[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (932)

Search Parameters:
Keywords = crosslink density

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 2520 KiB  
Article
Three-Dimensional Printing Multi-Drug Delivery Core/Shell Fiber Systems with Designed Release Capability
by Hao Wei, Yongxiang Luo, Ruisen Ma and Yuxiao Li
Pharmaceutics 2023, 15(9), 2336; https://doi.org/10.3390/pharmaceutics15092336 - 18 Sep 2023
Cited by 3 | Viewed by 1444
Abstract
A hydrogel system with the ability to control the delivery of multiple drugs has gained increasing interest for localized disease treatment and tissue engineering applications. In this study, a triple-drug-loaded model based on a core/shell fiber system (CFS) was fabricated through the co-axial [...] Read more.
A hydrogel system with the ability to control the delivery of multiple drugs has gained increasing interest for localized disease treatment and tissue engineering applications. In this study, a triple-drug-loaded model based on a core/shell fiber system (CFS) was fabricated through the co-axial 3D printing of hydrogel inks. A CFS with drug 1 loaded in the core, drug 2 in the shell part, and drug 3 in the hollow channel of the CFS was printed on a rotating collector using a co-axial nozzle. Doxorubicin (DOX), as the model drug, was selected to load in the core, with the shell and channel part of the CFS represented as drugs 1, 2, and 3, respectively. Drug 2 achieved the fastest release, while drug 3 showed the slowest release, which indicated that the three types of drugs printed on the CFS spatially can achieve sequential triple-drug release. Moreover, the release rate and sustained duration of each drug could be controlled by the unique core/shell helical structure, the concentration of alginate gels, the cross-linking density, the size and number of the open orifices in the fibers, and the CFS. Additionally, a near-infrared (NIR) laser or pH-responsive drug release could also be realized by introducing photo-thermal materials or a pH-sensitive polymer into this system. Finally, the drug-loaded system showed effective localized cancer therapy in vitro and in vivo. Therefore, this prepared CFS showed the potential application for disease treatment and tissue engineering by sequential- or stimulus-responsively releasing multi-drugs. Full article
Show Figures

Figure 1

Figure 1
<p>The photograph and SEM images of the fabricated CFS (<b>A</b>); the schematic diagram and photographs of drugs loading in the core (drug 1), shell (drug 2), and channel (drug 3) of the CFS (<b>B</b>); DOX release from the core part (drug 1) of the CFS cross-linked with different concentrations of CaCl<sub>2</sub>, <span class="html-italic">n</span> = 4 (<b>C</b>); DOX release from the shell layer (drug 2) of the CFS with different concentrations of alginate, <span class="html-italic">n</span> = 4 (<b>D</b>); DOX release from the channel (drug 3) of the CFS with different concentrations of alginate, <span class="html-italic">n</span> = 4 (<b>E</b>); DOX release from different part of the CFS, <span class="html-italic">n</span> = 4 (<b>F</b>); SEM images of the prepared PLGA microspheres (<b>G</b>) and PLGA microspheres embed in alginate shell of CFS (<b>H</b>); and DEX release from microspheres in the shell layer of the CFS, <span class="html-italic">n</span> = 4 (<b>I</b>).</p>
Full article ">Figure 2
<p>The schematic indicating the NIR-triggered on-demand drug release from prepared CFS (<b>A</b>), and the photographs of the DOX-loaded CFS before and after drug release (<b>B</b>); heating curves of the CFS with different concentrations of PDA (0.8 W cm<sup>−2</sup>, 808 nm NIR) (<b>C</b>) under different laser power (1% PDA) (<b>D</b>); photo-thermal conversion cycling test of the CFS (<b>E</b>); DOX release from the CFS with and without laser irradiation (0.8 W cm<sup>−2</sup>) (<b>F</b>); and DOX release rate with and without laser irradiation <span class="html-italic">n</span> = 4 (<b>G</b>).</p>
Full article ">Figure 3
<p>The flow cytometry of groups with CFS loading DOX (<b>A</b>), PTX (<b>B</b>), DOX + PTX (<b>C</b>), and without treatment (<b>D</b>). The cell viability of the groups with different treatments <span class="html-italic">n</span> = 4 (<b>E</b>). Live/dead staining of tumor cells (<b>F</b>). (Scale bar: 100 μm).</p>
Full article ">Figure 4
<p>The tumor volume changes curves <span class="html-italic">n</span> = 4 (<b>A</b>). The body weight curves of mice <span class="html-italic">n</span> = 4 (<b>B</b>). The digital photograph of tumors collected from the killed mice (<b>C</b>). H&amp;E-stained images of major organs (heart, liver, spleen, lung, and kidney) of mice (<b>D</b>) (Scale bar: 100 μm).</p>
Full article ">Figure 5
<p>FTIR spectra (<b>A</b>) and 1H NMR spectra (<b>B</b>) of PCL, MAC, and PCL–MAC; digital photograph of PCL–MAC hydrogel in different pHs for 0 h and 0.5 h (<b>C</b>); the swelling ratios curves of PCL–MAC hydrogel in different pHs, <span class="html-italic">n</span> = 4 (<b>D</b>); the SEM view of the PCL–MAC hydrogel in pH 7.4 (<b>E</b>) and 1.2 (<b>F</b>) after freeze-dry; the vancomycin release in PCL–MAC hydrogel in pH 7.4 and 1.2 conditions <span class="html-italic">n</span> = 4 (<b>G</b>).</p>
Full article ">
21 pages, 2767 KiB  
Article
Effect of Elastomeric Coating on the Properties and Performance of Myristic Acid (MA) Phase Change Material (PCM) Used for Photovoltaic Cooling
by Faisal Khaled Aldawood, Yamuna Munusamy, Mohamed Kchaou and Mohammad Alquraish
Coatings 2023, 13(9), 1606; https://doi.org/10.3390/coatings13091606 - 14 Sep 2023
Cited by 3 | Viewed by 1419
Abstract
Nitrile butadiene rubber (NBR) latex exhibits excellent tensile properties, chemical resistance, and thermal stability in applications such as gloves and safety shoes due to vulcanization. In this research work, attempts have been made to manipulate the vulcanization to produce thin and compact elastomeric [...] Read more.
