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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (262)

Search Parameters:
Keywords = polystyrene carbonate

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 4519 KiB  
Article
Impact of Polystyrene Microplastics on Soil Properties, Microbial Diversity and Solanum lycopersicum L. Growth in Meadow Soils
by Shuming Liu, Yan Suo, Jinghuizi Wang, Binglin Chen, Kaili Wang, Xiaoyu Yang, Yaokun Zhu, Jiaxing Zhang, Mengchu Lu and Yunqing Liu
Plants 2025, 14(2), 256; https://doi.org/10.3390/plants14020256 - 17 Jan 2025
Viewed by 561
Abstract
The pervasive presence of microplastics (MPs) in agroecosystems poses a significant threat to soil health and plant growth. This study investigates the effects of varying concentrations and sizes of polystyrene microplastics (PS-MPs) on the Solanum lycopersicum L.’s height, dry weight, antioxidant enzyme activities, [...] Read more.
The pervasive presence of microplastics (MPs) in agroecosystems poses a significant threat to soil health and plant growth. This study investigates the effects of varying concentrations and sizes of polystyrene microplastics (PS-MPs) on the Solanum lycopersicum L.’s height, dry weight, antioxidant enzyme activities, soil physicochemical properties, and rhizosphere microbial communities. The results showed that the PS0510 treatment significantly increased plant height (93.70 cm, +40.83%) and dry weight (2.98 g, +100%). Additionally, antioxidant enzyme activities improved across treatments for S. lycopersicum L. roots. Physicochemical analyses revealed enhanced soil organic matter and nutrient levels, including ammonium nitrogen, phosphorus, and effective potassium. Using 16S rRNA sequencing and molecular ecological network techniques, we found that PS-MPs altered the structure and function of the rhizosphere microbial community associated with S. lycopersicum L. The PS1005 treatment notably increased microbial diversity and displayed the most complex ecological network, while PS1010 led to reduced network complexity and more negative interactions. Linear discriminant analysis effect size (LEfSe) analysis identified biomarkers at various taxonomic levels, reflecting the impact of PS-MPs on microbial community structure. Mantel tests indicated positive correlations between microbial diversity and soil antioxidant enzyme activity, as well as relationships between soil physicochemical properties and enzyme activity. Predictions of gene function revealed that PS-MP treatments modified carbon and nitrogen cycling pathways, with PS1005 enhancing methanogenesis genes (mcrABG) and PS1010 negatively affecting denitrification genes (nirK, nirS). This study provides evidence of the complex effects of PS-MPs on soil health and agroecosystem functioning, highlighting their potential to alter soil properties and microbial communities, thereby affecting plant growth. Full article
(This article belongs to the Section Plant–Soil Interactions)
Show Figures

Figure 1

Figure 1
<p>Impact of PS-MPs on growth and antioxidant enzyme activities in <span class="html-italic">S. lycopersicum</span> L.: (<b>a</b>) plant height; (<b>b</b>) dry weight; (<b>c</b>) SOD; (<b>d</b>) POD; (<b>e</b>) CAT. Different lowercase letters indicate significant differences, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Physicochemical of soil: (<b>a</b>) organic matter; (<b>b</b>) NH<sub>4</sub>-N concentration; (<b>c</b>) NO<sub>3</sub>-N concentration; (<b>d</b>) P concentration; (<b>e</b>) available P; (<b>f</b>) available K; (<b>g</b>) pH value. Different lowercase letters indicate significant differences, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Biochemical of soil (<b>a</b>) S-SOD; (<b>b</b>) S-POD; (<b>c</b>) S-CAT. Different lowercase letters indicate significant differences, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Diversity and composition of rhizosphere soil bacterial community in <span class="html-italic">S. lycopersicum</span> L: (<b>a</b>) Simpson Index; (<b>b</b>) Shannon Index; (<b>c</b>) Venn diagram; (<b>d</b>) PCoA; (<b>e</b>) the abundance of microorganisms at the genus level in a Circos diagram. “*” represent significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Linear discriminant analysis effect size (LEfSe): (<b>a</b>) diagram depicting the evolutionary branching; (<b>b</b>) histogram displaying the distribution.</p>
Full article ">Figure 6
<p>Co-occurrence Networks analysis: (<b>a</b>) CK; (<b>b</b>) PS0505; (<b>c</b>) PS0510; (<b>d</b>) PS1005; (<b>e</b>) PS1010.</p>
Full article ">Figure 7
<p>Correlation analysis of bacterial communities, plant growth, enzyme activities and the physicochemical and biochemical properties of soil: (<b>a</b>) Mantel’s test investigated the correlation between the genus level bacteria and the indices of Shannon, Simpson, Chao1, and ACE; (<b>b</b>) VPA analysis; (<b>c</b>) Pearson correlations of genus, soil physicochemical, and biochemical properties. “*”, “**” and “***” represent significant differences (<span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 8
<p>PLS-PM showing the direct and indirect effects of adding PS-MPs to the soil–microbe–plant system. The width of arrows is proportional to the strength of the pathway coefficients. Red and blue arrows indicate positive and negative causality, respectively, with solid lines representing significant effects (<span class="html-italic">p</span> &lt; 0.05), and dashed lines representing non-significant effects (<span class="html-italic">p</span> &gt; 0.05). “*” indicates <span class="html-italic">p</span> &lt; 0.05; “***” indicates <span class="html-italic">p</span> &lt; 0.001. (<b>a</b>) PLS-PM; (<b>b</b>) Effect of direct, indirect and total. Soil physicochemical effects included NH<sub>4</sub>-N, P, AK, SOM, and pH; soil enzymes included S-SOD, S-POD, and S-CAT; microbial effects included the Simpson Index and Shannon Index; plant effects included SOD, POD, CAT, Height, and Dry weight.</p>
Full article ">Figure 9
<p>Relative abundance of the pathways involved in the C and N cycle. (<b>a</b>) C cycle; (<b>b</b>) N cycle. The pie chart illustrates the proportional representation of various pathways within each metagenomic sample. The dimensions of the pie charts reflect the overall relative abundance of each pathway.</p>
Full article ">
31 pages, 4220 KiB  
Article
Improving the Long-Term Mechanical Properties of Thermoplastic Short Natural Fiber Compounds by Using Alternative Matrices
by Renato Lemos Cosse, Tobias van der Most, Vincent S. D. Voet, Rudy Folkersma and Katja Loos
Biomimetics 2025, 10(1), 46; https://doi.org/10.3390/biomimetics10010046 - 13 Jan 2025
Viewed by 747
Abstract
Wood plastic composites (WPCs) offer a means to reduce the carbon footprint by incorporating natural fibers to enhance the mechanical properties. However, there is limited information on the mechanical properties of these materials under hostile conditions. This study evaluated composites of polypropylene (PP), [...] Read more.
Wood plastic composites (WPCs) offer a means to reduce the carbon footprint by incorporating natural fibers to enhance the mechanical properties. However, there is limited information on the mechanical properties of these materials under hostile conditions. This study evaluated composites of polypropylene (PP), polystyrene (PS), and polylactic acid (PLA) processed via extrusion and injection molding. Tests were conducted on tensile and flexural strength and modulus, heat deflection temperature (HDT), and creep analysis under varying relative humidity conditions (10% and 90%) and water immersion, followed by freeze—thaw cycles. The addition of fibers generally improved the mechanical properties but increased water absorption. HDT and creep were dependent on the crystallinity of the composites. PLA and PS demonstrated a superior overall performance, except for their impact properties, where PP was slightly better than PLA. Full article
(This article belongs to the Special Issue Advances in Biomaterials, Biocomposites and Biopolymers 2024)
Show Figures

