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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline

Search Results (113)

Search Parameters:
Keywords = organoclay

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 9544 KiB  
Article
Preparation and Characterization of Melamine Aniline Formaldehyde-Organo Clay Nanocomposite Foams (MAFOCF) as a Novel Thermal Insulation Material
by Ahmet Gürses and Elif Şahin
Polymers 2024, 16(24), 3578; https://doi.org/10.3390/polym16243578 - 21 Dec 2024
Viewed by 280
Abstract
The main purpose of this study is to prepare a melamine aniline formaldehyde foam, an MAF copolymer, with lower water sensitivity and non-flammability properties obtained by the condensation reaction of melamine, aniline, and formaldehyde. In addition, the preparation of MAFF composites with organoclay [...] Read more.
The main purpose of this study is to prepare a melamine aniline formaldehyde foam, an MAF copolymer, with lower water sensitivity and non-flammability properties obtained by the condensation reaction of melamine, aniline, and formaldehyde. In addition, the preparation of MAFF composites with organoclay reinforcement was determined as a secondary target in order to obtain better mechanical strength, heat, and sound insulation properties. For the synthesis of foams, the microwave irradiation technique, which offers advantages such as faster reactions, high yields and purities, and reduced curing times, was used together with the heating technique and the effect of organoclay content on the structural and textural properties of foams and both heat insulation and mechanical stability was investigated. Virgin melamine formaldehyde foam, MFF, melamine aniline formaldehyde foam, MAFFF, and melamine aniline formaldehyde–organoclay nanocomposite foams prepared with various organoclay contents, MAFOCFs, were characterized by HRTEM, FTIR, SEM, and XRD techniques. From spectroscopic and microscopic analyses, it was observed that organoclay flakes could be exfoliated without much change in the resin matrix with increasing clay content. In addition, it was determined that aniline formaldehyde, which is thought to enter the main polymer network as a bridge, caused textural changes in the polymeric matrix, and organoclay reinforcement also affected these changes. Although the highest compressive strength was obtained in MAFOCF5 foam with high organoclay content (0.40 MPa), it was determined that the compressive strengths in the nanocomposites were generally quite high despite their low bulk densities. In the prepared nanocomposite with 0.30% organoclay content (MAFOCF2), 0.33 MPa compressive strength and 0.051 thermal conductivity coefficient were measured. For virgin polymers and composites, bulk density, thermal conductivity, and compressive strength values were determined in the order of magnitude as MFF > MAFOCF1 > MAFOCF5 > MAFOCF6 > MAFF > MAFOCF3 > MAFOCF2 > MAFOCF4; MFF > MAFF > MAFOCF6 > MAFOCF5 > MAFOCF1 > MAFOCF4 > MAFOCF3 > MAFOCF2 and MAFOCF5 > MAFOCF4 > MAFOCF2 > MAFF > MAFOCF6 > MFF > MAFOCF1 > MAFOCF3. As a result, both compressive strength and thermal conductivity values indicate that nanocomposite foam with 0.20 wt% organoclay content can be a promising new insulation material. Full article
(This article belongs to the Special Issue Advances and Applications of Block Copolymers II)
Show Figures