Nitrile butadiene rubber (NBR) latex exhibits excellent tensile properties, chemical resistance, and thermal stability in applications such as gloves and safety shoes due to vulcanization. In this research work, attempts have been made to manipulate the vulcanization to produce thin and compact elastomeric NBR coating on myristic acid (MA) phase change material (PCM) to produce shape-stabilized PCM. The proposal for the use of latex-based elastomeric coating for PCM has been rarely considered in the literature due to a lack of understanding of the crosslink of elastomers. Thus, in this research, the effects of sulfur formulation on the coating performance of NBR on the PCM in terms of latent heat and thermal stability were determined. Leakage analysis indicates that the MA pellet coated with 0.5 phr of sulfur-cured NBR layer (MA/NBR-0.5) successfully eliminates the leakage issue. A tensile analysis revealed that a durable PCM coating layer must possess a combination of the following criteria: high tensile strength, ductility, and flexibility. Fourier transform infrared analysis (FTIR) and electron microscopy images showed the formation of thin, compact, and continuous NBR coating when 0.5 phr of sulfur was used. The further increment of sulfur loading between 1.0 and 1.5 phr causes the formation of defects on the coating layers, while non-vulcanized NBR layers seem to be very weak to withstand the phase-change process. The recorded latent heat values of melting and freezing of MA/NBR-0.5 are 142.30 ± 1.38 and 139.47 ± 1.23 J/g, respectively. The latent heat of the shape-stabilized MA/NBR-0.5 PCM is reduced by 32.24% from the pure MA latent heat density. This reduction is significantly lower than the reported latent heat reduction in shape-stabilized PCMs in other works. The thermal cycle test highlights the durability of the coated PCMs by withstanding up to 1000 thermal cycles (2.7 years) with less than 2% changes in latent heat value. Cooling performance test on photovoltaic (PV) module shows that the fabricated shape-stabilized PCM could reduce the temperature of the PV module up to 17 °C and increase the voltage generation by 7.92%. Actual performance analysis of shape-stabilized PCMs on the cooling of the PV module has been rarely reported and could be considered a strength of this work. Full article
(This article belongs to the Section Functional Polymer Coatings and Films)
Show Figures

Figure 1

Figure 1
<p>Dip coating method of MA PCMs. (<b>a</b>) Dip coating of MA PCM pellet with NBR latex, (<b>b</b>) drying of coated PCM on Teflon sheet.</p>
Full article ">Figure 2
<p>Leakage test setup.</p>
Full article ">Figure 3
<p>Leakage comparison of (<b>a</b>) uncoated MA pellet; (<b>b</b>) MA/NBR-0.5; (<b>c</b>) MA/NBR-1.0; and (<b>d</b>) MA/NBR-1.5 pellet after 60 thermal cyclic processes.</p>
Full article ">Figure 4
<p>(<b>a</b>) Ultimate tensile strength, (<b>b</b>) Young’s modulus, and (<b>c</b>) elongation at break of pure NBR and NBR films with different sulfur formulations.</p>
Full article ">Figure 5
<p>FTIR spectra of pure MA, pure NBR, and MA/NBR pellets with different sulfur formulation.</p>
Full article ">Figure 6
<p>FESEM surface morphology of (<b>a</b>) pure NBR; (<b>b</b>) MA/NBR-0.5; (<b>c</b>) MA/NBR-1.0; (<b>d</b>) MA/NBR-1.5 pellets with magnification of 5000×.</p>
Full article ">Figure 7
<p>FESEM cross-section morphology of (<b>a</b>) pure NBR; (<b>b</b>) MA/NBR-0.5; (<b>c</b>) MA/NBR-1.0; (<b>d</b>) MA/NBR-1.5 pellets with magnification of 250×.</p>
Full article ">Figure 7 Cont.
<p>FESEM cross-section morphology of (<b>a</b>) pure NBR; (<b>b</b>) MA/NBR-0.5; (<b>c</b>) MA/NBR-1.0; (<b>d</b>) MA/NBR-1.5 pellets with magnification of 250×.</p>
Full article ">Figure 8
<p>(<b>a</b>) Heating curve and (<b>b</b>) cooling curve of pure MA and MA/NBR pellets.</p>
Full article ">Figure 9
<p>DSC curves for (<b>a</b>) MA/NBR-0.5; (<b>b</b>) MA/NBR-1.0; and (<b>c</b>) MA/NBR-1.5 pellets before and after 1000 thermal cycle tests.</p>
Full article ">Figure 10
<p>Cooling effect of PCM on PV module.</p>
Full article ">Figure 11
<p>Cooling effect of PCM on PV module for voltage generation.</p>
Full article ">
23 pages, 12187 KiB  
Article
Embodied Energy in the Production of Guar and Xanthan Biopolymers and Their Cross-Linking Effect in Enhancing the Geotechnical Properties of Cohesive Soil
by M. Ashok Kumar, Arif Ali Baig Moghal, Kopparthi Venkata Vydehi and Abdullah Almajed
Buildings 2023, 13(9), 2304; https://doi.org/10.3390/buildings13092304 - 10 Sep 2023
Cited by 5 | Viewed by 2171
Abstract
Traditional soil stabilization techniques, such as cement and lime, are known for their menacing effect on the environment through heavy carbon emissions. Sustainable soil stabilization methods are grabbing attention, and the utilization of biopolymers is surely one among them. Recent studies proved the [...] Read more.
Traditional soil stabilization techniques, such as cement and lime, are known for their menacing effect on the environment through heavy carbon emissions. Sustainable soil stabilization methods are grabbing attention, and the utilization of biopolymers is surely one among them. Recent studies proved the efficiency of biopolymers in enhancing the geotechnical properties to meet the requirements of the construction industry. The suitability of biopolymer application in different soils is still unexplored, and the carbon footprint analysis (CFA) of biopolymers is crucial in promoting the biopolymers as a promising sustainable soil stabilization method. This study attempts to investigate the out-turn of cross-linked biopolymer on soils exhibiting different plasticity characteristics (Medium & High compressibility) and to determine the Embodied carbon factor (ECF) for the selected biopolymers. Guar (G) and Xanthan (X) biopolymers were cross-linked at different proportions to enhance the geotechnical properties of soils. Atterberg’s limits, Compaction characteristics, and Unconfined Compressive Strength were chosen as performance indicators, and their values were analyzed at different combinations of biopolymers before and after cross-linking. The test results have shown that Atterberg’s limits of the soils increased with the addition of biopolymers, and it is attributed to the formation of hydrogels in the soil matrix. Compaction test results reveal that the Optimum Moisture Content (OMC) of biopolymer-modified soil increased, and Maximum Dry Density (MDD) reduced due to the resistance offered by hydrogel against compaction effort. Soils amended with biopolymers and cured for 14, 28, and 60 days have shown an appreciable improvement in Unconfined Compressive Strength (UCS) results. Microlevel analysis was carried out using SEM (Scanning Electron Microscopy) and FTIR (Fourier-transform infrared spectroscopy) to formulate the mechanism responsible for the alteration in targeted performance indicators due to the cross-linking of biopolymers in the soil. The embodied energy in the production of both Guar and Xanthan biopolymers was calculated, and the obtained ECF values were 0.087 and 1.67, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>Variation of Atterberg’s limits for soil S1 and S2 at different biopolymer dosages.</p>
Full article ">Figure 2
<p>Variation of Optimum Moisture Content (OMC) for soil S1 and S2 at different biopolymer dosages.</p>
Full article ">Figure 3
<p>Variation of Maximum Dry Density (MDD) for soil S1 and S2 at different biopolymer dosages.</p>
Full article ">Figure 4
<p>Mechanism of biopolymer interaction with soil particles.</p>
Full article ">Figure 5
<p>Variation of Unconfined Compression Strength at different curing periods for soil S1 and S2.</p>
Full article ">Figure 6
<p>SEM images showing changes in the microstructure arrangement of 28 days cured sample of soil S1 (<b>a</b>) Virgin soil (<b>b</b>) 2G (<b>c</b>) 2X (<b>d</b>) 1G + 1X.</p>
Full article ">Figure 7
<p>SEM images showing changes in the microstructure arrangement of 28 days cured sample of soil S2 (<b>a</b>) Virgin soil (<b>b</b>) 2G (<b>c</b>) 2X (<b>d</b>) 1G + 1X.</p>
Full article ">Figure 8
<p>Variation of FTIR frequencies for soil S1 at different biopolymer dosages.</p>
Full article ">Figure 9
<p>Variation of FTIR frequencies for soil S2 at different biopolymer dosages.</p>
Full article ">Figure 10
<p>Different stages in the production of guar gum [<a href="#B42-buildings-13-02304" class="html-bibr">42</a>].</p>
Full article ">Figure 11
<p>Seed structure of Guar Gum [<a href="#B42-buildings-13-02304" class="html-bibr">42</a>].</p>
Full article ">Figure 12
<p>Outline of the guar gum manufacturing process [<a href="#B42-buildings-13-02304" class="html-bibr">42</a>].</p>
Full article ">Figure 13
<p>Outline of the xanthan gum manufacturing process [<a href="#B44-buildings-13-02304" class="html-bibr">44</a>].</p>
Full article ">
15 pages, 4084 KiB  
Article
Magnetic and Viscoelastic Response of Magnetorheological Elastomers Based on a Combination of Iron Nano- and Microparticles
by Imperio Anel Perales-Martínez, Luis Manuel Palacios-Pineda, Alex Elías-Zúñiga, Daniel Olvera-Trejo, Karina Del Ángel-Sánchez, Isidro Cruz-Cruz, Claudia Angélica Ramírez-Herrera and Oscar Martínez-Romero
Polymers 2023, 15(18), 3703; https://doi.org/10.3390/polym15183703 - 8 Sep 2023
Cited by 1 | Viewed by 1162
Abstract
In this paper, we discuss the creation of a hybrid magnetorheological elastomer that combines nano- and microparticles. The mixture contained 45 wt.% fillers, with combinations of either 0% nanoparticles and 100% microparticles or 25% nanoparticles and 75% microparticles. TGA and FTIR testing confirmed [...] Read more.
