Plasti-Corder.</p>
Full article ">Figure 2
<p>The proposed mechanism of cross-linking of citric acid with starch; green numbers represent the carbon numbering: citric anhydride formation (<b>1</b>); esterification reaction between citric anhydride and starch (<b>2</b>); formation of starch citrate by cross-linking (<b>3</b>).</p>
Full article ">Figure 2 Cont.
<p>The proposed mechanism of cross-linking of citric acid with starch; green numbers represent the carbon numbering: citric anhydride formation (<b>1</b>); esterification reaction between citric anhydride and starch (<b>2</b>); formation of starch citrate by cross-linking (<b>3</b>).</p>
Full article ">Figure 3
<p>Possible starch hydrolysis reaction with citric acid.</p>
Full article ">Figure 4
<p>FTIR spectra of samples S1–S4.</p>
Full article ">Figure 5
<p>FTIR spectra of samples S5–S8.</p>
Full article ">Figure 6
<p>FTIR maps of samples S1–S4 (red areas indicate the highest absorbance, while blue areas correspond to the lowest absorbance).</p>
Full article ">Figure 7
<p>FTIR maps of samples S5–S8 (red areas indicate the highest absorbance, while blue areas correspond to the lowest absorbance).</p>
Full article ">Figure 8
<p>TG-DSC curves of samples S1–S4 (<b>a</b>) and S5–S8 (<b>b</b>).</p>
Full article ">Figure 9
<p>Detail of TG-DSC curves of samples S1–S4 (<b>a</b>) and S5–S8 (<b>b</b>).</p>
Full article ">Figure 9 Cont.
<p>Detail of TG-DSC curves of samples S1–S4 (<b>a</b>) and S5–S8 (<b>b</b>).</p>
Full article ">
21 pages, 3611 KiB  
Article
Polymer Bionanocomposites Based on a P3BH/Polyurethane Matrix with Organomodified Montmorillonite—Mechanical and Thermal Properties, Biodegradability, and Cytotoxicity
by Beata Krzykowska, Łukasz Uram, Wiesław Frącz, Miroslava Kovářová, Vladimir Sedlařík, Dominika Hanusova, Maciej Kisiel, Joanna Paciorek-Sadowska, Marcin Borowicz and Iwona Zarzyka
Polymers 2024, 16(18), 2681; https://doi.org/10.3390/polym16182681 - 23 Sep 2024
Viewed by 580
Abstract
In the present work, hybrid nanobiocomposites based on poly(3-hydroxybutyrate), P3HB, with the use of aromatic linear polyurethane as modifier and organic nanoclay, Cloisite 30B, as a nanofiller were produced. The aromatic linear polyurethane (PU) was synthesized in a reaction of diphenylmethane 4,4′-diisocyanate and [...] Read more.
In the present work, hybrid nanobiocomposites based on poly(3-hydroxybutyrate), P3HB, with the use of aromatic linear polyurethane as modifier and organic nanoclay, Cloisite 30B, as a nanofiller were produced. The aromatic linear polyurethane (PU) was synthesized in a reaction of diphenylmethane 4,4′-diisocyanate and polyethylene glycol with a molecular mass of 1000 g/mole. The obtained nanobiocomposites were characterized by the small-angle X-ray scattering technique, scanning electron microscopy, Fourier infrared spectroscopy, thermogravimetry, and differential scanning calorimetry, and moreover, their selected mechanical properties, biodegradability, and cytotoxicity were tested. The effect of the organomodified montmorillonite presence in the biocomposites on their properties was investigated and compared to those of the native P3HB and the P3HB-PU composition. The obtained hybrid nanobiocomposites have an exfoliated structure. The presence and content of Cloisite 30B influence the P3HB-PU composition’s properties, and 2 wt.% Cloisite 30B leads to the best improvement in the aforementioned properties. The obtained results indicate that the thermal stability and mechanical properties of P3HB were improved, particularly in terms of increasing the degradation temperature, reducing hardness, and increasing impact strength, which were also confirmed by the morphological analysis of these bionanocomposites. However, the presence of organomodified montmorillonite in the obtained polymer biocomposites decreased their biodegradability slightly. The produced hybrid polymer nanobiocomposites have tailored mechanical and thermal properties and processing conditions for their expected application in the production of biodegradable, short-lived products for agriculture. Moreover, in vitro studies on human skin fibroblasts and keratinocytes showed their satisfactory biocompatibility and low cytotoxicity, which make them safe when in contact with the human body, for instance, in biomedical applications. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of nanocomposite production.</p>