Figure 1

Figure 1
<p>HRTEM images of virgin melamine formaldehyde foam (MFF) (<b>a</b>), virgin melamine aniline co-polymer foam (MAFF) (<b>b</b>), and melamine aniline copolymer organoclay nanocomposite foams (MAFOCFs1-6) (<b>c</b>–<b>h</b>).</p>
Full article ">Figure 1 Cont.
<p>HRTEM images of virgin melamine formaldehyde foam (MFF) (<b>a</b>), virgin melamine aniline co-polymer foam (MAFF) (<b>b</b>), and melamine aniline copolymer organoclay nanocomposite foams (MAFOCFs1-6) (<b>c</b>–<b>h</b>).</p>
Full article ">Figure 1 Cont.
<p>HRTEM images of virgin melamine formaldehyde foam (MFF) (<b>a</b>), virgin melamine aniline co-polymer foam (MAFF) (<b>b</b>), and melamine aniline copolymer organoclay nanocomposite foams (MAFOCFs1-6) (<b>c</b>–<b>h</b>).</p>
Full article ">Figure 2
<p>SEM patterns of virgin melamine formaldehyde foam (MFF) (<b>a</b>), virgin melamine aniline copolymer foam (MAFF) (<b>b</b>), and melamine aniline copolymer organoclay nanocomposite foams (MAFOCFs1-6) (<b>c</b>–<b>h</b>).</p>
Full article ">Figure 3
<p>FTIR analysis of virgin melamine formaldehyde foam (MFF) (<b>a</b>), virgin melamine aniline formaldehyde foam (MAFF) (<b>b</b>), and melamine aniline formaldehyde-organoclay nanocomposite foams (MAFOCFs1-6) (<b>c</b>–<b>h</b>).</p>
Full article ">Figure 4
<p>XRD diffractograms of virgin melamine formaldehyde foam (MF), virgin melamine aniline formaldehyde foam (MAFF), and melamine aniline formaldehyde–organoclay nanocomposite foams (MAFOCFs1-6).</p>
Full article ">
16 pages, 3008 KiB  
Article
Adsorption of Cr(VI) Using Organoclay/Alginate Hydrogel Beads and Their Application to Tannery Effluent
by Mayra X. Muñoz-Martinez, Iván F. Macías-Quiroga and Nancy R. Sanabria-González
Gels 2024, 10(12), 779; https://doi.org/10.3390/gels10120779 - 28 Nov 2024
Viewed by 616
Abstract
The tanning industry is among the most environmentally harmful activities globally due to the pollution of lakes and rivers from its effluents. Hexavalent chromium, a metal in tannery effluents, has adverse effects on human health and ecosystems, requiring the development of removal techniques. [...] Read more.
The tanning industry is among the most environmentally harmful activities globally due to the pollution of lakes and rivers from its effluents. Hexavalent chromium, a metal in tannery effluents, has adverse effects on human health and ecosystems, requiring the development of removal techniques. This study assessed the efficacy of organobentonite/alginate hydrogel beads in removing Cr(VI) from a fixed-bed adsorption column system. The synthesized organobentonite (OBent) was encapsulated in alginate, utilizing calcium chloride as a crosslinking agent to generate hydrogel beads. The effects of the volumetric flow rate, bed height, and initial Cr(VI) concentration on a synthetic sample were analyzed in the experiments in fixed-bed columns. The fractal-like modified Thomas model showed a good fit to the experimental data for the asymmetric breakthrough curves, confirmed by the high R2 correlation coefficients and low χ2 values. The application of organoclay/alginate hydrogel beads was confirmed with a wastewater sample from an artisanal tannery industry in Belén (Nariño, Colombia), in which a Cr(VI) removal greater than 99.81% was achieved. Organobentonite/alginate hydrogels offer the additional advantage of being composed of a biodegradable polymer (sodium alginate) and a natural material (bentonite-type clay), resulting in promising adsorbents for the removal of Cr(VI) from aqueous solutions in both synthetic and real water samples. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Color changes of the prepared hydrogel beads with different organoclay concentrations.</p>
Full article ">Figure 2
<p>X-ray diffraction patterns of alginate, sodium bentonite, organobentonite, and the hydrogel beads with 2 wt.% organoclay.</p>
Full article ">Figure 3
<p>Size distribution (diameter) of the organobentonite/alginate hydrogel beads.</p>
Full article ">Figure 4
<p>Effect of the organobentonite concentration in the hydrogel beads on Cr(VI) removal. Conditions: [Cr(VI)] = 25 mg/L, V = 50 mL, pH = 3.4, stirring speed = 150 rpm, organobentonite mass in hydrogel beads = 220 ± 3 mg, T = 20 ± 1 °C.</p>
Full article ">Figure 5
<p>Effect of pH on Cr(VI) removal by OBent (2%)/Alg hydrogel beads. Conditions: [Cr(VI)] = 25 mg/L, V = 50 mL, stirring speed = 150 rpm, organobentonite mass in the hydrogel beads = 220 ± 3 mg, T = 20 ± 1 °C.</p>
Full article ">Figure 6
<p>Breakthrough curves for Cr(VI) adsorption on OBent (2%)/Alg at different conditions (Dash line + symbol) and fits to the fractal-like modified Thomas model. (<b>a</b>) Effect of the volumetric flow rate, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> = 20 mg/L, pH = 3.4, and <math display="inline"><semantics> <mrow> <mi>h</mi> </mrow> </semantics></math> = 10 cm; (<b>b</b>) effect of the bed height, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> = 20 mg/L, pH = 3.4, and <math display="inline"><semantics> <mrow> <mi>Q</mi> </mrow> </semantics></math> = 3 mL/min; and (<b>c</b>) effect of the initial Cr(VI) concentration, pH = 3.4, <math display="inline"><semantics> <mrow> <mi>Q</mi> </mrow> </semantics></math> = 3.0 mL/min, and h = 15 cm.</p>
Full article ">Figure 7
<p>(<b>a</b>) Leather tanning tank; (<b>b</b>) sample of tannery wastewater.</p>
Full article ">Figure 8
<p>Successive Cr(VI) adsorption cycles on OBent (2%)/Alg hydrogel beads.</p>
Full article ">Figure 9
<p>Appearance of hydrogel beads. (<b>a</b>) Before and after Cr(III) removal; (<b>b</b>) before and after Cr(VI) removal.</p>
Full article ">
16 pages, 4165 KiB  
Article
Sorption Properties of Bentonite-Based Organoclays with Amphoteric and Nonionic Surfactants in Relation to Polycyclic Aromatic Hydrocarbons
by Tamara Dudnikova, Marina Burachevskaya, Tatyana Minkina, Saglara Mandzhieva, Inna Zamulina, Leonid Perelomov and Maria Gertsen