In this paper, we discuss the creation of a hybrid magnetorheological elastomer that combines nano- and microparticles. The mixture contained 45 wt.% fillers, with combinations of either 0% nanoparticles and 100% microparticles or 25% nanoparticles and 75% microparticles. TGA and FTIR testing confirmed the materials’ thermal and chemical stability, while an SEM analysis determined the particles’ size and morphology. XRD results were used to determine the crystal size of both nano- and microparticles. The addition of reinforcing particles, particularly nanoparticles, enhanced the stiffness of the composite materials studied, but their overall strength was only minimally affected. The computed interaction parameter relative to the volume fraction was consistent with the previous literature. Furthermore, the study observed a magnetic response increment in composite materials reinforced with nanoparticles above 30 Hz. The isotropic material containing only microparticles had a lower storage modulus than the isotropic sample with nanoparticles without a magnetic field. However, when a magnetic field was applied, the material with only microparticles exhibited a higher storage modulus than the samples with nanoparticles. Full article
(This article belongs to the Special Issue Magnetic Polymer Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Publications reported by Web of Science. The gray line represents all the papers that deal with magnetorheological materials (MRMs) from 1979 until this year. The blue line indicates the studies performed on magnetorheological elastomers regardless of the kind of reinforced material. Finally, the orange line denotes only those papers that investigate magnetorheological elastomers manufactured with nanoparticles (MRE Nano). The vertical axis is on a log scale.</p>
Full article ">Figure 2
<p>FTIR spectra of magnetorheological samples manufactured to different combinations of nano and microparticles.</p>
Full article ">Figure 3
<p>Diffractograms of (<b>a</b>) nano and microparticle powder and (<b>b</b>) different ratios of iron nano/microparticles as fillers of PDMS-based magnetorheological materials. The inlet figure represents a zoom-in for the (110) crystallographic plane of CIPs, and its Gaussian fit was applied to find the full width at half maximum (FWHM), denoted as β, to determine crystalline size using the Scherrer equation.</p>
Full article ">Figure 4
<p>Size distribution and micrography of magnetic particles: (<b>a</b>) nanosize and (<b>b</b>) microsize recorded by SEM analysis.</p>
Full article ">Figure 5
<p>Optical microscopy image that shows the magnetic particle alignment into the PDMS matrix of the sample Ani 25 N 75 M.</p>
Full article ">Figure 6
<p>(<b>a</b>) TGA analysis and (<b>b</b>) DTGA curves used to investigate the influence of the combination of nano and microparticles in the material thermal stability.</p>
Full article ">Figure 7
<p>Stress versus stretch response curves were obtained from the bare polymeric matrix sample and the composite material samples reinforced with nano and microparticles. Notice that the material stiffness increases with the presence of filling material. These curves were used to find the sample’s shear modulus and the crosslink chain density using Equations (3) and (4) along with the swelling test data. (<b>a</b>) General view of the tensile test curves, (<b>b</b>) a zoomed view in the zone of low stretch.</p>
Full article ">Figure 8
<p>(<b>a</b>) Young’s modulus and tensile strength and (<b>b</b>) Tensile strength and ultimate stretch for the bare and composite material obtained from the tensile test. All samples have 45% wt. of magnetic particles.</p>
Full article ">Figure 9
<p>Parallel plate rheological test in the frequency domain. The strain is kept constant at <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math>, with a constant normal force of 1.5 N and a constant field intensity of 0 T and 1 T. (<b>a</b>) Samples with only microparticles, (<b>b</b>) samples with nano- and microparticles, (<b>c</b>) anisotropic samples, and (<b>d</b>) isotropic samples.</p>
Full article ">Figure 10
<p>Storage modulus versus magnetic field. Notice that at higher magnetic fields, the isotropic specimens made with iron microparticles attained the highest storage modulus value.</p>
Full article ">Figure 11
<p>Cole–Cole plots showing the influence of the magnetic flux density on the composite materials. (<b>a</b>) Samples with only microparticles, (<b>b</b>) samples with nano- and microparticles, (<b>c</b>) anisotropic samples, and (<b>d</b>) isotropic samples.</p>
Full article ">Figure 12
<p>Interaction parameter values obtained from swelling tests as a function of the volume fraction. Comparison with the data from Chahal [<a href="#B39-polymers-15-03703" class="html-bibr">39</a>], Schuld [<a href="#B40-polymers-15-03703" class="html-bibr">40</a>], and Palacios-Pineda [<a href="#B24-polymers-15-03703" class="html-bibr">24</a>].</p>
Full article ">
17 pages, 19021 KiB  
Article
Atomistic Construction of Silicon Nitride Ceramic Fiber Molecular Model and Investigation of Its Mechanical Properties Based on Molecular Dynamics Simulations
by Yiqiang Hong, Yu Zhu, Youpei Du, Zhe Che, Guoxin Qu, Qiaosheng Li, Tingting Yuan, Wei Yang, Zhen Dai, Weijian Han and Qingsong Ma
Materials 2023, 16(18), 6082; https://doi.org/10.3390/ma16186082 - 5 Sep 2023
Cited by 3 | Viewed by 1124
Abstract
Molecular simulations are currently receiving significant attention for their ability to offer a microscopic perspective that explains macroscopic phenomena. An essential aspect is the accurate characterization of molecular structural parameters and the development of realistic numerical models. This study investigates the surface morphology [...] Read more.