Full article ">Figure 2
<p>SAXS plots of nanocomposites containing 1, 2, and 3 wt.% of Cloisite 30B and 10 wt.% of PU (designated as C10-1, C10-2, and C10-3, respectively). The figure shows the reference diffractograms of the unfilled P3HB and the polymer composition of P3HB and 10 wt.% PU and pure Cloisite 30B.</p>
Full article ">Figure 3
<p>Set of FTIR spectra of a native P3HB, PU, and P3HB-PU polymer biocomposition containing 10 wt.% PU and bionanocomposites based on the P3HB-PU polymer matrix with 1, 2, and 3 wt.% of Cloisite 30B (C10-1, C10-2, and C10-3, respectively).</p>
Full article ">Figure 4
<p>Scheme of hydrogen bond formation between P3HB and PU chains.</p>
Full article ">Figure 5
<p>SEM micrographs of P3HB (<b>a</b>) and its polymer biocomposition containing 10 wt.% polyurethane (P3HB-PU, C10) (<b>b</b>), and nanobiocomposites with 10 wt.% PU and 1 wt.% (<b>c</b>), 2 wt.% (<b>d</b>), and 3 wt. % of Cloisite 30B (<b>e</b>) (C10-1, C10-2, and C10-3, respectively).</p>
Full article ">Figure 6
<p>Graphs of the (<b>a</b>) impact strength, (<b>b</b>) hardness, (<b>c</b>) tensile strength, and (<b>d</b>) relative elongation at break of the extruded products—P3HB, P3HB-PU (C10) polymer composition, and polymer nanobiocomposites with 1, 2, and 3 wt.% Cloisite 30B (C10-1, C10-2, and C10-3, respectively).</p>
Full article ">Figure 7
<p>DSC thermal curves of P3HB, P3HB-PU (C10) polymer composition, and polymer nanobiocomposites with 1, 2, and 3 wt.% Cloisite 30B (C10-1, C10-2, and C10-3, respectively) upon heating the samples at 10 °C/min after prior cooling at the same rate.</p>
Full article ">Figure 8
<p>Microscopic images of human fibroblasts (BJs) and immortalized keratinocytes (HaCaTs) stained with crystal violet after 24 h incubation with the studied materials. Round images present the plate wells containing the stained cells, with brighter parts indicating the localization of the samples. The top and bottom rows show cells from the central parts of the reactivity zones.</p>
Full article ">Figure 9
<p>The viability of normal human fibroblasts (BJs) and immortalized human keratinocytes (HaCaTs) after 24 h incubation with the studied samples, estimated with the direct contact assay. Results are expressed as medians. The lower (25%) and upper (75%) quartile ranges are presented as whiskers. Asterisk * indicates differences between the control and samples (<span class="html-italic">p</span> &lt; 0.05, Kruskal–Wallis test). Symbol ▼ means significant differences between cell lines for a particular sample (<span class="html-italic">p</span> &lt; 0.05, Mann–Whitney U test).</p>
Full article ">
20 pages, 4050 KiB  
Article
Reversed Mg-Based Smectites: A New Approach for CO2 Adsorption
by Francisco Franco, Juan Antonio Cecilia, Laura Pardo, Salima Essih, Manuel Pozo, Lucía dos Santos-Gómez and Rosario M. P. Colodrero
Nanomaterials 2024, 14(18), 1532; https://doi.org/10.3390/nano14181532 - 21 Sep 2024
Viewed by 479
Abstract
Addressing climate change requires transitioning to cleaner energy sources and adopting advanced CO2 capture techniques. Clay minerals are effective in CO2 adsorption due to their regenerative properties. Recent advancements in nanotechnology further improve their efficiency and potential for use in carbon [...] Read more.