Minerals 2024, 14(11), 1132; https://doi.org/10.3390/min14111132 - 8 Nov 2024
Viewed by 643
Abstract
Polycyclic aromatic hydrocarbons (PAHs) are a major scientific challenge due to their profound impact on public and environmental health. Therefore, studying ways to detoxify PAHs is important. In this research, the adsorption ability of bentonite modified with five surfactants, including amphoteric (cocoamphodiacetate disodium [...] Read more.
Polycyclic aromatic hydrocarbons (PAHs) are a major scientific challenge due to their profound impact on public and environmental health. Therefore, studying ways to detoxify PAHs is important. In this research, the adsorption ability of bentonite modified with five surfactants, including amphoteric (cocoamphodiacetate disodium and sodium cocoiminodipropionate) and nonionic (lauramine oxide, cocamide diethanolamine, and alkylpolyglucoside) substances for the adsorption of high-molecular benzo(a)pyrene and low-molecular naphthalene from the PAH group was studied. The bentonite and bentonite-based organoclays were characterized using X-ray diffraction and Fourier transform infrared spectroscopy. The results showed that the maximum adsorption of benzo(a)pyrene by organoclays increased compared with the initial mineral. The adsorption of benzo(a)pyrene is higher than that of naphthalene. The adsorption process of benzo(a)pyrene by bentonite and organoclays is predominantly monolayer, as it is better described by the Langmuir model (R2 0.77–0.98), while naphthalene is predominantly multilayer, described by the Freundlich model (R2 0.86–0.96). According to the effectiveness of sorption capacities of organoclays—including the degree of sorption, Langmuir and Freundlich constants, the value of maximum adsorption, Gibbs free energy, and the index of favorability of the adsorption process—the most effective modification was found. For the adsorption of benzo(a)pyrene the best was cocoamphodiacetate disodium, and for naphthalene it was sodium cocoiminodipropionate. Full article
(This article belongs to the Special Issue Organo-Clays: Preparation, Characterization and Applications)
Show Figures

Figure 1

Figure 1
<p>Results of X-ray diffractometric analysis of commercial bentonite used for organoclay synthesis.</p>
Full article ">Figure 2
<p>Results of IR spectroscopic analysis of bentonite and organoclays synthesized with various surfactants. * Surfactant 1—sodium cocoiminodipropionate, Surfactant 2—lauramine oxide, Surfactant 3—cocamide diethethanolamine, Surfactant 4—disodium cocoamphodiacetate, Surfactant 5—alkylpolyglucoside.</p>
Full article ">Figure 3
<p>Isotherms of adsorption of benzo(a)pyrene (<b>A</b>) and naphthalene (<b>B</b>) by bentonite and organoclays by various surfactants using the Langmuir equation (continuous line) and Freundlich equation (dotted line). * Surfactant 1—sodium cocoiminodipropionate, Surfactant 2—lauramine oxide, Surfactant 3—cocamide diethethanolamine, Surfactant 4—disodium cocoamphodiacetate, Surfactant 5—alkylpolyglucoside.</p>
Full article ">
32 pages, 4197 KiB  
Article
Chlorophyll-Amended Organoclays for the Detoxification of Ochratoxin A
by Johnson O. Oladele, Meichen Wang, Xenophon Xenophontos, Kendall Lilly, Phanourios Tamamis and Timothy D. Phillips
Toxins 2024, 16(11), 479; https://doi.org/10.3390/toxins16110479 - 6 Nov 2024
Viewed by 775
Abstract
Climate change has been associated with outbreaks of mycotoxicosis following periods of drought, enhanced fungal growth, and increased exposure to mycotoxins. For detoxification, the inclusion of clay-based materials in food and drinking water has resulted in a very promising strategy to reduce mycotoxin [...] Read more.
Climate change has been associated with outbreaks of mycotoxicosis following periods of drought, enhanced fungal growth, and increased exposure to mycotoxins. For detoxification, the inclusion of clay-based materials in food and drinking water has resulted in a very promising strategy to reduce mycotoxin exposure. In this strategy, mycotoxins are tightly sorbed to high-affinity clay particles in the gastrointestinal tract, thus decreasing bioavailability, uptake to blood, and potential toxicity. This study investigated the ability of chlorophyll and chlorophyllin-amended montmorillonite clays to decrease the toxicity of ochratoxin A (OTA). The sorption mechanisms of OTA binding to surfaces of sorbents, as well as binding parameters such as capacity, affinity, enthalpy, and free energy, were examined. Chlorophyll-amended organoclay (CMCH) demonstrated the highest binding (72%) and was better than the chlorophyllin-amended hydrophilic clay (59%), possibly due to the hydrophobicity of OTA (LogP 4.7). In silico studies using molecular dynamics simulations showed that CMCH improves OTA binding in comparison to parent clay in line with experiments. Simulations depicted that chlorophyll amendments on clay facilitated OTA molecules binding both directly, through enhancing OTA binding on the clay, or predominantly indirectly, through OTA molecules interacting with bound chlorophyll amendments. Simulations uncovered the key role of calcium ions in OTA binding, particularly in neutral conditions, and demonstrated that CMCH binding to OTA is enhanced under both neutral and acidic conditions. Furthermore, the protection of various sorbents against OTA-induced toxicity was carried out using two living organisms (Hydra vulgaris and Caenorhabditis elegans) which are susceptible to OTA toxicity. This study showed the significant detoxification of OTA (33% to 100%) by inclusion of sorbents. Organoclay (CMCH) at 0.5% offered complete protection. These findings suggest that the chlorophyll-amended organoclays described in this study could be included in food and feed as OTA binders and as potential filter materials for water and beverages to protect against OTA contaminants during outbreaks and emergencies. Full article
(This article belongs to the Special Issue Toxins: 15th Anniversary)
Show Figures