Molecular simulations are currently receiving significant attention for their ability to offer a microscopic perspective that explains macroscopic phenomena. An essential aspect is the accurate characterization of molecular structural parameters and the development of realistic numerical models. This study investigates the surface morphology and elemental distribution of silicon nitride fibers through TEM and EDS, and SEM and EDS analyses. Utilizing a customized molecular dynamics approach, molecular models of amorphous and multi-interface silicon nitride fibers with complex structures were constructed. Tensile simulations were conducted to explore correlations between performance and molecular structural composition. The results demonstrate successful construction of molecular models with amorphous, amorphous–crystalline interface, and mixed crystalline structures. Mechanical property characterization reveal the following findings: (1) The nonuniform and irregular amorphous structure causes stress concentration and crack formation under applied stress. Increased density enhances material strength but leads to higher crack sensitivity. (2) Incorporating a crystalline reinforcement phase without interfacial crosslinking increases free volume and relative tensile strength, improving toughness and reducing crack susceptibility. (3) Crosslinked interfaces effectively enhance load transfer in transitional regions, strengthening the material’s tensile strength, while increased density simultaneously reduces crack propagation. Full article
(This article belongs to the Section Materials Simulation and Design)
Show Figures

Figure 1

Figure 1
<p>TEM and EDS, and SEM and EDS element content and distribution map of silicon nitride fiber.</p>
Full article ">Figure 2
<p>Construction algorithm flowchart and elementary structure of amorphous silicon nitride ceramic fiber model.</p>
Full article ">Figure 3
<p>Flow chart and structure diagram of molecular model construction of amorphous and crystal structure transition interface, the color standard is based on the actual image. (<b>a1</b>): Flow chart of structure generation algorithm, (<b>b1</b>–<b>b4</b>): interface structure generation diagram, (<b>c1</b>–<b>c4</b>): interface structure density projection diagram, (<b>d1</b>,<b>d2</b>): raw material structure diagram, (<b>e1</b>): interface structure model diagram.</p>
Full article ">Figure 4
<p>Construction of structure diagram using mixed molecular model of amorphous and crystal structure.</p>
Full article ">Figure 5
<p>Mechanical properties of amorphous and crystal transition structures; the red ball represents the two atoms where the bond breaks, the figure (I–VI) show the structural transformation process of interface stretching.</p>
Full article ">Figure 6
<p>Measurement of stress–strain behavior of silicon nitride ceramic fiber molecular model. (<b>a</b>): the tensile curve of different density structure, (<b>b</b>): the tensile curve of non-cross-linked structure, (<b>c</b>): the tensile curve of cross-linked structure.</p>
Full article ">Figure 7
<p>Stress–strain model of amorphous structure, the red ball represents the two atoms where the bond breaks, while the yellow and blue ball represent the crystal structure and there is no bond fracture on it. (<b>a-1</b>–<b>a-6</b>): Stretching molecular model of density 1.8 g/cm<sup>3</sup>, (<b>b-1</b>–<b>b-6</b>): Stretching molecular model of density 2.0 g/cm<sup>3</sup>, (<b>c-1</b>–<b>c-6</b>): Stretching molecular model of density 2.4 g/cm<sup>3</sup>.</p>
Full article ">Figure 8
<p>Stress–strain model of noncrosslinked crystal structure, the red ball represents the two atoms where the bond breaks, while the yellow and blue balls represent the crystal structure and there is no bond fracture on it. (<b>a-1</b>–<b>a-6</b>): Stretching molecular model of 10 no cross-linked crystals, (<b>b-1</b>–<b>b-6</b>): Stretching molecular model of 20 no cross-linked crystals, (<b>c-1</b>–<b>c-6</b>): Stretching molecular model of 30 no cross-linked crystals.</p>
Full article ">Figure 9
<p>Stress–strain model of crosslinked crystal structure, the red ball represents the two atoms where the bond breaks, while the yellow and blue balls represent the crystal structure and there is no bond fracture on it. (<b>a-1</b>–<b>a-6</b>): Stretching molecular model of 10 cross-linked crystals, (<b>b-1</b>–<b>b-6</b>): Stretching molecular model of 20 cross-linked crystals, (<b>c-1</b>–<b>c-6</b>): Stretching molecular model of 30 cross-linked crystals.</p>
Full article ">Figure 10
<p>Tensile fracture behavior of silicon nitride ceramic fiber molecular model. (<b>a-1</b>–<b>a-3</b>): fracture curve of different structures, (<b>b-1</b>–<b>b-3</b>): first derivative curve of fracture curve, (<b>c-1</b>–<b>c-3</b>): molecular model of fracture structure.</p>
Full article ">Figure 11
<p>Free volume change of silicon nitride ceramic fiber molecular model during drawing. (<b>a-1</b>–<b>a-3</b>): free volume curve of different structures, (<b>b-1</b>–<b>b-3</b>): first derivative curve of free volume curve, (<b>c-1</b>–<b>c-3</b>): molecular model of free volume.</p>
Full article ">
17 pages, 5928 KiB  
Article
Fibrinogen-Based Bioink for Application in Skin Equivalent 3D Bioprinting
by Aida Cavallo, Tamer Al Kayal, Angelica Mero, Andrea Mezzetta, Lorenzo Guazzelli, Giorgio Soldani and Paola Losi
J. Funct. Biomater. 2023, 14(9), 459; https://doi.org/10.3390/jfb14090459 - 5 Sep 2023
Cited by 4 | Viewed by 1964
Abstract
Three-dimensional bioprinting has emerged as an attractive technology due to its ability to mimic native tissue architecture using different cell types and biomaterials. Nowadays, cell-laden bioink development or skin tissue equivalents are still at an early stage. The aim of the study is [...] Read more.