Figure 1

Figure 1
<p>Differences in preparation between (<b>A</b>) PLA1+F35, (<b>B</b>) PLA1+MA+DCP+F35, and (<b>C</b>) PLA1+G-MA+F35. MA and DCP are amplified 10×.</p>
Full article ">Figure 2
<p>Macro-scale wetting of wood fibers by the polymer matrix. (<b>A</b>) PLA1+F17.5; (<b>B</b>) PLA1+F35; (<b>C</b>) PLA1+C35; (<b>D</b>) PLA2+F35; (<b>E</b>) PLA1+MA+DCP+F35; (<b>F</b>) PLA1+G-MA+F35.</p>
Full article ">Figure 3
<p>Side-by-side images of all the injection-molded samples.</p>
Full article ">Figure 4
<p>Water absorption behavior of submerged tensile bars fabricated from neat polymers and compatibilized polymer composites.</p>
Full article ">Figure 5
<p>Water absorption of submersed PLA1, PLA2, and their respective composite tensile bars for 28 days. The composites that underwent the test were PLA2+F35, PLA1+G-MA+F35, PLA1+MA+DCP+F35, PLA1+F35, PLA1+C35, and PLA1+F35.</p>
Full article ">Figure 6
<p>Tensile modulus of pure polymers and compatibilized compounds under various moisture.</p>
Full article ">Figure 7
<p>Tensile strength of pure polymers and compatibilized compounds under various moisture conditions.</p>
Full article ">Figure 8
<p>Tensile modulus of pure PLA and PLA compounds under various moisture conditions.</p>
Full article ">Figure 9
<p>Tensile strength of pure PLA and PLA compounds under various moisture conditions.</p>
Full article ">Figure 10
<p>Flexural creep of pure polymers and their compatibilized compounds.</p>
Full article ">Figure 11
<p>Flexural creep of PLA and PLA-based composites.</p>
Full article ">Figure 12
<p>Flexural creep of crystallized PLA and PLA compounds compared to that of the PP composite.</p>
Full article ">
14 pages, 3242 KiB  
Article
Effect of Sodium Hypochlorite Disinfection on Polyvinylidene Fluoride Membranes in Microplastic Ultrafiltration
by Guanghua Wang, Tongyu Li, Wenxuan Yin, Jianhua Zhou and Dongwei Lu
Water 2025, 17(1), 99; https://doi.org/10.3390/w17010099 - 2 Jan 2025
Viewed by 853
Abstract
With the widespread use of plastic products, microplastic (MP) pollution has become an important factor threatening the water environment and human health. Ultrafiltration (UF) technology, based on organic polymer membranes, is a common method to remove MPs in water treatment processes, offering high [...] Read more.
With the widespread use of plastic products, microplastic (MP) pollution has become an important factor threatening the water environment and human health. Ultrafiltration (UF) technology, based on organic polymer membranes, is a common method to remove MPs in water treatment processes, offering high removal efficiency and scalability. However, in water treatment plants (WTPs), oxidation pretreatment is often applied before UF, and the presence of oxidants can affect membrane performance. In this study, we constructed a polyvinylidene fluoride (PVDF) ultrafiltration membrane for a gravity filtration system to investigate the impact of sodium hypochlorite oxidation pretreatment on the removal of polystyrene (PS) MPs under gravity filtration. As a result, pre-chlorination reduced PS microplastic deposition on membranes by improving flux stability (15.1%) but significantly decreased the removal rate (from 36.6% to 22.6%). Pre-oxidation facilitated a shift in fouling behavior toward intermediate blocking while reducing standard blocking and enhancing irreversible fouling recovery. However, continuous chlorine exposure increased membrane porosity and pore size, substituted fluorine with chlorine, and led to organic carbon leaching, indicating pre-oxidation jeopardizes membrane stability and separation performance. These findings provide insights into the development of novel strategies aimed at enhancing the efficiency and sustainability of membrane treatment processes in WTPs. Full article
(This article belongs to the Section Wastewater Treatment and Reuse)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of dead-end filtration unit.</p>
Full article ">Figure 2
<p>Filtration performance of PVDF membrane for PS-MPs: (<b>a</b>) WCAs of membrane surface after filtration, (<b>b</b>) pure water flux, (<b>c</b>) normalized flux, and (<b>d</b>) removal efficiency. <span class="html-italic">J</span><sub>0</sub> and <span class="html-italic">J</span><sub>1</sub> represent the initial and final flux of the PS-MP solution, respectively.</p>
Full article ">Figure 3
<p>The surface morphology of the membrane after filtration of PS-MPs at different available chlorine concentrations: (<b>a</b>) pristine membrane, (<b>b</b>) 0 mg/L, (<b>c</b>) 1 mg/L, (<b>d</b>) 2 mg/L, and (<b>e</b>) 3 mg/L.</p>
Full article ">Figure 4
<p>Membrane fouling in the prechlorination–ultrafiltration system: (<b>a</b>) IFR, FRR, and PR, (<b>b</b>) coefficient of determination (R<sup>2</sup>), and (<b>c</b>) membrane blocking coefficient of Hermia’s models.</p>
Full article ">Figure 5
<p>The surface morphology and pore size distribution of the membrane after chlorination at different concentrations: (<b>a</b>) 0, (<b>b</b>) 300 mg/L, (<b>c</b>) 600 mg/L, (<b>d</b>) 1000 mg/L, (<b>e</b>) 1500 mg/L, and (<b>f</b>) porosity.</p>
Full article ">Figure 6
<p>Properties of the chlorinated membranes: (<b>a</b>,<b>b</b>) FTIR spectra, (<b>c</b>) the decrease rate of peak intensity at 1179 and 1404 cm<sup>−1</sup>, and (<b>d</b>) total organic carbon content in leaching solutions.</p>
Full article ">
17 pages, 18470 KiB  
Article
Photonic Band Gap Engineering by Varying the Inverse Opal Wall Thickness
by Dániel Attila Karajz, Levente Halápi, Tomasz Stefaniuk, Bence Parditka, Zoltán Erdélyi, Klára Hernádi, Csaba Cserháti and Imre Miklós Szilágyi
Int. J. Mol. Sci. 2024, 25(23), 12996; https://doi.org/10.3390/ijms252312996 - 3 Dec 2024
Viewed by 738
Abstract
We demonstrate the band gap programming of inverse opals by fabrication of different wall thickness by atomic layer deposition (ALD). The opal templates were synthesized using polystyrene and carbon nanospheres by the vertical deposition method. The structure and properties of the TiO2 [...] Read more.
We demonstrate the band gap programming of inverse opals by fabrication of different wall thickness by atomic layer deposition (ALD). The opal templates were synthesized using polystyrene and carbon nanospheres by the vertical deposition method. The structure and properties of the TiO2 inverse opal samples were investigated using Scanning Electron Microscope (SEM) and Focused Ion Beam Scanning Electron Microscopy (FIB-SEM), Energy Dispersive X-ray analysis (EDX), X-ray Diffraction (XRD) and Finite Difference Time Domain (FDTD) simulations. The photonic properties can be well detected by UV-Vis reflectance spectroscopy, while diffuse reflectance spectroscopy appears to be less sensitive. The samples showed visible light photocatalytic properties using Raman microscopy and UV-Visible spectrophotometry, and a newly developed digital photography-based detection method to track dye degradation. In our work, we stretch the boundaries of a working inverse opal to make it commercially more available while avoiding fully filling and using cheaper, but lower-quality, carbon nanosphere sacrificial templates. Full article
(This article belongs to the Special Issue Fabrication and Application of Photocatalytically Active Materials)
Show Figures