Figure 1

Figure 1
<p>Sorbent screening with reduction percentage of ochratoxin by the sorbents. * <span class="html-italic">p</span> &lt; 0.01 when compared to CM or SM. CM: calcium montmorillonite; CMCH: chlorophyll-amended calcium montmorillonite; CMCHin: chlorophyllin-amended calcium montmorillonite; SM: sodium montmorillonite; SMCH: chlorophyll-amended sodium montmorillonite; SMCHin: chlorophyllin-amended sodium montmorillonite.</p>
Full article ">Figure 2
<p>Adsorption isotherms of OTA onto binding surfaces of (<b>A</b>) CM-amended clays and (<b>B</b>) SM-amended clays at pH 2. CM: calcium montmorillonite; CMCH: chlorophyll-amended calcium montmorillonite; CMCHin: chlorophyllin-amended calcium montmorillonite; SM: sodium montmorillonite; SMCH: chlorophyll-amended sodium montmorillonite; SMCHin: chlorophyllin-amended sodium montmorillonite.</p>
Full article ">Figure 3
<p>Desorption isotherms of OTA onto binding surfaces of (<b>A</b>) CM-amended clays and (<b>B</b>) SM-amended clays at pH 6. CM: calcium montmorillonite; CMCH: chlorophyll-amended calcium montmorillonite; CMCHin: chlorophyllin-amended calcium montmorillonite; SM: sodium montmorillonite; SMCH: chlorophyll-amended sodium montmorillonite; SMCHin: chlorophyllin-amended sodium montmorillonite.</p>
Full article ">Figure 4
<p>Effect of contact time on the adsorption of OTA. CM: calcium montmorillonite; CMCH: chlorophyll-amended calcium montmorillonite; CMCHin: chlorophyllin-amended calcium montmorillonite; SMCH: chlorophyll-amended sodium montmorillonite.</p>
Full article ">Figure 5
<p>Toxicity effects of OTA exposure to hydra (<b>A</b>), protection with SM-derived sorbents (<b>B</b>), and CM-derived sorbents (<b>C</b>). CM: calcium montmorillonite; CMCH: chlorophyll-amended calcium montmorillonite; CMCHin: chlorophyllin-amended calcium montmorillonite; SM: sodium montmorillonite; SMCH: chlorophyll-amended sodium montmorillonite; SMCHin: chlorophyllin-amended sodium montmorillonite.</p>
Full article ">Figure 5 Cont.
<p>Toxicity effects of OTA exposure to hydra (<b>A</b>), protection with SM-derived sorbents (<b>B</b>), and CM-derived sorbents (<b>C</b>). CM: calcium montmorillonite; CMCH: chlorophyll-amended calcium montmorillonite; CMCHin: chlorophyllin-amended calcium montmorillonite; SM: sodium montmorillonite; SMCH: chlorophyll-amended sodium montmorillonite; SMCHin: chlorophyllin-amended sodium montmorillonite.</p>
Full article ">Figure 6
<p>Effect of toxicity of OTA on the body length (<b>A</b>), nose touch response (<b>B</b>), and survival rate of <span class="html-italic">Caenorhabditis elegans</span> after 24 h and 48 h of exposure (<b>C</b>). Data represent the average value from triplicate analysis  ±  the standard deviation. * indicates a significant difference (<span class="html-italic">p</span>  ≤  0.05) compared to the vehicle control group.</p>
Full article ">Figure 7
<p>Protective effect of parent clays and amended clays against OTA toxicity on the body length (<b>A</b>,<b>D</b>), nose touch response (<b>B</b>,<b>E</b>), and survival rate of <span class="html-italic">Caenorhabditis elegans</span> (<b>C</b>,<b>F</b>). Data represent the average value from triplicate analysis  ±  the standard deviation. * indicates a significant difference (<span class="html-italic">p</span>  ≤  0.05) compared to the vehicle control group. # indicates a significant difference (<span class="html-italic">p</span>  ≤  0.05) compared to the OTA-alone group. 1: 0.2% and 2: 0.5% clay inclusions; CM: calcium montmorillonite; CMCH: chlorophyll-amended calcium montmorillonite; CMCHin: chlorophyllin-amended calcium montmorillonite; SM: sodium montmorillonite; SMCH: chlorophyll-amended sodium montmorillonite; SMCHin: chlorophyllin-amended sodium montmorillonite.</p>
Full article ">Figure 7 Cont.
<p>Protective effect of parent clays and amended clays against OTA toxicity on the body length (<b>A</b>,<b>D</b>), nose touch response (<b>B</b>,<b>E</b>), and survival rate of <span class="html-italic">Caenorhabditis elegans</span> (<b>C</b>,<b>F</b>). Data represent the average value from triplicate analysis  ±  the standard deviation. * indicates a significant difference (<span class="html-italic">p</span>  ≤  0.05) compared to the vehicle control group. # indicates a significant difference (<span class="html-italic">p</span>  ≤  0.05) compared to the OTA-alone group. 1: 0.2% and 2: 0.5% clay inclusions; CM: calcium montmorillonite; CMCH: chlorophyll-amended calcium montmorillonite; CMCHin: chlorophyllin-amended calcium montmorillonite; SM: sodium montmorillonite; SMCH: chlorophyll-amended sodium montmorillonite; SMCHin: chlorophyllin-amended sodium montmorillonite.</p>
Full article ">Figure 8
<p>Average (%) probability of interaction between OTA molecules in the presence of CMCH, CMPHO, and CM. Blue corresponds to acidic conditions; red and yellow correspond to neutral conditions, simulating monoanionic and dianionic OTA, respectively. Average values are calculated from triplicate runs. Error bars denote standard deviation values calculated from the triplicate runs.</p>
Full article ">Figure 9
<p>Average (%) probability of OTA molecules interacting with CMCH, CMPHO, and CM in acidic condition (<b>A</b>) and neutral condition (<b>B</b>,<b>C</b>), simulating monoanionic and dianionic OTA, respectively. Direct, direct-assisted, and indirect-assisted interactions are shown in blue, red, and yellow, respectively. Average values are calculated from triplicate runs. Error bars denote standard deviation values calculated from the triplicate runs.</p>
Full article ">Figure 9 Cont.
<p>Average (%) probability of OTA molecules interacting with CMCH, CMPHO, and CM in acidic condition (<b>A</b>) and neutral condition (<b>B</b>,<b>C</b>), simulating monoanionic and dianionic OTA, respectively. Direct, direct-assisted, and indirect-assisted interactions are shown in blue, red, and yellow, respectively. Average values are calculated from triplicate runs. Error bars denote standard deviation values calculated from the triplicate runs.</p>
Full article ">Figure 10
<p>Average (%) probability of CM participating in interactions with different OTA groups (defined in <a href="#app1-toxins-16-00479" class="html-app">Figure S2</a>). Group 1 corresponds to blue, group 2 corresponds to red, group 3 corresponds to yellow, group 4 corresponds to green, and group 5 corresponds to orange. Additionally, the average (%) probability of CM-bound OTA molecules interacting with calcium is shown in cyan. The results were normalized, i.e., they were calculated given an interaction between OTA and CM. Values correspond to parent (CM) and amended clays (CMCH and CMPHO) in acidic conditions (left), as well as neutral conditions (middle) and (right), of simulations including monoanionic and dianionic OTA, respectively. Average values are calculated from triplicate runs. Error bars denote standard deviation values calculated from the triplicate runs.</p>
Full article ">Figure 11
<p>Average (%) probability of CMCH (<b>A</b>) and CMPHO (<b>B</b>), respectively, participating in interactions with different OTA groups (defined in <a href="#app1-toxins-16-00479" class="html-app">Figure S2</a>). Group 1 corresponds to blue, group 2 corresponds to red, group 3 corresponds to yellow, group 4 corresponds to green, and group 5 corresponds to orange. Additionally, the average (%) probability of CMCH-bound (<b>A</b>) or CMPHO-bound (<b>B</b>) OTA molecules interacting with calcium is shown in cyan. The results were normalized, i.e., they were calculated given an interaction between OTA and CMCH and CMPHO. Values correspond to systems in acidic conditions (left), as well as neutral conditions (middle) and (right), of simulations including monoanionic and dianionic OTA, respectively. Average values are calculated from triplicate runs. Error bars denote standard deviation values calculated from the triplicate runs. All values shown above are normalized over the total number of interactions per system.</p>
Full article ">Figure 12
<p>Simulation snapshots of CMCH in complex with OTA in (<b>A</b>) acidic conditions, as well as (<b>B</b>,<b>C</b>) neutral conditions, investigating OTA in monoanionic and dianionic states, respectively. Panels (<b>D</b>–<b>F</b>) show zoomed-in representation of particular interactions, marked as (i) and (ii), that occur within panels (<b>A</b>–<b>C</b>), respectively. CM, chlorophyll, and calcium are shown in vdW representation, while OTA is shown in licorice representation. Atoms are colored by atom type, except for carbon atoms of chlorophyll in green and calcium in tan. Calcium ions that are at a greater distance than 3.5 Å from all OTA molecules were omitted. Hydrogen atoms are also omitted for clarity.</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 890
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 ">
20 pages, 4197 KiB  
Article
Removal of Lead Cations by Novel Organoclays Derived from Bentonite and Amphoteric and Nonionic Surfactants
by Maria Gertsen, Leonid Perelomov, Anna Kharkova, Marina Burachevskaya, S. Hemalatha and Yury Atroshchenko
Toxics 2024, 12(10), 713; https://doi.org/10.3390/toxics12100713 - 30 Sep 2024
Cited by 1 | Viewed by 1126
Abstract
For many decades, natural and modified clay minerals have been used as adsorbents to clean up aquatic and soil ecosystems contaminated with organic and inorganic pollutants. In this study, organoclays based on bentonite and various amphoteric and nonionic surfactants were synthesized and tested [...] Read more.
For many decades, natural and modified clay minerals have been used as adsorbents to clean up aquatic and soil ecosystems contaminated with organic and inorganic pollutants. In this study, organoclays based on bentonite and various amphoteric and nonionic surfactants were synthesized and tested as effective sorbents for lead ions. The maximum values of R were obtained when describing the sorption processes using the Langmuir model, which ranged from 0.97 to 0.99. The adsorption of lead ions by these organoclays was investigated using different sorption models including the Langmuir, Freundlich, and BET. It was found that, according to the values of limiting adsorption to the Langmuir equation, the synthesized organoclays formed an increasing series: organoclay with cocamide diethanolamine < bentonite < organoclay with lauramine oxide < organoclay with sodium cocoiminodipropionate < organoclay with disodium cocoamphodiacetate < organoclay with alkyl polyglucoside. The Gibbs energy for all of the analyzed samples was calculated and found to be negative, indicating the spontaneity of the cation adsorption process in the forward direction. The maximum value of the adsorption capacity of lead cations on organoclay-based bentonite with alkyl polyglucoside was 1.49 ± 0.05 mmol/g according to the Langmuir model, and 0.523 ± 0.003 mmol/g as determined by the BET model. In the process of modifying bentonite, there was an increase in negative values of the zeta potential for organoclays compared to the initial mineral, which clearly enhanced their electrostatic interactions with the positively charged lead ions. It was hypothesized, based on the physicochemical principles, that exchange adsorption is the main mechanism for lead absorption. Based on chemical approaches, organoclays based on amphoteric surfactants absorb lead mainly through the mechanisms of electrostatic attraction, ion exchange, and complexation as well as the formation of insoluble precipitates. Organoclays based on nonionic surfactants, on the other hand, absorb lead through mechanisms of complexation (including chelation) and the formation of insoluble chemical precipitates. The comparison of isotherms from different models allows us to find the most accurate match between the model and the experimental data, and to better understand the nature of the processes involved. Full article
Show Figures