Three-dimensional bioprinting has emerged as an attractive technology due to its ability to mimic native tissue architecture using different cell types and biomaterials. Nowadays, cell-laden bioink development or skin tissue equivalents are still at an early stage. The aim of the study is to propose a bioink to be used in skin bioprinting based on a blend of fibrinogen and alginate to form a hydrogel by enzymatic polymerization with thrombin and by ionic crosslinking with divalent calcium ions. The biomaterial ink formulation, composed of 30 mg/mL of fibrinogen, 6% of alginate, and 25 mM of CaCl2, was characterized in terms of homogeneity, rheological properties, printability, mechanical properties, degradation rate, water uptake, and biocompatibility by the indirect method using L929 mouse fibroblasts. The proposed bioink is a homogeneous blend with a shear thinning behavior, excellent printability, adequate mechanical stiffness, porosity, biodegradability, and water uptake, and it is in vitro biocompatible. The fibrinogen-based bioink was used for the 3D bioprinting of the dermal layer of the skin equivalent. Three different normal human dermal fibroblast (NHDF) densities were tested, and better results in terms of viability, spreading, and proliferation were obtained with 4 × 106 cell/mL. The skin equivalent was bioprinted, adding human keratinocytes (HaCaT) through bioprinting on the top surface of the dermal layer. A skin equivalent stained by live/dead and histological analysis immediately after printing and at days 7 and 14 of culture showed a tissuelike structure with two distinct layers characterized by the presence of viable and proliferating cells. This bioprinted skin equivalent showed a similar native skin architecture, paving the way for its use as a skin substitute for wound healing applications. Full article
(This article belongs to the Special Issue Recent Advances in Tissue Regeneration and Biomaterials Manufacturing)
Show Figures

Figure 1

Figure 1
<p>Designed pattern to calculate (<b>a</b>) spreading ratio and (<b>b</b>) shape fidelity on 4 and 8 overlapped layers; scale bar, 5 mm.</p>
Full article ">Figure 2
<p>(<b>a</b>) Extrusion force measured for the fibrinogen-based biomaterial ink using the dedicated setup implemented to assess the solution homogeneity; (<b>b</b>) filament-like shape of extruded biomaterial ink (indicated by black arrows) during the extrusion process.</p>
Full article ">Figure 3
<p>Rheological characterization of biomaterial ink at 25 and 37 °C: (<b>a</b>) flow curve as a function of shear rate; (<b>b</b>) curves of storage (G′) and loss (G″) modulus vs. angular frequency.</p>
Full article ">Figure 4
<p>Biomaterial ink printability assessment in terms of (<b>a</b>) filament collapse test, (<b>b</b>) spreading ratio, shape fidelity, (<b>c</b>) 4 layers, (<b>d</b>) 8 layers, and (<b>e</b>) printable angles; scale bar, 5 mm.</p>
Full article ">Figure 5
<p>(<b>a</b>) Mechanical compression tests on three crosslinked 3D-bioprinted constructs using the fibrinogen-based biomaterial ink, (<b>b</b>) degradation rate, (<b>c</b>) water uptake, and (<b>d</b>) swelling ratio of biomaterial ink; * <span class="html-italic">p</span> &lt; 0.05 between two consecutive time points.</p>
Full article ">Figure 6
<p>Surface and section SEM images of a crosslinked 3D-bioprinted construct using 50 UT/mL of bovine thrombin in 50 mM CaCl<sub>2</sub>.</p>
Full article ">Figure 7
<p>Live/dead staining of 3D-bioprinted samples with cell densities of (<b>a</b>–<b>f</b>) 1 × 10<sup>6</sup> NHDF/mL, (<b>g</b>–<b>i</b>) 2 × 10<sup>6</sup> NHDF/mL, and (<b>j</b>–<b>l</b>) 4 × 10<sup>6</sup> NHDF/mL at days 1, 7, and 14 of in vitro culture; scale bar, 100 µm.</p>
Full article ">Figure 8
<p>XTT assay on samples biofabricated using 1, 2, and 4 × 10<sup>6</sup> NHDF/mL at days 0, 3, 7, and 14 of in vitro culture. The absorbance values measured immediately after printing were assumed as 100% of cell viability. Data are presented as mean ± SD (n = 3). * <span class="html-italic">p</span> &lt; 0.05 statistically significant differences among the samples biofabricated with 4 × 10<sup>6</sup> NHDF/mL at days 3, 7, and 14 with respect to day 0; # <span class="html-italic">p</span> &lt; 0.05 statistically significant differences among the samples biofabricated with 4 × 10<sup>6</sup> NHDF/mL at days 3 and 7; <b>°</b> <span class="html-italic">p</span> &lt; 0.05 statistically significant differences among the samples biofabricated with 4 × 10<sup>6</sup> NHDF/mL at days 7 and 14.</p>
Full article ">Figure 9
<p>Staining of 3D-bioprinted construct with calcein AM of (<b>a</b>,<b>b</b>) NHDF cells at days 7 and 14 of in vitro culture and of (<b>c</b>,<b>d</b>) HaCaT cell at days 7 and 14 of in vitro culture; (<b>e</b>,<b>f</b>) calcein AM and Hoechst 33342 staining of HaCaT keratinocytes; scale bar, 100 µm.</p>
Full article ">Figure 10
<p>Representative image of H&amp;E staining at day 14 of culture of equivalent fabricated by HaCaT seeding 24 h after dermal layer bioprinting. HaCaT and NHDF cells are indicated by black and red arrows, respectively; scale bar, 200 µm.</p>
Full article ">
18 pages, 7060 KiB  
Article
A Study of Isosorbide Synthesis from Sorbitol for Material Applications Using Isosorbide Dimethacrylate for Enhancement of Bio-Based Resins
by Vojtěch Jašek, Jan Fučík, Jiří Krhut, Ludmila Mravcova, Silvestr Figalla and Radek Přikryl
Polymers 2023, 15(17), 3640; https://doi.org/10.3390/polym15173640 - 4 Sep 2023
Cited by 3 | Viewed by 2082
Abstract
Bio-based cross-linkers can fulfill the role of enhancing additives in bio-sourced curable materials that do not compare with artificial resin precursors. Isosorbide dimethacrylate (ISDMMA) synthesized from isosorbide (ISD) can serve as a cross-linker from renewable sources. Isosorbide is a bicyclic carbon molecule produced [...] Read more.
Bio-based cross-linkers can fulfill the role of enhancing additives in bio-sourced curable materials that do not compare with artificial resin precursors. Isosorbide dimethacrylate (ISDMMA) synthesized from isosorbide (ISD) can serve as a cross-linker from renewable sources. Isosorbide is a bicyclic carbon molecule produced by the reaction modification of sorbitol and the optimal conditions of this reaction were studied in this work. The reaction temperature of 130 °C and 1% w/w amount of para-toluenesulfonic acid (p-TSA) were determined as optimal and resulted in a yield of 81.9%. Isosorbide dimethacrylate was synthesized via nucleophilic substitution with methacrylic anhydride (MAA) with the conversion of 94.1% of anhydride. Formed ISD and ISDMMA were characterized via multiple verification methods (FT-IR, MS, 1H NMR, and XRD). Differential scanning calorimetry (DSC) proved the curability of ISDMMA (activation energy Ea of 146.2 kJ/mol) and the heat-resistant index of ISDMMA (Ts reaching value of 168.9) was determined using thermogravimetric analysis (TGA). Characterized ISDMMA was added to the precursor mixture containing methacrylated alkyl 3-hydroxybutyrates (methyl ester M3HBMMA and ethyl ester E3HBMMA), and the mixtures were cured via photo-initiation. The amount of ISDMMA cross-linker increased all measured parameters obtained via dynamic mechanical analysis (DMA), such as storage modulus (E’) and glass transition temperature (Tg), and the calculated cross-linking densities (νe). Therefore, the enhancement influence of bio-based ISDMMA on resins from renewable sources was confirmed. Full article
(This article belongs to the Special Issue Resin-Based Polymer Materials and Related Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The dependence of formed and condensed reaction water during the dehydration reaction resulting in the production of isosorbide on the reaction time at different temperatures; (<b>b</b>) Monitoring of the synthesis of isosorbide via LC–MS analysis for differing amounts of catalyst (p-TSA) at 130 °C.</p>