Figure 1

Figure 1
<p>SEM images (<b>left</b>) and FIB-SEM images (<b>right</b>) of TiO<sub>2</sub> inverse opal samples. CSIO–inverse opal made using carbon nanospheres opal as template PSIO–inverse opal made using polystyrene nanospheres opal as template 1, 2 and 3–16.5 nm, 34.3 nm and 49.7 nm deposited TiO<sub>2</sub> layer by ALD.</p>
Full article ">Figure 2
<p>XRD diffractograms of the TiO<sub>2</sub> inverse opal samples.</p>
Full article ">Figure 3
<p>UV-Vis absorbance spectra of the TiO<sub>2</sub> inverse opal samples in reflectance and diffuse reflectance modes.</p>
Full article ">Figure 4
<p>(<b>Left</b>): The refractive index and extinction coefficient of thin TiO<sub>2</sub> layer, deposited using ALD process [<a href="#B44-ijms-25-12996" class="html-bibr">44</a>]. The inset presents the schematic of the opal structure, used in numerical simulations. (<b>Right</b>): Absorbance of PSIO samples calculated using FDTD method. The thicknesses of TiO<sub>2</sub> layers, in particular, inverse opal geometries, correspond to the ones measured in the experiment: 16 nm (blue curve), 34 nm (red curve), 49 nm (orange curve). The shaded parts of the graph correspond to three different types of the optical response of the structure (marked for PSIO-1 sample).</p>
Full article ">Figure 5
<p>The electric field distributions calculated for a single cell of PSIO−1 sample for <span class="html-italic">λ</span> = 275 nm (<b>left</b>), <span class="html-italic">λ</span> = 428 nm (<b>middle</b>) and <span class="html-italic">λ</span> = 750 nm (<b>right</b>) wavelength. The colormaps are normalized to the amplitude of the illuminating plane-wave. The white line indicates the plane tangent to the first layer of nanospheres.</p>
Full article ">Figure 6
<p>Absorbance of CSIO-1 sample calculated using FDTD method. Different curves correspond to samples with different degrees of disorder: (blue) no disarrangement, (red) Δ = 20 nm, (yellow) Δ = 40 nm.</p>
Full article ">Figure 7
<p>Raman spectra of the methylene blue dried on the surface of the samples before and after the 19-h irradiation with the highlight of the most intensive methylene blue peak.</p>
Full article ">Figure 8
<p>Images taken by the Sony A6000 (<b>top</b>) and Nikon D3400 (<b>bottom</b>) cameras during the photocatalysis experiment.</p>
Full article ">Figure 9
<p>Summary of the photocatalysis tests using Sony A6000 (<b>left</b>) and Nikon D3400 (<b>right</b>) cameras. The spectrums show how the absorbance of red, green and blue color of the pixels changes over the 4 h long experiment.</p>
Full article ">Figure 10
<p>Relative absorbance of the red and green colours of the pixels as a function of time during the photocatalytic test for every sample.</p>
Full article ">Figure 11
<p>Visible light photocatalysis of rhodamine 6 G by the reference and photocatalysts.</p>
Full article ">Figure 12
<p>Representation of the interstitial sites and the equations used in the calculation of their size.</p>
Full article ">
16 pages, 3797 KiB  
Article
Novel Synthesis of Polystyrenesulfonate@AC Based on Olive Tree Leaves Biomass for the Photo-Degradation of Methylene Blue from Aqueous Solution
by Ibrahim Hotan Alsohaimi
Polymers 2024, 16(23), 3321; https://doi.org/10.3390/polym16233321 - 27 Nov 2024
Viewed by 600
Abstract
Water pollution poses significant environmental challenges, particularly from dyes used in various industrial processes. Effective removal methods are essential to mitigate their impact on aquatic environments. Activated carbon (AC) is widely used for its adsorption properties, and further modifications can enhance its efficiency. [...] Read more.
Water pollution poses significant environmental challenges, particularly from dyes used in various industrial processes. Effective removal methods are essential to mitigate their impact on aquatic environments. Activated carbon (AC) is widely used for its adsorption properties, and further modifications can enhance its efficiency. In this study, we developed polystyrene sulfonate-modified activated carbon (AC@PSS) using a facile and efficient method to improve the photo-degradation of methylene blue (MB) in aquatic environments. The modification enhanced the activated carbon’s surface features and adsorption, improving its photocatalytic activity. The photocatalysts were characterized using XRD, SEM, FTIR, and TGA. Based on Tauc’s equation, the band gap value of AC@PSS was 4.0 eV. The photocatalytic efficacy of the AC@PSS catalyst was assessed by studying the degradation of MB dye under UV-rich solar irradiation. The influence of various variables on the photo-degradation of MB dye such as pH (2–12), reaction time (0–160 min), catalyst dosage (20–80 mg), and dye concentration (10–300 mg/L) was investigated. The AC@PSS catalyst demonstrated impressive degradation efficacy for MB dye of 98% in 160 min at pH 11, a temperature of 25 °C, a catalyst dose of 60 mg, and initial MB content of 10 mg/L. The superior performance of the AC@PSS catalyst could be due to the effective separation of photogenerated electron holes. Accordingly, the photo-degradation of MB is affected by the photo-produced radical OH. Finally, we conclude that synthesizing AC@PSS is highly effective for the degradation of MB dye. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the synthesis of the AC@PSS composite.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) SEM images of AC and (<b>c</b>,<b>d</b>) AC@PSS composites.</p>
Full article ">Figure 3
<p>N<sub>2</sub> isotherms of AC and AC@PSS composites.</p>
Full article ">Figure 4
<p>FTIR spectrum of AC and AC@PSS composites.</p>
Full article ">Figure 5
<p>(<b>a</b>) XRD patterns of AC and AC@PSS composites (the inset is the XRD pattern of PSS) and (<b>b</b>) TGA profiles of AC and AC@PSS composites.</p>
Full article ">Figure 6
<p>(<b>a</b>) UV–vis absorption spectra and (<b>b</b>) Tauc’s profile for the band gap assessment of the AC@PSS composite.</p>
Full article ">Figure 7
<p>Photocatalytic efficacy (C<sub>t</sub>/C<sub>o</sub> against time) of MB irradiated by UV-rich solar light (without AC@PSS), MB adsorption without irradiation, and MB irradiated by UV-rich solar light in the presence of AC@PSS.</p>
Full article ">Figure 8
<p>(<b>a</b>) Effect of pH and (<b>b</b>) catalyst dosage on the photo-degradation efficiency of MB dye using the AC@PSS composite.</p>
Full article ">Figure 9
<p>(<b>a</b>) The effect of time on the photo-degradation efficiency of MB dye, (<b>b</b>) the influence of initial MB concentration on photo-degradation efficiency, and (<b>c</b>) photocatalytic degradation kinetics of MB using the AC@PSS composite.</p>
Full article ">Figure 10
<p>Photocatalytic degradation of MB by the AC@PSS composite in the presence of EDTA as a (h<sup>+</sup>) scavenger, silver nitrate as an (e<sup>–</sup>) scavenger, tert-butyl alcohol as an (<sup>•</sup>OH) scavenger, and methyl alcohol as an (<sup>•</sup>O<sub>2</sub>) scavenger.</p>
Full article ">Figure 11
<p>A possible mechanism of the photocatalytic degradation of the MB dyes onto the AC@PSS composite.</p>
Full article ">
21 pages, 3951 KiB  
Article
Smart Coating of Carbon Steel Using Polystyrene Clay Nanocomposites Loaded with Cerium and Silanol Inhibitors: Characterization and Electrochemical Study
by Layla A. Al Juhaiman, Mona A. Al Jufareen, Saeed M. Al-Zahrani, Ubair Abdus Samad and Tahani S. Al-Garni
Polymers 2024, 16(22), 3196; https://doi.org/10.3390/polym16223196 - 17 Nov 2024
Viewed by 943
Abstract
Local Khulays clay was modified to prepare polystyrene clay nanocomposite (PCN) coatings on carbon steel. The PCN coatings were added to microcapsules (MCs) loaded with the corrosion inhibitor PCN(MC). The microcapsules were prepared by the encapsulation of rare-earth metal Ce+3 ions and [...] Read more.
Local Khulays clay was modified to prepare polystyrene clay nanocomposite (PCN) coatings on carbon steel. The PCN coatings were added to microcapsules (MCs) loaded with the corrosion inhibitor PCN(MC). The microcapsules were prepared by the encapsulation of rare-earth metal Ce+3 ions and isobutyl silanol into polystyrene via the double emulsion solvent evaporation (DESE) technique. From characterization techniques, Fourier-transform infrared spectroscopy (FT-IR), X-ray diffraction (XRD), transmission electron microscopy (TEM), and scanning electron microscopy (SEM) with EDX. SEM and FT-IR confirmed the success of the preparation of the PCN(MC). Nanoindentation tests were performed on the thin-film samples. A significant reduction in both the hardness and the reduced modulus was observed for the PCN film compared to the PS film. Electrochemical impedance spectroscopy (EIS) and electrochemical frequency modulation (EFM) all showed an enhanced protection efficiency (%PE) of 3% PCN(MC) over 3% PCN at high temperatures and at different times. The smart coatings were proven by applying the thermal and the mechanical triggers for the 3% PCN(MC) coating. The mechanism of the release of inhibitors was discussed. The self-healing properties of 3% PCN(MC) were evaluated. The enhanced properties of the developed PCN(MC) coatings make them attractive for potential applications in the oil and other industries. Full article
Show Figures

Figure 1

Figure 1
<p>The XRD patterns of RC (<b>A</b>); RC, NaC, and OC (<b>B</b>); and OC, PS, and 3% PCN (<b>C</b>).</p>
Full article ">Figure 2
<p>TEM images of 3% PCN at two magnifications (<b>A</b>,<b>B</b>).</p>
Full article ">Figure 3
<p>FT-IR spectra of RC, NaC, OC, and CPC (<b>A</b>); FTIR of OC, PS, and 3% PCN (<b>B</b>); and FT-IR of MC, PS, W1 IBTMS, and Ce(NO<sub>3</sub>)<sub>3</sub> (<b>C</b>).</p>
Full article ">Figure 4
<p>SEM images showing the diameter of MCs (top photo) and EDX analysis of MCs (bottom photo).</p>
Full article ">Figure 4 Cont.
<p>SEM images showing the diameter of MCs (top photo) and EDX analysis of MCs (bottom photo).</p>
Full article ">Figure 5
<p>Load vs. depth curves for thin-film samples.</p>
Full article ">Figure 6
<p>The Nyquist plot of 3% PCN (<b>top photo</b>) and 3% PCN(MC) in 3.5% NaCl at different temperatures (<b>bottom photo</b>).</p>
Full article ">Figure 7
<p>Schematic diagram of the coated C-steel-equivalent circuit.</p>
Full article ">
24 pages, 4263 KiB  
Article
Production of Aviation Fuel-Range Hydrocarbons Through Catalytic Co-Pyrolysis of Polystyrene and Southern Pine
by Ayden Kemp, Tawsif Rahman, Hossein Jahromi and Sushil Adhikari
Catalysts 2024, 14(11), 806; https://doi.org/10.3390/catal14110806 - 9 Nov 2024
Viewed by 1954
Abstract
Sustainable aviation fuels (SAFs), produced from waste and renewable sources, are a promising means for reducing net greenhouse gas emissions from air travel while still maintaining the quality of air transportation expected. In this work, the catalytic co-pyrolysis of polystyrene and pine with [...] Read more.
Sustainable aviation fuels (SAFs), produced from waste and renewable sources, are a promising means for reducing net greenhouse gas emissions from air travel while still maintaining the quality of air transportation expected. In this work, the catalytic co-pyrolysis of polystyrene and pine with red mud (bauxite residue) and ZSM-5 catalysts at temperatures of 450 °C, 500 °C, and 550 °C was investigated as a method for producing aromatic hydrocarbons with carbon numbers ranging from 7 to 17 for use as additives to blend with SAF produced through other methods to add the required quantity of aromatic molecules to these blends. The maximum yield of kerosene-range aromatic hydrocarbons was 620 mg per gram of feedstock (62% of feedstock was converted to kerosene-range hydrocarbons) obtained at 550 °C in the presence of ZSM-5. Additionally, it was noted that a positive synergy exists between pine and polystyrene feedstocks during co-pyrolysis that cracks solid and liquid products into gaseous products similarly to that of a catalyst. The co-pyrolysis of pine and polystyrene without a catalyst produced on average 17% or 36.3 mg more kerosene-range hydrocarbons than predicted, with a maximum yield of 266 mg of C7–C17 aromatic hydrocarbons per gram of feedstock (26.6% conversion of initial feedstock) obtained at 550 °C. Full article
Show Figures