Figure 1

Figure 1
<p>The results of the X-ray diffraction analysis of bentonite used in the synthesis of organoclays.</p>
Full article ">Figure 2
<p>IR spectra of the synthesized organoclays using amphoteric surfactants.</p>
Full article ">Figure 3
<p>IR spectra of the synthesized organoclays using nonionic surfactants.</p>
Full article ">Figure 4
<p>Adsorption isotherms of Pb<sup>2+</sup> ions on the initial and surfactant-modified bentonite forms: (<b>a</b>) organoclays modified with amphoteric surfactants; (<b>b</b>) organoclays modified with nonionic surfactants.</p>
Full article ">Figure 5
<p>Adsorption isotherms of Pb<sup>2+</sup> ions on the initial and surfactant-modified forms of bentonite: (<b>a</b>) organoclays modified with amphoteric surfactants; (<b>b</b>) organoclays modified with nonionic surfactants.</p>
Full article ">Figure 6
<p>Adsorption isotherms of Pb<sup>2+</sup> ions on the original and surfactant-modified forms of bentonite: (<b>a</b>) organoclays modified with amphoteric surfactants; (<b>b</b>) organoclays modified with nonionic surfactants.</p>
Full article ">Figure 7
<p>Schematic illustration of the possible mechanism for the adsorption of Pb<sup>2+</sup> onto organoclays modified with amphoteric and nonionic surfactants: 1—ion exchange; 2—complexation; 3—precipitation; 4—long-range forces (physical adsorption).</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 718
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 ">
9 pages, 4541 KiB  
Article
Mechanical Properties of Clay-Reinforced Polyamide 6 Nanocomposite Liner Materials of Type IV Hydrogen Storage Vessels
by Dávid István Kis, Attila Bata, János Takács and Eszter Kókai
Nanomaterials 2024, 14(17), 1385; https://doi.org/10.3390/nano14171385 - 25 Aug 2024
Viewed by 884
Abstract
This study focuses on polyamide 6/organo-modified montmorillonite (PA6/OMMT) nanocomposites as potential liner materials, given the growing interest in enhancing the performance of type IV composite overwrapped hydrogen storage pressure vessels. The mechanical properties of PA6/OMMT composites with varying filler concentrations were investigated across [...] Read more.
This study focuses on polyamide 6/organo-modified montmorillonite (PA6/OMMT) nanocomposites as potential liner materials, given the growing interest in enhancing the performance of type IV composite overwrapped hydrogen storage pressure vessels. The mechanical properties of PA6/OMMT composites with varying filler concentrations were investigated across a temperature range relevant to hydrogen storage conditions (−40 °C to +85 °C). Liner collapse, a critical issue caused by rapid gas discharge, was analyzed using an Ishikawa diagram to identify external and internal factors. Mechanical testing revealed that higher OMMT content generally increased stiffness, especially at elevated temperatures. The Young’s modulus and first yield strength exhibited non-linear temperature dependencies, with 1 wt. per cent OMMT content enhancing yield strength at all tested temperatures. Dynamic mechanical analysis (DMA) indicated that OMMT improves the storage modulus, suggesting effective filler dispersion, but it also reduces the toughness and heat resistance, as evidenced by lower glass transition temperatures. This study underscores the importance of optimizing OMMT content to balance mechanical performance and thermal stability for the practical application of PA6/OMMT nanocomposites in hydrogen storage pressure vessels. Full article
(This article belongs to the Section Energy and Catalysis)
Show Figures