Full article ">Figure 2
<p>FT-IR spectrum of synthesized isosorbide.</p>
Full article ">Figure 3
<p>MS fragmentation spectrum of synthesized isosorbide.</p>
Full article ">Figure 4
<p>XRD spectrum of synthesized isosorbide.</p>
Full article ">Figure 5
<p><sup>1</sup>H NMR spectrum of isosorbide (d-chloroform, 500 MHz): δ (ppm) = 4.69 (dd,1H), 4.37 (m, 2H), 4.29 (q, 1H), 3.87 (m, 3H), 3.52 (dd, 1H), 2.67 (s, 1H), 1.69 (s, 1H).</p>
Full article ">Figure 6
<p>GC-FID analyzed the quantification of decreasing methacrylic anhydride (MAA) and increasing methacrylic acid (MA).</p>
Full article ">Figure 7
<p>FT-IR spectrum of synthesized isosorbide dimethacrylate.</p>
Full article ">Figure 8
<p>MS full scan spectrum of synthesized isosorbide dimethacrylate.</p>
Full article ">Figure 9
<p><sup>1</sup>H NMR spectrum of isosorbide dimethacrylate (d-chloroform, 500 MHz): δ (ppm) 6.22–6.16 (d, J = 5.6 Hz, 1H); 6.13–6.10 (d, J = 5.7 Hz, 1H); 5.73–5.59 (dd, J = 12.7, 5.3 Hz, 2H); 5.29–5.16 (m, 2H); 4.92–4.90 (dt, J = 7.0, 4.8 Hz, 1H); 4.55–4.52 (m, 1H); 4.05–3.87 (m, 4H); 2.02–1.94 (dd, J = 20.3, 5.8 Hz, 6H).</p>
Full article ">Figure 10
<p>(<b>a</b>) DSC graphs comparing the exothermic polymerization enthalpy release of isosorbide dimethacrylate at different heating ramps; (<b>b</b>) graphical representation of Kissinger’s theory of cured isosorbide dimethacrylate.</p>
Full article ">Figure 11
<p>TGA analysis curves of dependence of the weight and derivative weight of the isosorbide dimethacrylate (ISDMMA) sample on increasing temperature.</p>
Full article ">Figure 12
<p>Methacrylated alkyl monoesters of 3-hydroxybutanoic acid used for the preparation of resins containing isosorbide dimethacrylate.</p>
Full article ">Figure 13
<p>(<b>a</b>) DMA storage modulus curves of enhanced methacrylated methyl 3-hydroxybutyrate (M3HBMMA) resins; (<b>b</b>) DMA damping factor curves of enhanced M3HBMMA resins.</p>
Full article ">Figure 14
<p>(<b>a</b>) DMA storage modulus curves of enhanced methacrylated ethyl3-hydroxybutyrate (E3HBMMA) resins; (<b>b</b>) DMA damping factor curves of enhanced E3HBMMA resins.</p>
Full article ">Scheme 1
<p>The reaction mechanism of the dehydration of D-glucitol to 1,4-anhydro-D-glucitol in the presence of catalyst <span class="html-italic">p</span>-TSA (<span class="html-italic">para</span>-toluenesulfonic acid).</p>
Full article ">Scheme 2
<p>The reaction mechanism of the dehydration of 1,4-anhydro-D-glucitol to 1,4:3,6-dianhydro-D-glucitol in the presence of catalyst <span class="html-italic">p</span>-TSA (<span class="html-italic">para</span>-toluenesulfonic acid).</p>
Full article ">Scheme 3
<p>The mechanism of the nucleophilic substitution reaction leading to the formation of isosorbide dimethacrylate using methacrylic anhydride as the reactant.</p>
Full article ">
17 pages, 3622 KiB  
Article
Magnetically Controlled Hyaluronic Acid–Maghemite Nanocomposites with Embedded Doxorubicin
by Vasily Spiridonov, Zukhra Zoirova, Yuliya Alyokhina, Nikolai Perov, Mikhail Afanasov, Denis Pozdyshev, Daria Krjukova, Alexander Knotko, Vladimir Muronetz and Alexander Yaroslavov
Polymers 2023, 15(17), 3644; https://doi.org/10.3390/polym15173644 - 4 Sep 2023
Viewed by 1226
Abstract
The controllable delivery of drugs is a key task of pharmacology. For this purpose, a series of polymer composites was synthesized via the cross-linking of hyaluronate and a hyaluronate/polyacrylate mixture with Fe2O3 nanoparticles. The cross-linking imparts magnetic properties to the [...] Read more.
The controllable delivery of drugs is a key task of pharmacology. For this purpose, a series of polymer composites was synthesized via the cross-linking of hyaluronate and a hyaluronate/polyacrylate mixture with Fe2O3 nanoparticles. The cross-linking imparts magnetic properties to the composites, which are more pronounced for the ternary hyaluronate/polyacrylate/γ-Fe2O3 composites compared with the binary hyaluronate/Fe2O3 composites. When dispersed in water, the composites produce microsized hydrogel particles. Circulation of the ternary microgels in an aqueous solution at a speed of 1.84 cm/s can be stopped using a permanent external magnet with a magnetic flux density of 400 T. The composite hydrogels can absorb the antitumor antibiotic doxorubicin (Dox); the resulting constructs show their cytotoxicity to tumor cells to be comparable to the cytotoxicity of Dox itself. The addition of the hyaluronidase enzyme induces degradation of the binary and ternary microgels down to smaller particles. This study presents prospectives for the preparation of magnetically controlled biodegradable polymer carriers for the encapsulation of bioactive substances. Full article
(This article belongs to the Special Issue Polymer-Containing Nanomaterials: Synthesis, Properties, Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray diffraction patterns of the binary composites I (1), II (2), III (3) and IV (4).</p>
Full article ">Figure 2
<p>Mössbauer spectrum of the binary composite IV measured at 78 K.</p>
Full article ">Figure 3
<p>TEM image of the binary composite IV (<b>left</b>) and size distribution of the composite particles (<b>right</b>).</p>
Full article ">Figure 4
<p>IR spectra of HYAL (1) and the binary composite IV (2).</p>
Full article ">Figure 5
<p>Mössbauer spectrum of the Fe-containing composite IVpa with η = 38 wt% measured at 78 K.</p>
Full article ">Figure 6
<p>TEM image of the ternary composite IVpa (<b>left</b>) and size distribution of the composite IVpa particles (<b>right</b>).</p>
Full article ">Figure 7
<p>Magnetization curves of the composites at 300 °K. (<b>a</b>) I (1), II (2), III (3) and IV (4); (<b>b</b>) Ipa (1), IIIpa (2) and IVpa (3).</p>
Full article ">Figure 8
<p>(<b>a</b>) Microgel collected with the magnet on the inner surface of capillary. (<b>b</b>) No effect of the magnet on the microgel circulation. Composite IV (<b>a</b>) and composite IVpa (<b>b</b>); flow rate 18.4 cm/s, magnetic flux density 400 T.</p>
Full article ">Figure 9
<p>(<b>a</b>) Kinetics of hyaluronidase-induced degradation of HYAL (1, control) and the binary composites: I (2), II (3), III (4) and IV (5). (<b>b</b>) Kinetics of hyaluronidase-induced degradation of HYAL (6, control) and the ternary composites: Ipa (7), IpaI (8), IIIpa (9) and IVpa (10). Hyaluronidase conc. 1 μg/mL, HYAL conc. 1 mg/mL, pH 7.4, 37 °C.</p>
Full article ">Figure 10
<p>(<b>a</b>) Kinetics of hyaluronidase-induced degradation of HYAL (1, control) and the binary composites: I (2), II (3), III (4) and IV (5). (<b>b</b>) Kinetics of hyaluronidase-induced degradation of HYAL (6, control) and the ternary composites: Ipa (7), IIpa (8), IIIpa (9) and IVpa (10). Kinetics was controlled by measuring size of polymer and composite particles using dynamic light scattering. Hyaluronidase conc. 1 μg/mL, HYAL conc. 1 mg/mL, pH 7.4, 37 °C.</p>
Full article ">Figure 11
<p>Kinetics of hyaluronidase-induced increase in the fluorescence of Dox incorporated in HYAL (1) and the binary composites: I (2), II (3), III (4) and IV (5). Hyaluronidase conc. 1 μg/mL, HYAL conc. 0.1 mg/mL, [-COOH]/[Dox] = 5, pH 7.4, 37 °C.</p>
Full article ">
14 pages, 8083 KiB  
Article
Improved Thermal Insulation and Mechanical Strength of Styrene-Butadiene Rubber through the Combination of Filled Silica Aerogels and Modified Glass Fiber
by Guofeng Wang, Wenwen Yu, Sitong Zhang, Kaijie Yang, Wenying Liu, Jiaqi Wang and Fuyong Liu
Materials 2023, 16(17), 5947; https://doi.org/10.3390/ma16175947 - 30 Aug 2023
Cited by 1 | Viewed by 1471
Abstract
To improve heat dissipation capability and enhance mechanical properties, a series of silica aerogel (SA) and modified glass fiber (GF)-filled SBR composites were prepared. It was found that the addition of SA successfully reduced the thermal conductivity of SBR by 35%, owing to [...] Read more.