Figure 1

Figure 1
<p>Pyrolysis product yields: (<b>a</b>) polystyrene pyrolysis; (<b>b</b>) pine pyrolysis; (<b>c</b>) polystyrene and pine co-pyrolysis; (<b>d</b>) theoretically predicted values for polystyrene and pine co-pyrolysis.</p>
Full article ">Figure 2
<p>Distribution of compounds in pyrolysis oil by carbon number: (<b>a</b>) PS at 500 °C; (<b>b</b>) PS with RM at 500 °C; (<b>c</b>) PS with ZSM-5 at 500 °C; (<b>d</b>) pine at 500 °C; (<b>e</b>) pine with RM at 500 °C; (<b>f</b>) pine with ZSM-5 at 500 °C. Numbers atop bars indicate the total area percentage of all compounds for a given carbon number. A list of compounds identified by GC/MS, those containing oxygen, and those excluded from concentration calculations due to insufficient match quality are provided in <a href="#app1-catalysts-14-00806" class="html-app">Table S7</a>.</p>
Full article ">Figure 3
<p>Distribution of compounds in pyrolysis oil by carbon number: (<b>a</b>) pine with RM at 450 °C; (<b>b</b>) pine with RM at 550 °C; (<b>c</b>) pine with ZSM-5 at 450 °C; (<b>d</b>) pine with ZSM-5 at 550 °C; (<b>e</b>) PS with ZSM-5 at 450 °C; (<b>f</b>) PS with ZSM-5 at 550 °C. Numbers atop bars indicate the total area percentage of all compounds for a given carbon number. A list of compounds identified by GC/MS, those containing oxygen, and those excluded from concentration calculations due to insufficient match quality are provided in <a href="#app1-catalysts-14-00806" class="html-app">Table S7</a>.</p>
Full article ">Figure 4
<p>Distribution of compounds in pyrolysis oil by carbon number: (<b>a</b>) PS/pine blend at 450 °C; (<b>b</b>) PS/pine blend with RM at 450 °C; (<b>c</b>) PS/pine blend with ZSM-5 at 450 °C; (<b>d</b>) PS/pine blend at 550 °C; (<b>e</b>) PS/pine blend with RM at 550 °C; (<b>f</b>) PS/pine blend with ZSM-5 at 550 °C. Numbers atop bars indicate the total area percentage of all compounds for a given carbon number. A list of compounds identified by GC/MS, those containing oxygen, and those excluded from concentration calculations due to insufficient match quality are provided in <a href="#app1-catalysts-14-00806" class="html-app">Table S7</a>.</p>
Full article ">Figure 5
<p>Reaction pathways for the pyrolysis of pine and polystyrene and co-pyrolysis of these feedstocks [<a href="#B18-catalysts-14-00806" class="html-bibr">18</a>,<a href="#B19-catalysts-14-00806" class="html-bibr">19</a>,<a href="#B21-catalysts-14-00806" class="html-bibr">21</a>,<a href="#B22-catalysts-14-00806" class="html-bibr">22</a>,<a href="#B23-catalysts-14-00806" class="html-bibr">23</a>,<a href="#B24-catalysts-14-00806" class="html-bibr">24</a>,<a href="#B26-catalysts-14-00806" class="html-bibr">26</a>,<a href="#B46-catalysts-14-00806" class="html-bibr">46</a>,<a href="#B47-catalysts-14-00806" class="html-bibr">47</a>].</p>
Full article ">Figure 6
<p>Diagram of fixed bed pyrolysis reactor setup.</p>
Full article ">
15 pages, 3317 KiB  
Article
A Label-Free Electrochemical Aptamer Sensor for Sensitive Detection of Cardiac Troponin I Based on AuNPs/PB/PS/GCE
by Liying Jiang, Dongyang Li, Mingxing Su, Yirong Qiu, Fenghua Chen, Xiaomei Qin, Lan Wang, Yanghai Gui, Jianbo Zhao, Huishi Guo, Xiaoyun Qin and Zhen Zhang
Nanomaterials 2024, 14(19), 1579; https://doi.org/10.3390/nano14191579 - 30 Sep 2024
Cited by 1 | Viewed by 1097
Abstract
Cardiac troponin I (cTnI) monitoring is of great value in the clinical diagnosis of acute myocardial infarction (AMI). In this paper, a highly sensitive electrochemical aptamer sensor using polystyrene (PS) microspheres as the electrode substrate material in combination with Prussian blue (PB) and [...] Read more.
Cardiac troponin I (cTnI) monitoring is of great value in the clinical diagnosis of acute myocardial infarction (AMI). In this paper, a highly sensitive electrochemical aptamer sensor using polystyrene (PS) microspheres as the electrode substrate material in combination with Prussian blue (PB) and gold nanoparticles (AuNPs) was demonstrated for the sensitive and label-free determination of cTnI. PS microspheres were synthesized by emulsion polymerization and then dropped onto the glassy carbon electrode (GCE); PB and AuNPs were electrodeposited on the electrode in corresponding electrolyte solutions step by step. The PS microsphere substrate provided a large surface area for the loading mass of the biological affinity aptamers, while the PB layer improved the electrical conductivity of the modified electrode, and the electroactive AuNPs exhibited excellent catalytic performance for the subsequent electrochemical measurements. In view of the above mentioned AuNPs/PB/PS/GCE sensing platform, the fabricated label-free electrochemical aptamer sensor exhibited a wide detection range of 10 fg/mL~1.0 μg/mL and a low detection limit of 2.03 fg/mL under the optimal conditions. Furthermore, this biosensor provided an effective detection platform for the analysis of cTnI in serum samples. The introduction of this sensitive electrochemical aptamer sensor provides a reference for clinically sensitive detection of cTnI. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of (<b>A</b>) PS, (<b>B</b>) PB/PS, and (<b>C</b>) AuNPs/PB/PS microspheres. (<b>D</b>) EDAX spectrum of AuNPs/PB/PS. (<b>E</b>) The mapping images of a single sphere for AuNPs/PB/PS, (magenta) Au, (green) Fe, (red) C, and (yellow) O elemental spectra.</p>
Full article ">Figure 2
<p>(<b>A</b>) XRD pattern of AuNPs/PB/PS (red curve). (<b>B</b>) FT-IR spectra of PS (black curve), PB/PS (red curve), and AuNPs/PB/PS (blue curve).</p>
Full article ">Figure 3
<p>(<b>A</b>,<b>C</b>) CV curves and (<b>B</b>,<b>D</b>) EIS plots of (a) GCE, (b) PS/GCE, (c) PB/PS/GCE, and (d) AuNPs/PB/PS/GCE. (<b>C</b>) CV curves and (<b>D</b>) EIS plots of (e) Tro4/AuNPs/PB/PS/GCE, (f) MCH/Tro4/AuNPs/PB/PS/GCE, and (g) cTnI/MCH/Tro4/AuNPs/PB/PS/GCE. Dotted line, raw data; solid line, fitted result.</p>
Full article ">Figure 4
<p>CV plots of (<b>A</b>) AuNPs/PB/PS/GCE and (<b>B</b>) AuNPs/PB/GCE at different sweep rates. From inside to outside: 10 (curve a), 20 (curve b), 40 (curve c), 60 (curve d), 80 (curve e), 100 (curve f), 150 (curve g), 200 (curve h), 300 (curve i), 400 (curve j), and 500 mV/s (curve k). (<b>C</b>) Plot of the oxidation peak current versus the square root of the sweep rate (ν<sup>1/2</sup>) for the (a) AuNPs/PB/PS/GCE and (b) AuNPs/PB/GCE. (<b>D</b>) A histogram comparing the corresponding electrochemically active areas of GCE (yellow column), AuNPs/PB/GCE (violet column), and AuNPs/PB/PS/GCE (red column).</p>
Full article ">Figure 5
<p>(<b>A</b>) Effect of different concentrations of KNO<sub>3</sub> on the peak current density. (<b>B</b>) Effect of different incubation time of Tro4 on the peak current density. (<b>C</b>) Effect of different blocking time of MCH on the peak current density. (<b>D</b>) Effect of the binding time between Tro4/AuNPs/PB/PS/GCE and cTnI on the peak current density.</p>
Full article ">Figure 6
<p>(<b>A</b>) Typical DPV current density response of Tro4/AuNPs/PB/PS/GCE to a series of cTnI concentrations in 0.8 M KNO<sub>3</sub> solution, 10 fg/mL (curve a), 0.1 pg/mL (curve b), 1 pg/mL (curve c), 10 pg/mL (curve d), 0.1 ng/mL (curve e), 1 ng/mL (curve f), 10 ng/mL (curve g), 0.1 μg/mL (curve h), 1.0 μg/mL (curve i). (<b>B</b>) The linear correlation between the current density value and the negative logarithm of the cTnI concentration (−lg <span class="html-italic">C</span><sub>[cTnI]</sub>). (<b>C</b>) Selectivity and anti-interference performance of the aptamer sensor. (<b>D</b>) Stability of the sensor over a period of 55 days; the error bars indicate the standard deviation (<span class="html-italic">n</span> = 5).</p>
Full article ">Figure 7
<p>The feasibility of the aptamer sensor in simulated AMI samples.</p>
Full article ">Scheme 1
<p>Schematic diagram of the construction procedure of the Tro4/AuNPs/PB/PS/GCE aptamer sensor and its electrochemical sensing for cTnI.</p>
Full article ">
17 pages, 5007 KiB  
Article
The Effect of Fly Ash Additive on the Thermal Conductivity of Polystyrene Concrete
by Rassul B. Tlegenov, Rimma K. Niyazbekova, Assel E. Jexembayeva, Kinga Korniejenko, Lyazat B. Aruova, Saule S. Aldabergenova and Aslan S. Maykonov
Buildings 2024, 14(9), 2850; https://doi.org/10.3390/buildings14092850 - 10 Sep 2024
Cited by 1 | Viewed by 1083
Abstract
The use of fly ash in compositions as a substitute for a part of cement is economically favorable and ecologically feasible in connection with large accumulations of waste at the enterprises of the energy sector. In addition, the technology of cement production provides [...] Read more.
The use of fly ash in compositions as a substitute for a part of cement is economically favorable and ecologically feasible in connection with large accumulations of waste at the enterprises of the energy sector. In addition, the technology of cement production provides high-temperature treatment of mineral substances in kilns with significant emissions of carbon dioxide. One of the most effective directions of the utilization of fly ash is their use in concrete composites. The use of this material will provide the required temperature and humidity conditions in residential premises, solve the problem of “cold bridges” in structures, minimize heat losses of the structure, and increase the energy efficiency of buildings in general. At the same time, polystyrene concrete, due to its structural structure and the presence of thermally conductive concrete, has limited opportunities for thermal and physical–mechanical properties. To improve the operational properties of polystyrene concrete, it is proposed to use composite binders, including fly ash from the thermal power station of Astana. The main aim of this study is to develop compositions of polystyrene concrete with reduced thermal conductivity and improved physical and mechanical properties. The objectives of this study include the determination of characteristics of fly ash from Astana, formulation of polystyrene concrete mixtures with different proportions of fly ash, and evaluation of their thermal conductivity properties. These tasks are in line with the objectives of the ISO 50001 standard to improve energy efficiency and reduce environmental impact. The results showed that the addition of fly ash from Astana to polystyrene concrete leads to a marked reduction in thermal conductivity, contributing to improved energy efficiency of the building envelope. Optimal results were achieved by using 15% of Astana fly ash as an additive in polystyrene concrete, which led to a significant reduction in thermal conductivity of 51.47%. This reduction is in line with improving the energy efficiency of building materials, especially in cold climates. Full article
(This article belongs to the Section Building Materials, and Repair & Renovation)
Show Figures