Figure 1

Figure 1
<p>Ishikawa diagram with the effecting factors of liner collapse.</p>
Full article ">Figure 2
<p>Tensile specimens after testing between −40 °C and 85 °C.</p>
Full article ">Figure 3
<p>Young’s modulus, first yield strength, and area under tensile curve vs. temperature values of (<b>a</b>) neat PA6; (<b>b</b>) PA6/OMMT-1 per cent; (<b>c</b>) PA6/OMMT-2,5 per cent; (<b>d</b>) PA6/OMMT-5 per cent; (<b>e</b>) PA6/OMMT-10 per cent; and (<b>f</b>) Young’s modulus plotted against clay content.</p>
Full article ">Figure 4
<p>Results by DMA measurement of neat PA6 and PA6/OMMT composites: (<b>a</b>) storage modulus, (<b>b</b>) loss modulus, (<b>c</b>) damping as a function of temperature, (<b>d</b>) Cole–Cole plot of loss modulus vs. storage modulus.</p>
Full article ">Figure 5
<p>Tg originated from the α relaxation peak of loss modulus and tan δ curves.</p>
Full article ">
23 pages, 3480 KiB  
Article
Influence of Montmorillonite Organoclay Fillers on Hygrothermal Response of Pultruded E-Glass/Vinylester Composites
by Vistasp M. Karbhari
Polymers 2024, 16(15), 2157; https://doi.org/10.3390/polym16152157 - 29 Jul 2024
Viewed by 768
Abstract
Pultruded fiber reinforced polymer composites used in civil, power, and offshore/marine applications use fillers as resin extenders and for process efficiency. Although the primary use of fillers is in the form of an extender and processing aid, the appropriate selection of filler can [...] Read more.
Pultruded fiber reinforced polymer composites used in civil, power, and offshore/marine applications use fillers as resin extenders and for process efficiency. Although the primary use of fillers is in the form of an extender and processing aid, the appropriate selection of filler can result in enhancing mechanical performance characteristics, durability, and multifunctionality. This is of special interest in structural and high voltage applications where the previous use of specific fillers has been at levels that are too low to provide these enhancements. This study investigates the use of montmorillonite organoclay fillers of three different particle sizes as substitutes for conventional CaCO3 fillers with the intent of enhancing mechanical performance and hygrothermal durability. The study investigates moisture uptake and kinetics and reveals that uptake is well described by a two-stage process that incorporates both a diffusion dominated initial phase and a second slower phase representing relaxation and deterioration. The incorporation of the organoclay particles substantially decreases uptake levels in comparison to the use of CaCO3 fillers while also enhancing stage I, diffusion, dominated stability, with the use of the 1.5 mm organoclay fillers showing as much as a 41.5% reduction in peak uptake as compared to the CaCO3 fillers at the same 20% loading level (by weight of resin). The mechanical performance was characterized using tension, flexure, and short beam shear tests. The organoclay fillers showed a significant improvement in each, albeit with differences due to particle size. Overall, the best performance after exposure to four different temperatures of immersion in deionized water was shown by the 4.8 mm organoclay filler-based E-glass/vinylester composite system, which was the only one to have less than a 50% deterioration over all characteristics after immersion for a year in deionized water at the highest temperature investigated (70 °C). The fillers not only enhance resistance to uptake but also increase tortuosity in the path, thereby decreasing the overall effect of uptake. The observations demonstrate that the use of the exfoliated organoclay particles with intercalation, which have been previously used in very low amounts, and which are known to be beneficial in relation to enhanced thermal stability, flame retardancy, and decreased flammability, provide enhanced mechanical characteristics, decreased moisture uptake, and increased hygrothermal durability when used at particle loading levels comparable to those of conventional fillers, suggesting that these novel systems could be considered for critical structural applications. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Moisture uptake in deionized water at 23 °C. (<b>b</b>) Moisture uptake in deionized water at 40 °C. (<b>c</b>) Moisture uptake in deionized water at 55 °C. (<b>d</b>) Moisture uptake in deionized water at 70 °C. Dashed lines show predictions of the two-stage model.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) Moisture uptake in deionized water at 23 °C. (<b>b</b>) Moisture uptake in deionized water at 40 °C. (<b>c</b>) Moisture uptake in deionized water at 55 °C. (<b>d</b>) Moisture uptake in deionized water at 70 °C. Dashed lines show predictions of the two-stage model.</p>
Full article ">Figure 2
<p>Schematic of moisture uptake models.</p>
Full article ">Figure 3
<p>Maximum uptake as a function of filler and temperature of immersion.</p>
Full article ">Figure 4
<p>Diffusion coefficient as a function of filler type and temperature of immersion.</p>
Full article ">Figure 5
<p>Uptake transition ratio as a function of filler type and temperature of immersion.</p>
Full article ">Figure 6
<p>Relaxation/deterioration coefficient as a function of filler type and temperature of immersion.</p>
Full article ">Figure 7
<p>Normalized tensile strength as a function of time and temperature of immersion in deionized water.</p>
Full article ">Figure 8
<p>Comparison of tensile strength variation as a function of filler type at different temperatures of immersion. Scale is in MPa.</p>
Full article ">Figure 9
<p>Normalized tensile modulus as a function of time and temperature of immersion in deionized water.</p>
Full article ">Figure 10
<p>Normalized flexural strength as a function of time and temperature of immersion in deionized water.</p>
Full article ">Figure 11
<p>Normalized longitudinal SBS strength as a function of time and temperature of immersion in deionized water.</p>
Full article ">Figure 12
<p>Normalized transverse SBS strength as a function of time and temperature of immersion in deionized water.</p>
Full article ">Figure 13
<p>Comparison of SBS strength variation as a function of filler type at different temperatures of immersion. Scale is in MPa.</p>
Full article ">
13 pages, 1335 KiB  
Article
New Methodology for Modifying Sodium Montmorillonite Using DMSO and Ethyl Alcohol
by Adriana Stoski, Bruno Rafael Machado, Bruno Henrique Vilsinski, Lee Marx Gomes de Carvalho, Edvani Curti Muniz and Carlos Alberto Policiano Almeida
Materials 2024, 17(12), 3029; https://doi.org/10.3390/ma17123029 - 20 Jun 2024
Viewed by 823
Abstract
Modified clays with organic molecules have many applications, such as the adsorption of pollutants, catalysts, and drug delivery systems. Different methodologies for intercalating these structures with organic moieties can be found in the literature with many purposes. In this paper, a new methodology [...] Read more.
Modified clays with organic molecules have many applications, such as the adsorption of pollutants, catalysts, and drug delivery systems. Different methodologies for intercalating these structures with organic moieties can be found in the literature with many purposes. In this paper, a new methodology of modifying Sodium Montmorillonite clays (Na-Mt) with a faster drying time was investigated by X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), BET, and thermogravimetric analysis (TG and DTG). In the modification process, a mixture of ethyl alcohol, DMSO, and Na-Mt were kept under magnetic stirring for one hour. Statistical analysis was applied to evaluate the effects of the amount of DMSO, temperature, and sonication time on the modified clay (DMSO-SMAT) using a 23-factorial design. XRD and FTIR analyses showed the DMSO intercalation into sodium montmorillonite Argel-T (SMAT). An average increase of 0.57 nm for the interplanar distance was found after swelling with DMSO intercalation. BET analysis revealed a decrease in the surface area (from 41.8933 m2/g to 2.1572 m2/g) of Na-Mt when modified with DMSO. The porosity increased from 1.74 (SMAT) to 1.87 nm (DMSO-SMAT) after the application of the methodology. Thermal analysis showed a thermal stability for the DMSO-SMAT material, and this was used to calculate the DMSO-SMAT formula of Na[Al5Mg]Si12O30(OH)6 · 0.54 DMSO. Statistical analysis showed that only the effect of the amount of DMSO was significant for increasing the interlayer space of DMSO-SMAT. In addition, at room temperature, the drying time of the sample using this methodology was 30 min. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of unmodified Mt (SMAT) and modified Mt (DMSO-SMAT) samples.</p>
Full article ">Figure 2
<p>FTIR spectra of modified Mt (DMSO-SMAT) and unmodified Mt (SMAT) samples with emphasis on the region between 960 and 900 cm<sup>−1</sup>.</p>
Full article ">Figure 3
<p>TG and DTG curves for modified Mt (DMSO-SMAT) and unmodified Mt (SMAT) samples.</p>
Full article ">Figure 4
<p>Interaction effects between: (<b>a</b>) temperature and amount of the DMSO; (<b>b</b>) temperature and sonication time and (<b>c</b>) amount of the DMSO and sonication time for the 2<sup>3</sup>-factorial designs for interplanar space increase.</p>
Full article ">Figure 5
<p>Appearance of the samples using the procedure of <a href="#sec2dot2-materials-17-03029" class="html-sec">Section 2.2</a>: (<b>a</b>) 10 min and (<b>b</b>) 30 min of drying with ethanol; (<b>c</b>) 10 min and (<b>d</b>) 30 min without ethanol.</p>
Full article ">
16 pages, 19173 KiB  
Article
Synthesis and Characterization of Nanocomposite Hydrogels Based on Poly(Sodium 4-Styrene Sulfonate) under Very-High Concentration Regimen of Clays (Bentonite and Kaolinite)
by Tulio A. Lerma, Enrique M. Combatt and Manuel Palencia
Gels 2024, 10(6), 405; https://doi.org/10.3390/gels10060405 - 18 Jun 2024
Cited by 1 | Viewed by 944
Abstract
The aim of this work was to synthesize and study the functional properties of polymer-clay nanocomposite (PCNCs) based on poly(sodium 4-styrene sulfonate) (NaPSS) and two types of clay in the dispersed phase: bentonite and kaolinite, in order to advance in the development of [...] Read more.
The aim of this work was to synthesize and study the functional properties of polymer-clay nanocomposite (PCNCs) based on poly(sodium 4-styrene sulfonate) (NaPSS) and two types of clay in the dispersed phase: bentonite and kaolinite, in order to advance in the development of new geomimetic materials for agricultural and environmental applications. In this study, the effect of adding high concentrations of clay (10–20 wt. %) on the structural and functional properties of a polymer–clay nanocomposite was evaluated. The characterization by infrared spectroscopy made it possible to show that the PCNCs had a hybrid nature structure through the identification of typical vibration bands of the clay matrix and NaPSS. In addition, scanning electron microscopy allowed us to verify its hybrid composition and an amorphous particle-like morphology. The thermal characterization showed degradation temperatures higher than ~300 °C with Tg values higher than 100 °C and variables depending on the clay contents. In addition, the PCNCs showed a high water-retention capacity (>2900%) and cation exchange capacity (>112 meq/100 g). Finally, the results demonstrated the ability of geomimetic conditioners to mimic the structure and functional properties of soils, suggesting their potential application in improving soil quality for plant growth. Full article
Show Figures