To improve heat dissipation capability and enhance mechanical properties, a series of silica aerogel (SA) and modified glass fiber (GF)-filled SBR composites were prepared. It was found that the addition of SA successfully reduced the thermal conductivity of SBR by 35%, owing to the heat shield of the nanoscale porous structure of SA. Moreover, the addition of modified glass fiber (MGF) yielded a significant increase in the tensile and tear strength of SBR/SA composite rubber of 37% and 15%, respectively. This enhancement was more pronounced than the improvement observed with unmodified GF, and was attributed to the improved dispersion of fillers and crosslinking density of the SBR matrix. Rheological analysis revealed that the addition of SA and MGF weakened the ω dependence. This was due to the partial relaxation of immobilized rubber chains and limited relaxation of rubber chains adsorbed on the MGF. Furthermore, the strain amplification effect of MGF was stronger than that of GF, leading to a more pronounced reinforcing effect. Full article
(This article belongs to the Topic Rubbers and Elastomers Materials)
Show Figures

Figure 1

Figure 1
<p>FT-IR infrared spectra of TESPT, MGF and GF.</p>
Full article ">Figure 2
<p>TGA curves of TESPT, MGF and GF.</p>
Full article ">Figure 3
<p>Thermal conductivity of SBR/SA/GF (<b>a</b>) and SBR/SA/MGF composites (<b>b</b>).</p>
Full article ">Figure 4
<p><span class="html-italic">G</span>′ as a function of <span class="html-italic">f</span> for SBR/SA/MGF (<b>a</b>), and SBR/SA/GF rubber composites (<b>b</b>).</p>
Full article ">Figure 5
<p>SEM images of SBR/SA/GF10% (<b>a</b>) and SBR/SA/MGF10% (<b>b</b>).</p>
Full article ">Figure 6
<p><span class="html-italic">G</span>′ (<b>a</b>) and <span class="html-italic">G</span>″ (<b>b</b>) as a function of <span class="html-italic">γ</span> for SBR rubber composites.</p>
Full article ">Figure 7
<p>The <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>φ</mi> </mrow> </mfenced> </mrow> </semantics></math> (<b>a</b>) and <math display="inline"><semantics> <mrow> <mi>f</mi> <mfenced separators="|"> <mrow> <mi>φ</mi> </mrow> </mfenced> </mrow> </semantics></math> (<b>b</b>) as a function of glass fiber contents.</p>
Full article ">Figure 8
<p>The vertical and horizontal translation of <span class="html-italic">G</span>′-<span class="html-italic">γ</span> relation curve.</p>
Full article ">Figure 9
<p>Vulcanization characteristic curves of rubber composites (<b>a</b>), ML and ΔS of SBR/SA/GF composites (<b>b</b>), and ML and ΔS of SBR/SA/MGF composites (<b>c</b>).</p>
Full article ">Figure 10
<p>Dynamic mechanical properties of SBR/SA/GF, SBR/SA/MGF composites, <span class="html-italic">E</span>′~T (<b>a</b>) and tan<span class="html-italic">δ</span>~T (<b>b</b>).</p>
Full article ">Figure 11
<p>The molecular weight between the crosslinking points of rubbers calculated by DMA.</p>
Full article ">Figure 12
<p>Tensile strength (<b>a</b>) and tear strength (<b>b</b>) of SBR/SA/GF and SBR/SA/MGF composites.</p>
Full article ">Figure 13
<p>SEM images of SBR/SA/GF5% (<b>a</b>), SBR/SA/GF10% (<b>b</b>), SBR/SA/GF15% (<b>c</b>), SBR/SA/GF20% (<b>d</b>), and SBR/SA/MGF5% (<b>e</b>), SBR/SA/MGF10% (<b>f</b>), SBR/SA/MGF15% (<b>g</b>), SBR/SA/MGF20% (<b>h</b>).</p>
Full article ">Figure 13 Cont.
<p>SEM images of SBR/SA/GF5% (<b>a</b>), SBR/SA/GF10% (<b>b</b>), SBR/SA/GF15% (<b>c</b>), SBR/SA/GF20% (<b>d</b>), and SBR/SA/MGF5% (<b>e</b>), SBR/SA/MGF10% (<b>f</b>), SBR/SA/MGF15% (<b>g</b>), SBR/SA/MGF20% (<b>h</b>).</p>
Full article ">
13 pages, 3408 KiB  
Article
3D Porous VOx/N-Doped Carbon Nanosheet Hybrids Derived from Cross-Linked Dicyandiamide–Chitosan Hydrogels for Superior Supercapacitor Electrode Materials
by Jinghua Liu, Xiong He, Jiayang Cai, Jie Zhou, Baosheng Liu, Shaohui Zhang, Zijun Sun, Pingping Su, Dezhi Qu and Yudong Li
Polymers 2023, 15(17), 3565; https://doi.org/10.3390/polym15173565 - 28 Aug 2023
Cited by 4 | Viewed by 991
Abstract
Three-dimensional porous carbon materials with moderate heteroatom-doping have been extensively investigated as promising electrode materials for energy storage. In this study, we fabricated a 3D cross-linked chitosan-dicyandiamide-VOSO4 hydrogel using a polymerization process. After pyrolysis at high temperature, 3D porous VOx/N-doped [...] Read more.