Figure 1

Figure 1
<p>Flowchart for provided research.</p>
Full article ">Figure 2
<p>Particle size distribution: (<b>a</b>) fly ash sample; (<b>b</b>) ash samples from ash dump; (<b>c</b>) ash samples from hydraulic ash dump.</p>
Full article ">Figure 2 Cont.
<p>Particle size distribution: (<b>a</b>) fly ash sample; (<b>b</b>) ash samples from ash dump; (<b>c</b>) ash samples from hydraulic ash dump.</p>
Full article ">Figure 3
<p>X-ray diffraction results of fly ash with phase decoding were obtained by analyzing the position of the main diffraction reflexes and their coincidence with card values from the PDF-2 database.</p>
Full article ">Figure 4
<p>Phase composition of fly ash.</p>
Full article ">Figure 5
<p>Prepared samples of polystyrene concrete: (<b>a</b>) polystyrene concrete outside and inside, (<b>b</b>) polystyrene concrete outside, (<b>c</b>) samples during preliminary research.</p>
Full article ">Figure 6
<p>Thermal conductivity after 3 days.</p>
Full article ">Figure 7
<p>Thermal conductivity after 7 days.</p>
Full article ">Figure 8
<p>Thermal conductivity after 28 days.</p>
Full article ">Figure 9
<p>Microphotographs of quenched specimens without additives after 28 days of curing, ×1000.</p>
Full article ">
24 pages, 4098 KiB  
Article
Multi-Objective Optimization of Building Design Parameters for Cost Reduction and CO2 Emission Control Using Four Different Algorithms
by Ahmet Serhan Canbolat and Emre İsa Albak
Appl. Sci. 2024, 14(17), 7668; https://doi.org/10.3390/app14177668 - 30 Aug 2024
Cited by 2 | Viewed by 1353
Abstract
Thermal insulation applications on the exterior facades of buildings have been the subject of numerous studies from the past to the present. Some of these studies focus on the cost reduction effect of insulation, while others emphasize its ecological benefits. In this study, [...] Read more.
Thermal insulation applications on the exterior facades of buildings have been the subject of numerous studies from the past to the present. Some of these studies focus on the cost reduction effect of insulation, while others emphasize its ecological benefits. In this study, multi-objective optimization, the objectives of which are minimum cost and minimum CO2 emission, has been carried out with the NSGA-II method. In emission calculations, in addition to fuel-related emissions, the carbon footprint of all materials comprising the wall has also been included. The multi-objective optimization study examined four design variables: wall thickness, wall material (light concrete, reinforced concrete, and brick), insulation material (expanded polystyrene, extruded polystyrene, mineral wool, and polyurethane foam), and heating source (natural gas, electricity, fuel oil). Analyses have been carried out for four cities (Osmaniye, Bursa, Isparta, and Erzurum), which are located in different climatic regions, and considering solar radiation effects. An existing building has been taken as the base case scenario, and the study has determined the improvements in the total cost and the amount of CO2 released into the environment when the appropriate insulation material, insulation thickness, wall material, and heating source identified in the multi-objective optimization study have been used. At the cost-oriented optimum point in the study, the most suitable insulation material was found to be expanded polystyrene, the most suitable wall material was brick, and the most suitable heating source was natural gas. In the CO2-oriented optimum, in contrast to the cost-oriented approach, optimal results have been obtained when light concrete was selected as the wall material. Full article
Show Figures