Figure 1

Figure 1
<p>IR-ATR spectra of bentonite (dashed line) and bentonite–tClVS (solid line) (<b>A</b>) and kaolinite (dashed line) and kaolinite–tClVS (solid line) (<b>B</b>). IR-ATR-FEDS spectra of bentonite–tClVS (<b>C</b>) and kaolinite–tClVS (<b>D</b>) between 2500 and 3800 cm<sup>−1</sup>.</p>
Full article ">Figure 2
<p>SEM images (left imagen), EDS spectrum for the determination of elemental composition (bottom right image, EDS analysis zone corresponds to the cross enclosed in the red circle in the SEM image) and digital images (top right image) of bentonite and kaolinite modified with tClVS: kaolinite–tClVS (<b>A</b>) and bentonite–tClVS (<b>B</b>).</p>
Full article ">Figure 3
<p>IR-ATR-FEDS spectra of PCNCs: Kao-NaPSS-10 (<b>A</b>), Kao-NaPSS-20 (<b>B</b>), Bent-NaPSS-10 (<b>C</b>) and Bent-NaPSS-20 (<b>D</b>). Dashed line IR-ATR spectra and solid line FEDS spectra. (being Kao = kaolinite, Bent = bentonite, and NaPSS = sodium poly(styrene sulfonate), where clay percentages are given by 10 and 20 in the end of each notation).</p>
Full article ">Figure 4
<p>SEM images (left image), EDS spectrum for the determination of elemental composition (bottom center image, EDS analysis was carried out in mapping mode of the corresponding SEM image), results of EDS mapping (upper center images, each color represents an element, C (carbon), S (sulfur) and Si (silicon)) and digital images (right image) of PCNCs: Kao-NaPSS-10 (<b>A</b>), Kao-NaPSS-20 (<b>B</b>), Bent-NaPSS-10 (<b>C</b>) and Bent-NaPSS-20 (<b>D</b>).</p>
Full article ">Figure 5
<p>TGA results (solid line) and derivative thermogravimetric curve (dashed line, mass/temperature derivation) of PCNCs: Kao-NaPSS-10 (<b>A</b>), Kao-NaPSS-20 (<b>B</b>), Bent-NaPSS-10 (<b>C</b>) and Bent-NaPSS-20 (<b>D</b>).</p>
Full article ">Figure 6
<p>DSC results of PCNCs: Kao-NaPSS-10 (<b>A</b>), Kao-NaPSS-20 (<b>B</b>), Bent-NaPSS-10 (<b>C</b>) and Bent-NaPSS-20 (<b>D</b>). Analysis conditions: inert nitrogen atmosphere in a temperature range from room temperature to 300 °C and a heating ramp of 15 °C/min.</p>
Full article ">Figure 7
<p>Water sorption capacity (WSC) (<b>A</b>) and cation exchange capacity (CEC) (<b>B</b>) of PCNCs: Kao-NaPSS-10, Kao-NaPSS-20, Bent-NaPSS-10 and Bent-NaPSS-20. The value in the center of the boxes and the error bars corresponds to the mean value and standard deviation of the measurements, respectively.</p>
Full article ">
21 pages, 9002 KiB  
Review
Organoclays Based on Bentonite and Various Types of Surfactants as Heavy Metal Remediants
by Leonid Perelomov, Maria Gertsen, Marina Burachevskaya, S. Hemalatha, Architha Vijayalakshmi, Irina Perelomova and Yurii Atroshchenko
Sustainability 2024, 16(11), 4804; https://doi.org/10.3390/su16114804 - 5 Jun 2024
Cited by 3 | Viewed by 1795
Abstract
The rapid industrial development of civilization has led to the need for the development of new materials to clean up chemically contaminated wastewater and soils. Organoclays, based on smectite minerals and various types of surfactants, are one of the most effective sorbents for [...] Read more.
The rapid industrial development of civilization has led to the need for the development of new materials to clean up chemically contaminated wastewater and soils. Organoclays, based on smectite minerals and various types of surfactants, are one of the most effective sorbents for adsorbing organic and inorganic pollutants. Organoclays are clay minerals that have been modified by the intercalation or grafting of organic molecules. The main mechanism of interaction between organic substances and organoclays involves the adsorption of the substances onto the surface of the clay mineral, which has an expanded structural cell. Various types of surfactants can be used to synthesize organoclays, including cationic, anionic, and amphoteric surfactants. Each type of surfactant has different properties that affect the clay’s ability to sorb. Cationic forms of trace elements, such as heavy metals, can also be adsorbed by organoclays. Data on the adsorption of these substances by organoclays are provided, along with information on how to synthesize them using various surfactants. This review also discusses the main mechanisms of interaction between these substances and clays and the various methods used to create organoclays. It is clear that the adsorption of heavy metals by organoclays is not influenced by their structure or properties, as they belong to the category of surfactant, but rather by their overall chemical structure and characteristics. The wide variety of surfactant types leads to different effects on the adsorption properties of trace elements. Full article
Show Figures