Addressing climate change requires transitioning to cleaner energy sources and adopting advanced CO2 capture techniques. Clay minerals are effective in CO2 adsorption due to their regenerative properties. Recent advancements in nanotechnology further improve their efficiency and potential for use in carbon capture and storage. This study examines the CO2 adsorption properties of montmorillonite and saponite, which are subjected to a novel microwave-assisted acid treatment to enhance their adsorption capacity. While montmorillonite shows minimal changes, saponite undergoes significant alterations. Furthermore, the addition of silica pillars to smectites results in a new nanomaterial with a higher surface area (653 m2 g−1), denoted as reversed smectite, with enhanced CO2 adsorption capabilities, potentially useful for electrochemical devices for converting captured CO2 into value-added products. Full article
(This article belongs to the Special Issue Design of Nanomaterials for Electrochemical Devices)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of starting samples: saponite (blue line) and montmorillonite (black line).</p>
Full article ">Figure 2
<p>XRD patterns of the initial sample and those for the materials obtained after 4, 8, 12, and 16 min of microwave-assisted acid treatment: (<b>a</b>) Mont and Mont-H x min, (<b>b</b>) Sap and Sap-H x min, (<b>c</b>) Mont-PCH series, (<b>d</b>) Sap-PCH series.</p>
Full article ">Figure 3
<p>Structures of the Sap or normal smectite (<b>a</b>), Sap-PCH (<b>b</b>), and SAP-PCH-H or “reversed smectite” (<b>c</b>) (modified from [<a href="#B29-nanomaterials-14-01532" class="html-bibr">29</a>]).</p>
Full article ">Figure 4
<p>Variation of MgO/SiO<sub>2</sub> (<b>a</b>), Al<sub>2</sub>O<sub>3</sub>/SiO<sub>2</sub> (<b>b</b>), and Fe<sub>2</sub>O<sub>3</sub>/SiO<sub>2</sub> (<b>c</b>) along the experiments: saponite (black squares), saponite-PCH (red squares), montmorillonite (black triangles), montmorillonite-PCH (red triangles).</p>
Full article ">Figure 5
<p>FTIR spectra for (<b>a</b>) Mont series and (<b>b</b>) Sap series.</p>
Full article ">Figure 6
<p>Selected SEM micrographs of the studied materials. (<b>a</b>) Starting montmorillonite. (<b>b</b>) Mont-PCH. (<b>c</b>) Starting saponite. (<b>d</b>) Sap-PCH. (<b>e</b>) Montmorillonite activated during 16 min. (<b>f</b>) Mont-PCH activated during 16 min. (<b>g</b>) Saponite activated during 16 min. (<b>h</b>) Sap-PCH activated during 16 min.</p>
Full article ">Figure 7
<p>N<sub>2</sub> adsorption-desorption isotherms at −196 °C of the raw smectites and their respective samples after the microwave-assisted acid treatment. (<b>a</b>) Montmorillonite series. (<b>b</b>) Saponite series. Filled squares, N<sub>2</sub> adsorption; Empty squares, N<sub>2</sub> desorption.</p>
Full article ">Figure 8
<p>Representation of the pore volume vs the microwave-assisted acid treatment time (<b>a</b>) and micropore volume vs the microwave-assisted acid treatment time (<b>b</b>) for Mont, Sap, Mont-PCH, and Sap-PCH.</p>
Full article ">Figure 9
<p>Isotherms of CO<sub>2</sub> adsorption on selected samples of (<b>a</b>) montmorillonite series and (<b>b</b>) saponite series. Filled squares, CO<sub>2</sub> adsorption; Empty squares, CO<sub>2</sub> desorption.</p>
Full article ">Figure 10
<p>Correlation between the microporosity and the CO<sub>2</sub> adsorption capacity.</p>
Full article ">Scheme 1
<p>Flowchart of the treatment and synthesis of Porous Clay Heterostructures.</p>
Full article ">
Back to TopTop