Figure 1

Figure 1
<p>Wall Components of the Base Case Scenario.</p>
Full article ">Figure 2
<p>The flowchart of NSGA-II.</p>
Full article ">Figure 3
<p>Different climatic zones of Türkiye and selected cities for the study.</p>
Full article ">Figure 4
<p>Annual Average Heating and Cooling Degree Days for Different Zones.</p>
Full article ">Figure 5
<p>Effect of design variables on total cost.</p>
Full article ">Figure 6
<p>Effect of design variables on CO<sub>2</sub>.</p>
Full article ">Figure 7
<p>CO<sub>2</sub> and total cost values at optimum points obtained for different objective functions (Zone I).</p>
Full article ">Figure 8
<p>CO<sub>2</sub> and total cost values at optimum points obtained for different objective functions (Zone II).</p>
Full article ">Figure 9
<p>CO<sub>2</sub> and total cost values at optimum points obtained for different objective functions (Zone III).</p>
Full article ">Figure 10
<p>CO<sub>2</sub> and total cost values at optimum points obtained for different objective functions (Zone IV).</p>
Full article ">Figure 11
<p>Percentage improvement in total cost and CO<sub>2</sub> emissions for different insulation thicknesses under various optimization objectives across climate zones.</p>
Full article ">Figure 12
<p>Cost components variation for different insulation thicknesses in the hottest and coldest climate.</p>
Full article ">Figure 13
<p>Emission components variation for different insulation thicknesses in the hottest and coldest climate.</p>
Full article ">
21 pages, 3902 KiB  
Brief Report
Enhancing Magnetic Micro- and Nanoparticle Separation with a Cost-Effective Microfluidic Device Fabricated by Laser Ablation of PMMA
by Cristian F. Rodríguez, Paula Guzmán-Sastoque, Carolina Muñoz-Camargo, Luis H. Reyes, Johann F. Osma and Juan C. Cruz
Micromachines 2024, 15(8), 1057; https://doi.org/10.3390/mi15081057 - 22 Aug 2024
Cited by 2 | Viewed by 1776
Abstract
Superparamagnetic iron oxide micro- and nanoparticles have significant applications in biomedical and chemical engineering. This study presents the development and evaluation of a novel low-cost microfluidic device for the purification and hyperconcentration of these magnetic particles. The device, fabricated using laser ablation of [...] Read more.
Superparamagnetic iron oxide micro- and nanoparticles have significant applications in biomedical and chemical engineering. This study presents the development and evaluation of a novel low-cost microfluidic device for the purification and hyperconcentration of these magnetic particles. The device, fabricated using laser ablation of polymethyl methacrylate (PMMA), leverages precise control over fluid dynamics to efficiently separate magnetic particles from non-magnetic ones. We assessed the device’s performance through Multiphysics simulations and empirical tests, focusing on the separation of magnetite nanoparticles from blue carbon dots and magnetite microparticles from polystyrene microparticles at various total flow rates (TFRs). For nanoparticle separation, the device achieved a recall of up to 93.3 ± 4% and a precision of 95.9 ± 1.2% at an optimal TFR of 2 mL/h, significantly outperforming previous models, which only achieved a 50% recall. Microparticle separation demonstrated an accuracy of 98.1 ± 1% at a TFR of 2 mL/h in both simulations and experimental conditions. The Lagrangian model effectively captured the dynamics of magnetite microparticle separation from polystyrene microparticles, with close agreement between simulated and experimental results. Our findings underscore the device’s robust capability in distinguishing between magnetic and non-magnetic particles at both micro- and nanoscales. This study highlights the potential of low-cost, non-cleanroom manufacturing techniques to produce high-performance microfluidic devices, thereby expanding their accessibility and applicability in various industrial and research settings. The integration of a continuous magnet, as opposed to segmented magnets in previous designs, was identified as a key factor in enhancing magnetic separation efficiency. Full article
Show Figures

Figure 1

Figure 1
<p>Computational domain and boundary conditions. (<b>a</b>) Computational domain, and mesh used for the simulation. (<b>b</b>) Microfluidic device manufactured in PMMA using laser ablation. The particle inlet is shown in green and labeled ‘1’, and the water buffer inlet, facilitating the washing and separation of magnetic from non-magnetic particles, is in blue and labeled ‘2’. The outlet, marked in red and denoted by ‘3’, is where non-magnetic particles exit first, followed by magnetic particles after the magnet is removed.</p>
Full article ">Figure 2
<p>Manufacturing process of microfluidic devices using CO<sub>2</sub> laser ablation in PMMA. (1) The microfluidic device is manufactured in AutoCAD (AutoDesk Inc., Mill Valley, CA, USA). (2) The design is transferred to a CO<sub>2</sub> laser system for engraving and cutting. A 2 mm-thick PMMA sheet is engraved to a depth of 1 mm to create the microfluidic channels, while a 4 mm-thick PMMA sheet is cut to form the inlets and outlets. (3) Then, PMMA layers are cleaned with a 70% ethanol solution to remove any residues. (4) Next, layers are bonded using 96% ethanol, pressure, and heat at 110 °C for 3 min. (5) Finally, the inlets and outlets are assembled into the microfluidic device.</p>
Full article ">Figure 3
<p>Synthesis and functionalization of magnetite micro- and nanoparticles. (<b>a</b>) Schematic of the magnetite micro- and nanoparticle synthesis using the coprecipitation technique. The process involves the preparation of an iron chloride solution, followed by the addition of NaOH. Rapid addition yields micro-sized particles (~2405 nm), while slow, dropwise addition results in nano-sized particles (~155 nm). (<b>b</b>) Silanization of magnetite nanoparticles to functionalize their surface and facilitate further modifications. (<b>c</b>) Subsequent labeling of silanized magnetite nanoparticles with rhodamine B (Rhod-B). (<b>d</b>) Synthesis of carbon dots through a separate process involving heating, sonication, and filtration, yielding purified carbon dots.</p>
Full article ">Figure 4
<p>Evaluation of magnetic nanoparticle separation microfluidic device. (<b>a</b>) Simulated magnetic flux density distribution around the continuous magnet. (<b>b</b>) Distribution of the magnetic scalar potential across the microfluidic device, indicating areas of maximum potential (up to ±2 amperes). (<b>c</b>) Trajectory simulation of magnetic nanoparticles in the microfluidic channel, demonstrating their response to the applied magnetic field. (<b>d</b>) Photographic evidence of nanoparticle retention within the microfluidic device, aligning with areas of high magnetic flux density. (<b>e</b>) Comparative bar graph showcasing recall, precision, and accuracy metrics for nanoparticle separation at varying total flow rates (2, 20, and 200 mL/h), as obtained from both in silico simulations (dark shades) and experimental results (light shades).</p>
Full article ">Figure 5
<p>Microparticle separation efficacy: (<b>a</b>) Fluorescence microscope images: non-magnetic polystyrene particles exhibit red fluorescence, while magnetite particles are non-fluorescent and visible in the transmitted BF (Bright Field) overlay. Scale-bar 50 <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>. (<b>b</b>) Analysis of recall, precision, and accuracy, derived from the fluorescence microscope images, illustrating the device’s performance.</p>
Full article ">Figure 6
<p>Microparticle sedimentation in microchannels: Bright Field and fluorescent images show minimal sedimentation of rhodamine B-labeled polystyrene microparticles, primarily in curved regions. The sedimentation is minimal and does not significantly obstruct the microchannel or impact the device’s functionality. Black arrows show the direction flow.</p>
Full article ">Figure 7
<p>SWOT analysis of the magnetic separator microfluidic device. Strengths (S), weaknesses (W), opportunities (O), and threats (T).</p>
Full article ">
12 pages, 955 KiB  
Article
Charge Phenomena in the Elastic Backscattering of Electrons from Insulating Polymers
by Maurizio Dapor
Polymers 2024, 16(16), 2329; https://doi.org/10.3390/polym16162329 - 17 Aug 2024
Cited by 1 | Viewed by 729
Abstract
Elastic peak electron spectroscopy (EPES) analyzes the line shape of the elastic peak. The reduction in energy of the elastic peak electrons is the result of energy transfer to the target atoms, a phenomenon known as recoil energy. EPES differs from other electron [...] Read more.
Elastic peak electron spectroscopy (EPES) analyzes the line shape of the elastic peak. The reduction in energy of the elastic peak electrons is the result of energy transfer to the target atoms, a phenomenon known as recoil energy. EPES differs from other electron spectroscopies in its unique ability to identify hydrogen in polymers and hydrogenated carbon-based materials. This feature is particularly noteworthy as lighter elements exhibit stronger energy shifts. The energy difference between the positions of the elastic peak of carbon and the elastic peak of hydrogen tends to increase as the kinetic energy of the incident electrons increases. During electron irradiation of an insulating polymer, if the number of secondary electrons emitted from the surface is less than the number of electrons absorbed in the sample, the surface floats energetically until it stabilizes at a potential energy eVs. As a result, the interaction energy changes and modifies the energy difference between the elastic peaks of hydrogen and carbon. In this study, the charge effects are evaluated using the Monte Carlo method to simulate the EPES spectra of electrons interacting with polystyrene and polyethylene. Full article
(This article belongs to the Section Polymer Analysis and Characterization)
Show Figures