Figure 1

Figure 1
<p>Structure and classification of surfactants [<a href="#B50-sustainability-16-04804" class="html-bibr">50</a>].</p>
Full article ">Figure 2
<p>Schematic representation of Gemini surfactant.</p>
Full article ">
126 KiB  
Abstract
Removal of Uranium (VI) from the Water Environment Using Mechanochemical-Activated Organoclay
by Iryna Kovalchuk
Proceedings 2024, 102(1), 19; https://doi.org/10.3390/proceedings2024102019 - 3 Apr 2024
Viewed by 295
Abstract
The contamination of the environment in the uranium-mining region of Ukraine occurs as the result of the technological processes of mining and processing of uranium raw materials [...] Full article
13 pages, 5677 KiB  
Article
Preparation of Polyethylene Clay Composites via Melt Intercalation Using Hydrophobic and Superhydrophobic Organoclays and Comparison of Their Textural, Mechanical and Thermal Properties
by Ahmet Gürses and Kübra Güneş
Polymers 2024, 16(2), 272; https://doi.org/10.3390/polym16020272 - 19 Jan 2024
Cited by 4 | Viewed by 1384
Abstract
Polymer clay nanocomposites, which can exhibit many superior properties compared to virgin polymers, have gained increasing interest and importance in recent years. This study aimed to prepare composites of two organoclays with unusual ratios and different degrees of lyophilicity with low-density polyethylene and [...] Read more.
Polymer clay nanocomposites, which can exhibit many superior properties compared to virgin polymers, have gained increasing interest and importance in recent years. This study aimed to prepare composites of two organoclays with unusual ratios and different degrees of lyophilicity with low-density polyethylene and compare their textural structures and thermal and mechanical properties with those of virgin polymer. For this purpose, firstly, organoclays, hydrophobic and superhydrophobic organoclays (OC and SOC), were prepared by solution intercalation method using cetyltrimethylammonium bromide with and without addition of a hydrocarbon substance. Then, using both organoclays, polyethylene organoclay composites were prepared and characterized using X-ray powder diffraction (XRD), high-resolution transmission electron microscopy (HRTEM) and Fourier transform infrared spectroscopy (FTIR) techniques. Additionally, tensile and hardness tests were performed to determine the mechanical properties of the composites, and differential scanning calorimetry (DSC) thermograms were taken to examine their thermal behavior. XRD patterns and HRTEM images of hydrophobic and superhydrophobic organoclays and the composites show that the characteristic smectite peak of the clay shifts to the left and expands, that is, the interlayer space widens and, in the composites, it deforms immediately at low clay ratios. HRTEM images of the composites prepared especially with low clay ratios indicate that a heterogeneous dispersion of clay platelets occurs, indicating that nanocomposite formation has been achieved. On the contrary, in the composites prepared with high clay ratios, this dispersion behavior partially turns into aggregation. In the composites prepared using up to 20% by weight of superhydrophobic organoclay, extremely stable and continuous improvements in all mechanical properties were observed compared to those of the composites prepared using hydrophobic organoclay. This indicates that by using superhydrophobic organoclay, a ductile nanocomposite of polyethylene containing inorganic components in much higher than usual proportions can be prepared. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of hydrophobic organoclay and superhydrophobic organoclay preparation procedures.</p>
Full article ">Figure 2
<p>HRTEM images of (<b>a</b>) raw clay (RC), (<b>b</b>) hydrophobic clay (OC), (<b>c</b>) superhydrophobic clay (SOC), (<b>d</b>) virgin polymer (LDPE), (<b>e</b>) the composite prepared using hydrophobic organoclay (LDPEOCC1) and (<b>f</b>) the composite prepared using superhydrophobic organoclay (LDPESOCC1).</p>
Full article ">Figure 2 Cont.
<p>HRTEM images of (<b>a</b>) raw clay (RC), (<b>b</b>) hydrophobic clay (OC), (<b>c</b>) superhydrophobic clay (SOC), (<b>d</b>) virgin polymer (LDPE), (<b>e</b>) the composite prepared using hydrophobic organoclay (LDPEOCC1) and (<b>f</b>) the composite prepared using superhydrophobic organoclay (LDPESOCC1).</p>
Full article ">Figure 3
<p>FTIR spectra of the virgin polymer (LDPE) and the composites prepared using both hydrophobic organoclay and superhydrophobic organoclay (LDPEOCC1–5 (<b>a</b>) and LDPESOCC1–5 (<b>b</b>)).</p>
Full article ">Figure 3 Cont.
<p>FTIR spectra of the virgin polymer (LDPE) and the composites prepared using both hydrophobic organoclay and superhydrophobic organoclay (LDPEOCC1–5 (<b>a</b>) and LDPESOCC1–5 (<b>b</b>)).</p>
Full article ">Figure 4
<p>XRD patterns of raw clay (RC) and the composites prepared using both hydrophobic organoclay and superhydrophobic organoclay: (<b>a</b>) LDPEOCC1–5 and (<b>b</b>) LDPESOCC1–5.</p>
Full article ">Figure 5
<p>DSC thermograms of the virgin polymer (LDPE) and the composites prepared using both hydrophobic organoclay and superhydrophobic organoclay: (<b>a</b>) LDPEOCC1–5 and (<b>b</b>) LDPESOCC1–5.</p>
Full article ">
24 pages, 8933 KiB  
Article
A Commercial Clay-Based Material as a Carrier for Targeted Lysozyme Delivery in Animal Feed
by Marianna Guagliano, Cinzia Cristiani, Matteo Dell’Anno, Giovanni Dotelli, Elisabetta Finocchio, Maria Lacalamita, Ernesto Mesto, Serena Reggi, Luciana Rossi and Emanuela Schingaro
Nanomaterials 2023, 13(22), 2965; https://doi.org/10.3390/nano13222965 - 17 Nov 2023
Viewed by 1284
Abstract
The controlled supply of bioactive molecules is a subject of debate in animal nutrition. The release of bioactive molecules in the target organ, in this case the intestine, results in improved feed, as well as having a lower environmental impact. However, the degradation [...] Read more.
The controlled supply of bioactive molecules is a subject of debate in animal nutrition. The release of bioactive molecules in the target organ, in this case the intestine, results in improved feed, as well as having a lower environmental impact. However, the degradation of bioactive molecules’ in transit in the gastrointestinal passage is still an unresolved issue. This paper discusses the feasibility of a simple and cost-effective procedure to bypass the degradation problem. A solid/liquid adsorption procedure was applied, and the operating parameters (pH, reaction time, and LY initial concentration) were studied. Lysozyme is used in this work as a representative bioactive molecule, while Adsorbo®, a commercial mixture of clay minerals and zeolites which meets current feed regulations, is used as the carrier. A maximum LY loading of 32 mgLY/gAD (LY(32)-AD) was obtained, with fixing pH in the range 7.5–8, initial LY content at 37.5 mgLY/gAD, and reaction time at 30 min. A full characterisation of the hybrid organoclay highlighted that LY molecules were homogeneously spread on the carrier’s surface, where the LY–carrier interaction was mainly due to charge interaction. Preliminary release tests performed on the LY(32)-AD synthesised sample showed a higher releasing capacity, raising the pH from 3 to 7. In addition, a preliminary Trolox equivalent antioxidant capacity (TEAC) assay showed an antioxidant capacity for the LY of 1.47 ± 0.18 µmol TroloxEq/g with an inhibition percentage of 33.20 ± 3.94%. Full article
(This article belongs to the Section Nanocomposite Materials)
Show Figures

Figure 1

Figure 1
<p>Lysozyme–Adsorbo<sup>®</sup> adsorption process.</p>
Full article ">Figure 2
<p>Methodology workflow for the preliminary lysozyme release assay.</p>
Full article ">Figure 3
<p>Steps in the preliminary antioxidant activity assay of lysozyme.</p>
Full article ">Figure 4
<p>LY capture: (<b>a</b>) as a function of reaction time, and (<b>b</b>) as a function of the initial LY content.</p>
Full article ">Figure 5
<p>FT-IR skeletal spectra of (<b>a</b>) LY(6)-AD; (<b>b</b>) LY(22)-AD; (<b>c</b>) LY(32)-AD, upon subtraction of spectrum of pristine AD. Dashed line: FT IR skeletal spectrum of pure LY in KBr.</p>
Full article ">Figure 6
<p>TGA (<b>a</b>) and DTG (<b>b</b>) curves of LY-AD samples at increasing LY loading (pristine AD and LY are reported for comparison, dashed lines: LY decomposition maxima). Curves were intentionally shifted to highlight differences; plotting in the original version in <a href="#app1-nanomaterials-13-02965" class="html-app">Figure S8</a> (Colours in (a) as in (b)).</p>
Full article ">Figure 7
<p>Comparison of XRPD patterns of (<b>a</b>) LY(32)-AD and LY-AD-MM, AD reported for comparison, and (<b>b</b>) enlarged pattern of LY-AD-MM. The position of the most intense (I<sub>100</sub>) reflections of the single phases is reported (Z: zeolite; Chab: chabazite; Phil: phillipsite; Feld: feldspar; D: dolomite).</p>
Full article ">Figure 8
<p>Comparison of the synthetised LY(32)-AD organoclay and the mechanical mixture LY-AD-MM: (<b>a</b>) FT-IR spectra upon AD subtraction, and (<b>b</b>) DTG. In both plots: LY-AD-MM in blue, LY(32)-AD in red. Dashed line in plot (<b>a</b>) pure LY spectrum.</p>
Full article ">Figure 9
<p>SEM images (magnification = 2 µm) of (<b>a</b>) pristine AD, (<b>b</b>) pristine LY, (<b>c</b>) LY(32)-AD the synthesised organoclay, and (<b>d</b>) LY-AD-MM the mechanical mixture. Yellow circles: LY aggregates.</p>
Full article ">Figure 10
<p>SEM-EDX analysis of LY (32)-AD organoclay (<b>a</b>) SEM of the analysed portion, (<b>b</b>) EXD of the analysed portion, (<b>c</b>) EXD of Si, (<b>d</b>) EDX of Al, (<b>e</b>) EDX of Mg, and (<b>f</b>) EDX of S.</p>
Full article ">Figure 11
<p>(<b>a</b>) SEM analysis of the synthesised organoclay, LY(32)-AD, (<b>b</b>) EDX analysis of the synthesised organoclay, LY(32)-AD, (<b>c</b>) SEM analysis of the mechanical mixture, LY-AD-MM, (<b>d</b>) EDX analysis of the mechanical mixture, LY-AD-MM. Yellow circles: lysozyme aggregates.</p>
Full article ">
Back to TopTop