Figure 1

Figure 1
<p>A flow chart of the Monte Carlo simulation. It describes the motion of each electron of the incident beam. The procedure shown here was repeated 10<sup>10</sup> times to obtain each of the Monte Carlo simulated spectra presented in this paper.</p>
Full article ">Figure 2
<p>Elastic scattering cross-section of electrons hitting hydrogen and carbon [<a href="#B12-polymers-16-02329" class="html-bibr">12</a>] (solid lines) in comparison with the calculations of Mayol and Salvat [<a href="#B31-polymers-16-02329" class="html-bibr">31</a>] (symbols).</p>
Full article ">Figure 3
<p>Cumulative probabilities of elastic scattering of 1500 eV and 2000 eV electrons hitting hydrogen and carbon atoms [<a href="#B12-polymers-16-02329" class="html-bibr">12</a>].</p>
Full article ">Figure 4
<p>Inelastic mean free path of electrons impinging on polystyrene and polyethylene, according to calculations by Tanuma, Powell, and Penn [<a href="#B34-polymers-16-02329" class="html-bibr">34</a>].</p>
Full article ">Figure 5
<p>The Monte Carlo simulation of the EPES of 1500 eV electrons impinging on polystyrene, taking charge effects into account. Two values of the surface energy potential <math display="inline"><semantics> <mrow> <mi>e</mi> <msub> <mi>V</mi> <mi>s</mi> </msub> </mrow> </semantics></math>, i.e., 100 eV and 200 eV, are considered, and the corresponding MC spectra are compared with the spectrum obtained without considering the charge (<math display="inline"><semantics> <mrow> <mi>e</mi> <msub> <mi>V</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> eV). The experimental data of Filippi and Calliari [<a href="#B11-polymers-16-02329" class="html-bibr">11</a>] are also presented. The simulations were performed while reproducing the experimental conditions, i.e., the electron beam hit the sample surface at an angle of 30° in the surface normal direction, and the acceptance scattering angle was 138<math display="inline"><semantics> <mrow> <mspace width="3.33333pt"/> <mo>±</mo> <mspace width="3.33333pt"/> <msup> <mn>6</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. The MC simulation was performed assuming 5% hydrogen desorption (induced by the electron irradiation) and taking into account the Doppler broadening. The spectra shown here as a function of energy loss were normalized to a common height of the elastic carbon peak and aligned so that the elastic carbon peak was at 0 energy loss.</p>
Full article ">Figure 6
<p>The Monte Carlo simulation of the EPES of 2000 eV electrons impinging on polyethylene, taking charge effects into account. Two values of the surface energy potential <math display="inline"><semantics> <mrow> <mi>e</mi> <msub> <mi>V</mi> <mi>s</mi> </msub> </mrow> </semantics></math>, i.e., 100 eV and 200 eV, are considered, and the corresponding MC spectra are compared with the spectrum obtained without considering the charge (<math display="inline"><semantics> <mrow> <mi>e</mi> <msub> <mi>V</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> eV). The experimental data of Orosz et al. [<a href="#B13-polymers-16-02329" class="html-bibr">13</a>] are also presented. The simulations were performed while reproducing the experimental conditions, i.e., the electron beam hit the sample surface at an angle of 50° in the surface normal direction, and the detection scattering angle was 0° in the surface normal direction. The MC simulation was performed assuming 2.2% hydrogen desorption (induced by the electron irradiation) and taking into account the Doppler broadening. The spectra shown here as a function of energy loss were normalized to a common height of the elastic carbon peak and aligned so that the elastic carbon peak was at 0 energy loss.</p>
Full article ">
15 pages, 7026 KiB  
Article
Isoconversional Analysis of the Catalytic Pyrolysis of ABS, HIPS, PC and Their Blends with PP and PVC
by Maria-Anna Charitopoulou, Evangelia C. Vouvoudi and Dimitris S. Achilias
Polymers 2024, 16(16), 2299; https://doi.org/10.3390/polym16162299 - 14 Aug 2024
Viewed by 981
Abstract
Thermochemical recycling of plastics in the presence of catalysts is often employed to facilitate the degradation of polymers. The choice of the catalyst is polymer-oriented, while its selection becomes more difficult in the case of polymeric blends. The present investigation studies the catalytic [...] Read more.
Thermochemical recycling of plastics in the presence of catalysts is often employed to facilitate the degradation of polymers. The choice of the catalyst is polymer-oriented, while its selection becomes more difficult in the case of polymeric blends. The present investigation studies the catalytic pyrolysis of polymers abundant in waste electric and electronic equipment (WEEE), including poly(acrylonitrile-butadiene-styrene) (ABS), high-impact polystyrene (HIPS) and poly(bisphenol-A carbonate) (PC), along with their blends with polypropylene (PP) and poly(vinyl chloride) (PVC). The aim is to study the kinetic mechanism and estimate the catalysts’ effect on the activation energy of the degradation. The chosen catalysts were Fe2O3 for ABS, Al-MCM-41 for HIPS, Al2O3 for PC, CaO for Blend A (comprising ABS, HIPS, PC and PP) and silicalite for Blend B (comprising ABS, HIPS, PC, PP and PVC). Thermogravimetric experiments were performed in a N2 atmosphere at several heating rates. Information on the degradation mechanism (degradation steps, initial and final degradation temperature, etc.) was attained. It was found that for pure (co)polymers, the catalytic degradation occurred in one-step, whereas in the case of the blends, two steps were required. For the estimation of the activation energy of those degradations, isoconversional kinetic models (integral and differential) were employed. In all cases, the catalysts used were efficient in reducing the estimated Eα, compared to the values of Eα obtained from conventional pyrolysis. Full article
(This article belongs to the Section Polymer Physics and Theory)
Show Figures

Figure 1

Figure 1
<p>Charts of mass loss vs. temperature for (<b>a</b>) ABS/Fe<sub>2</sub>O<sub>3</sub> samples, (<b>c</b>) HIPS/Al-MCM-41 samples and (<b>e</b>) PC/Al<sub>2</sub>O<sub>3</sub> samples. Charts of differential TG curves vs. temperature for (<b>b</b>) ABS/Fe<sub>2</sub>O<sub>3</sub> samples, (<b>d</b>) HIPS/Al-MCM-41 samples and (<b>f</b>) PC/Al<sub>2</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 2
<p>Charts of mass loss vs. temperature for (<b>a</b>) Blend A/CaO samples and (<b>c</b>) Blend B/silicalite samples. Charts of differential TG curves vs. temperature for (<b>b</b>) Blend A/CaO samples and (<b>d</b>) Blend B/silicalite samples, respectively [<a href="#B33-polymers-16-02299" class="html-bibr">33</a>,<a href="#B35-polymers-16-02299" class="html-bibr">35</a>,<a href="#B36-polymers-16-02299" class="html-bibr">36</a>].</p>
Full article ">Figure 3
<p>Variation in (<b>a</b>) <span class="html-italic">α</span> vs. <span class="html-italic">T</span> and (<b>b</b>) d<span class="html-italic">α</span>/d<span class="html-italic">t</span> vs. <span class="html-italic">α</span> curves for ABS/Fe<sub>2</sub>O<sub>3</sub>; (<b>c</b>) <span class="html-italic">α</span> vs. <span class="html-italic">T</span> and (<b>d</b>) d<span class="html-italic">α</span>/d<span class="html-italic">t</span> vs. <span class="html-italic">α</span> curves for HIPS/Al-MCM-41; and (<b>e</b>) <span class="html-italic">α</span> vs. <span class="html-italic">T</span> and (<b>f</b>) d<span class="html-italic">α</span>/d<span class="html-italic">t</span> vs. <span class="html-italic">α</span> curves for PC/Al<sub>2</sub>O<sub>3</sub>, respectively.</p>
Full article ">Figure 4
<p>Variation in <span class="html-italic">α</span> vs. <span class="html-italic">T</span> (<b>a</b>) and d<span class="html-italic">α</span>/d<span class="html-italic">t</span> vs. <span class="html-italic">α</span> curves (<b>b</b>) for Blend A/CaO; and <span class="html-italic">α</span> vs. <span class="html-italic">T</span> (<b>c</b>) and d<span class="html-italic">α</span>/d<span class="html-italic">t</span> vs. <span class="html-italic">α</span> curves (<b>d</b>) for the Blend B/silicalite systems.</p>
Full article ">Figure 5
<p>Comparisons of the effect of the catalyst on the variation in the activation energy (<span class="html-italic">E<sub>α</sub></span>) with the extent of degradation (<span class="html-italic">α</span>) for (<b>a</b>) ABS/Fe<sub>2</sub>O<sub>3</sub> using the Friedman method, (<b>b</b>) ABS/Fe<sub>2</sub>O<sub>3</sub> using the KAS method, (<b>c</b>) HIPS/Al-MCM-41 using the Friedman method, (<b>d</b>) HIPS/Al-MCM-41 using the KAS method, (<b>e</b>) PC/Al<sub>2</sub>O<sub>3</sub> using the Friedman method, (<b>f</b>) PC/Al<sub>2</sub>O<sub>3</sub> using the KAS method, (<b>g</b>) Blend A/CaO using the Friedman method, (<b>h</b>) Blend A/CaO using the KAS method and, finally, (<b>i</b>) Blend B/silicalite using the Friedman method and (<b>j</b>) Blend B/silicalite using the KAS method.</p>
Full article ">
16 pages, 7650 KiB  
Article
Effect of High-Tenacity Polypropylene Fibers on the Carbonation Resistance of Expanded Polystyrene Concrete
by Shifang Wang, Shangquan Xu, Yong Han, Weiqi Dong, Zhicheng Zhang, Kaisheng Yu, Wei Lin, Ji Yuan, Haijie He, Hongjian Lin, Wen Xu and Zhiyuan Ren
Buildings 2024, 14(8), 2480; https://doi.org/10.3390/buildings14082480 - 11 Aug 2024
Cited by 2 | Viewed by 958
Abstract
Expanded polystyrene concrete (EPSC) is increasingly utilized in buildings as a green building material. To investigate the effect of high-tenacity polypropylene (HTPP) fibers on the carbonation resistance (CR) of EPSC, five groups of EPSC specimens with HTPP fiber volume fractions of 0%, 0.6%, [...] Read more.
Expanded polystyrene concrete (EPSC) is increasingly utilized in buildings as a green building material. To investigate the effect of high-tenacity polypropylene (HTPP) fibers on the carbonation resistance (CR) of EPSC, five groups of EPSC specimens with HTPP fiber volume fractions of 0%, 0.6%, 0.9%, 1.2%, and 1.5% were prepared. Rapid carbonation tests were conducted to measure the carbonation depth (CD) and uniaxial compression strength (UCS) of the specimens at different carbonation ages (3, 7, 14, and 28 days). The CD and UCS of the specimens were calculated and analyzed. The results indicated that the HTPP fibers dramatically improved the CR of EPSC, with a decrease in the CD of up to 29.5% at 28 days. A model for predicting the CD of EPSC was developed. The model for the strength after carbonation also showed good agreement with the experimental results. Scanning electron microscopy (SEM) was used to examine the microstructure of the HTPP-reinforced EPSC, while the mechanism of HTPP fibers to enhance the CR of EPSC was elucidated. The findings of this study provide valuable insights for the application of EPSC as a structural material. Full article
(This article belongs to the Section Building Materials, and Repair & Renovation)
Show Figures

Figure 1

Figure 1
<p>EPS particles.</p>
Full article ">Figure 2
<p>HTPP fibers.</p>
Full article ">Figure 3
<p>Preparation of EPSC specimens.</p>
Full article ">Figure 4
<p>Carbonation of group EP00: (<b>a</b>) 3 days; (<b>b</b>) 7 days; (<b>c</b>) 14 days; (<b>d</b>) 28 days.</p>
Full article ">Figure 5
<p>Carbonation of group EP06: (<b>a</b>) 3 days; (<b>b</b>) 7 days; (<b>c</b>) 14 days; (<b>d</b>) 28 days.</p>
Full article ">Figure 6
<p>Carbonation of group EP09: (<b>a</b>) 3 days; (<b>b</b>) 7 days; (<b>c</b>) 14 days; (<b>d</b>) 28 days.</p>
Full article ">Figure 7
<p>Carbonation of group EP12: (<b>a</b>) 3 days; (<b>b</b>) 7 days; (<b>c</b>) 14 days; (<b>d</b>) 28 days.</p>
Full article ">Figure 8
<p>Carbonation of group EP15: (<b>a</b>) 3 days; (<b>b</b>) 7 days; (<b>c</b>) 14 days; (<b>d</b>) 28 days.</p>
Full article ">Figure 9
<p>Comparison of CDs.</p>
Full article ">Figure 10
<p>The 28-day UCS.</p>
Full article ">Figure 11
<p>Strength at various carbonation ages.</p>
Full article ">Figure 12
<p>Relative strength versus carbonation age.</p>
Full article ">Figure 13
<p>Fitted curves of relative strength. (<b>a</b>) EP00, (<b>b</b>) EP06, (<b>c</b>) EP09, (<b>d</b>) EP12, (<b>e</b>) EP15.</p>
Full article ">Figure 13 Cont.
<p>Fitted curves of relative strength. (<b>a</b>) EP00, (<b>b</b>) EP06, (<b>c</b>) EP09, (<b>d</b>) EP12, (<b>e</b>) EP15.</p>
Full article ">Figure 14
<p>CD versus carbonation age. (<b>a</b>) EP00, (<b>b</b>) EP06, (<b>c</b>) EP09, (<b>d</b>) EP12, (<b>e</b>) EP15.</p>
Full article ">Figure 14 Cont.
<p>CD versus carbonation age. (<b>a</b>) EP00, (<b>b</b>) EP06, (<b>c</b>) EP09, (<b>d</b>) EP12, (<b>e</b>) EP15.</p>
Full article ">Figure 15
<p>SEM images of specimen internal structure. (<b>a</b>) EP00 (300×), (<b>b</b>) EP00 (90×), (<b>c</b>) EP12 (30×), (<b>d</b>) EP12 (90×), (<b>e</b>) EP12 (90×), (<b>f</b>) EP12 (900×).</p>
Full article ">Figure 15 Cont.
<p>SEM images of specimen internal structure. (<b>a</b>) EP00 (300×), (<b>b</b>) EP00 (90×), (<b>c</b>) EP12 (30×), (<b>d</b>) EP12 (90×), (<b>e</b>) EP12 (90×), (<b>f</b>) EP12 (900×).</p>
Full article ">Figure 16
<p>Relationship between carbonation service life of EPSC components and fiber content.</p>
Full article ">
16 pages, 2743 KiB  
Article
Evaluation of Microbial Degradation of Thermoplastic and Thermosetting Polymers by Environmental Isolates
by Pierluca Nuccetelli, Francesca Maisto, Lucia Kraková, Alfredo Grilli, Alžbeta Takáčová, Alena Opálková Šišková and Domenico Pangallo
Coatings 2024, 14(8), 982; https://doi.org/10.3390/coatings14080982 - 3 Aug 2024
Viewed by 1190
Abstract
In this study, a microbial–enzymatic strategy was pursued to address the challenge of degrading thermoplastic and thermosetting polymers. Environmental microorganisms were isolated, and their enzymatic activities were assessed using colorimetric assays to evaluate their potential for producing enzymes capable of degrading these polymers. [...] Read more.
In this study, a microbial–enzymatic strategy was pursued to address the challenge of degrading thermoplastic and thermosetting polymers. Environmental microorganisms were isolated, and their enzymatic activities were assessed using colorimetric assays to evaluate their potential for producing enzymes capable of degrading these polymers. Microorganisms demonstrating higher activity in the enzymatic assays were selected for a 30-day biodegradation experiment, in which epoxy resins, polyethylene terephthalate, or polystyrene served as the sole carbon source. The effectiveness of biodegradation was assessed through the ATR-FTIR analysis of the chemical composition and the SEM examination of surface characteristics before and after degradation. The results indicated that thermoplastic compounds were more susceptible to microbial degradation, exhibiting greater changes in absorbance. In particular, PET treated with Stenotrophomonas sp. showed the most significant efficacy, achieving a 60.18% reduction in the area under the curve with a standard error of ± 3.42 when analyzed by FTIR spectroscopy. Significant alterations in surface morphology were noticed in thermoplastic compounds. In contrast, thermosetting compounds demonstrated lower reactivity, as evidenced by the absence of band shifts in FTIR spectra and minor changes in bond absorbance and surface morphology. Full article
(This article belongs to the Section Functional Polymer Coatings and Films)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overlapped mode of PET (<b>a</b>), PS (<b>b</b>), RS1 (<b>c</b>), and RS2 (<b>d</b>) after 30-day microbial degradation carried out by 5 microorganisms. With control representing pristine PET sample, <span class="html-italic">Rahnella</span> sp. (S5), <span class="html-italic">Stenotrophomonas</span> sp. (S6), <span class="html-italic">Sporobolomyces roseus</span> (S9), <span class="html-italic">Bullera alba</span> (S10), and <span class="html-italic">Bacillus tropicus</span> (ORAS6).</p>
Full article ">Figure 2
<p>Images acquired through SEM. Sample of PET control (<b>a</b>,<b>b</b>), and samples of PET degraded by <span class="html-italic">Rahnella</span> sp. (<b>c</b>,<b>d</b>) and <span class="html-italic">Stenotrophomonas</span> sp. (<b>e</b>,<b>f</b>) observed under different magnifications.</p>
Full article ">Figure 3
<p>The area under the curve for six absorption peaks characteristic of PET, with PET_C representing the pristine sample. <span class="html-italic">Rahnella</span> sp. (PET_S5), <span class="html-italic">Stenotrophomonas</span> sp. (PET_S6), <span class="html-italic">Sporobolomyces roseus</span> (PET_S9), <span class="html-italic">Bullera alba</span> (PET_S10), and <span class="html-italic">Bacillus tropicus</span> (PET_ORAS6).</p>
Full article ">Figure 4
<p>Images acquired through SEM. Sample of PS control (<b>a</b>,<b>b</b>) and samples of PS degraded by <span class="html-italic">Rahnella</span> sp. (<b>c</b>,<b>d</b>) and <span class="html-italic">Bacillus tropicus</span> (<b>e</b>,<b>f</b>) observed under different magnifications.</p>
Full article ">Figure 5
<p>Images acquired through SEM. Sample of RS1 control observed at 1400× (<b>a</b>). Samples of RS1 degraded by <span class="html-italic">Bullera alba</span> observed at 1400× (<b>b</b>) and by <span class="html-italic">Bacillus tropicus</span> observed at 1400× (<b>c</b>). Sample of RS2 control observed at 1400× (<b>d</b>). Samples of RS2 degraded by <span class="html-italic">Rahnella</span> sp. observed at 1400× (<b>e</b>) and by <span class="html-italic">Bacillus tropicus</span> observed at 1400× (<b>f</b>).</p>
Full article ">
Back to TopTop