Three-dimensional porous carbon materials with moderate heteroatom-doping have been extensively investigated as promising electrode materials for energy storage. In this study, we fabricated a 3D cross-linked chitosan-dicyandiamide-VOSO4 hydrogel using a polymerization process. After pyrolysis at high temperature, 3D porous VOx/N-doped carbon nanosheet hybrids (3D VNCN) were obtained. The unique 3D porous skeleton, abundant doping elements, and presence of VOx 3D VNCN pyrolyzed at 800 °C (3D VNCN-800) ensured excellent electrochemical performance. The 3D VNCN-800 electrode exhibits a maximum specific capacitance of 408.1 F·g−1 at 1 A·g−1 current density and an admirable cycling stability with 96.8% capacitance retention after 5000 cycles. Moreover, an assembled symmetrical supercapacitor based on the 3D VNCN-800 electrode delivers a maximum energy density of 15.6 Wh·Kg−1 at a power density of 600 W·Kg−1. Our study demonstrates a potential guideline for the fabrication of porous carbon materials with 3D structure and abundant heteroatom-doping. Full article
(This article belongs to the Special Issue Environmentally Responsive Polymer Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Illustration of the polymerization process, (<b>b</b>) TGA curves of CS-DCDAVOSO<sub>4</sub> gel, CS-DCD gel, and CS-DCDA-VOSO<sub>4</sub> powder, (<b>c</b>) FT-TR spectra of CS-DCDA-VOSO<sub>4</sub> gel and 3D VNCN-800.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of 3D VNCN, 3D NCN, and VNC, (<b>b</b>) Raman spectra of 3D VNCN, 3D NCN, and VNC, (<b>c</b>) N<sub>2</sub> adsorption/desorption isotherms of 3D VNCN, 3D NCN, and VNC, (<b>d</b>) PSD curves of 3D VNCN, 3D NCN, and VNC.</p>
Full article ">Figure 3
<p>(<b>a</b>) XPS spectra of 3D VNCN and VNC, (<b>b</b>) V2p spectra of 3D VNCN and VNC, (<b>c</b>) N1s spectra of 3D VNCN and VNC, (<b>d</b>) the ratios of different nitrogen species determined from the N 1s XPS spectra.</p>
Full article ">Figure 4
<p>SEM images of (<b>a</b>) VNC, (<b>b</b>) 3D NCN, (<b>c</b>) CS-DCDA-VOSO<sub>4</sub> gel, (<b>d</b>,<b>e</b>) 3D VNCN-800, (<b>f</b>–<b>i</b>) elemental mapping of C, N, O, and V.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>d</b>) TEM images of 3D VNCN-800.</p>
Full article ">Figure 6
<p>(<b>a</b>) CV curves of 3D VNCN, 3D NCN, and VNC at 10 mV·s<sup>−1</sup>. (<b>b</b>) GCD curves of 3D VNCN, 3D NCN, and VNC at 0.5 A·g<sup>−1</sup>. (<b>c</b>,<b>d</b>) CV and GCD curves of 3D VNCN-800 at different scan rates and different current densities. (<b>e</b>,<b>f</b>) Cycling performance and Nyquist plots of 3D VNCN-800.</p>
Full article ">Figure 7
<p>(<b>a</b>) CV curves of 3D VNCN-800-SC at different scan rates. (<b>b</b>) GCD curves of 3D VNCN-800-SC at different current densities. (<b>c</b>) Relationship between the specific capacitance versus current densities of 3D VNCN-800-SC. (<b>d</b>) Ragone plots of 3D VNCN-800-SC.</p>
Full article ">
23 pages, 3397 KiB  
Article
Understanding Protein Functionality and Its Impact on Quality of Plant-Based Meat Analogues
by Jenna Flory, Ruoshi Xiao, Yonghui Li, Hulya Dogan, Martin J. Talavera and Sajid Alavi
Foods 2023, 12(17), 3232; https://doi.org/10.3390/foods12173232 - 28 Aug 2023
Cited by 1 | Viewed by 2855
Abstract
A greater understanding of protein functionality and its impact on processing and end-product quality is critical for the success of the fast-growing market for plant-based meat products. In this research, simple criteria were developed for categorizing plant proteins derived from soy, yellow pea, [...] Read more.
A greater understanding of protein functionality and its impact on processing and end-product quality is critical for the success of the fast-growing market for plant-based meat products. In this research, simple criteria were developed for categorizing plant proteins derived from soy, yellow pea, and wheat as cold swelling (CS) or heat swelling (HS) through various raw-material tests, including the water absorption index (WAI), least gelation concentration (LGC), rapid visco analysis (RVA), and % protein solubility. These proteins were blended together in different cold-swelling: heat-swelling ratios (0:100 to 90:10 or 0–90% CS) and extruded to obtain texturized vegetable proteins (TVPs). In general, the WAI (2.51–5.61 g/g) and protein solubility (20–46%) showed an increasing trend, while the LGC decreased from 17–18% to 14–15% with an increase in the % CS in raw protein blends. Blends with high CS (60–90%) showed a clear RVA cold viscosity peak, while low-CS (0–40%) blends exhibited minimal swelling. The extrusion-specific mechanical energy for low-CS blends (average 930 kJ/kg) and high-CS blends (average 949 kJ/kg) was similar, even though both were processed with similar in-barrel moisture, but the former had substantially lower protein content (69.7 versus 76.6%). Extrusion led to the aggregation of proteins in all treatments, as seen from the SDS-PAGE and SEC-HPLC analyses, but the protein solubility decreased the most for the high-CS (60–90%) blends as compared to the low-CS (0–40%) blends. This indicated a higher degree of crosslinking due to extrusion for high CS, which, in turn, resulted in a lower extruded TVP bulk density and higher water-holding capacity (average 187 g/L and 4.2 g/g, respectively) as compared to the low-CS treatments (average 226 g/L and 2.9 g/g, respectively). These trends matched with the densely layered microstructure of TVP with low CS and an increase in pores and a spongier structure for high CS, as observed using optical microscopy. The microstructure, bulk density, and WHC observations corresponded well with texture-profile-analysis (TPA) hardness of TVP patties, which decreased from 6949 to 3649 g with an increase in CS from 0 to 90%. The consumer test overall-liking scores (9-point hedonic scale) for TVP patties were significantly lower (3.8–5.1) as compared to beef hamburgers (7.6) (p < 0.05). The data indicated that an improvement in both the texture and flavor of the former might result in a better sensory profile and greater acceptance. Full article
(This article belongs to the Special Issue Functionality and Food Applications of Plant Proteins (Volume II))
Show Figures

Figure 1

Figure 1
<p>Screw profile for the pilot-scale extrusion trial.</p>
Full article ">Figure 2
<p>Water absorption index (WAI) for raw-material blends with varying amounts of cold-swelling proteins (0–90% CS). Different letters imply significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Rapid-visco-analysis (RVA) viscographs for individual ingredients.</p>
Full article ">Figure 4
<p>Rapid-visco-analysis (RVA) viscographs for raw-material blends with varying amounts of cold-swelling proteins (0–90% CS).</p>
Full article ">Figure 5
<p>Results from SDS-PAGE analysis, showing bands corresponding to different protein sub-units for treatment with varying amounts of cold-swelling proteins (0–90% CS) both before (<b>a</b>) and after (<b>b</b>) extrusion.</p>
Full article ">Figure 6
<p>SDS-PAGE results, showing the molecular weights of different protein subunits from individual ingredients.</p>
Full article ">Figure 7
<p>HPLC results for raw ingredients, showing peaks for protein fractions at different molecular weights.</p>
Full article ">Figure 8
<p>HPLC results for formulations with different cold-swelling protein concentrations (0–90% CS), showing peaks (representing protein fractions) before extrusion (<b>a</b>) compared to after extrusion (<b>b</b>).</p>
Full article ">Figure 9
<p>Water holding capacity (WHC) and off the dryer bulk density for extruded textured vegetable protein with different levels of cold-swelling proteins (0–90% CS). Error bars represent standard deviation. Different letters imply significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Texture profile analysis: hardness and chewiness of plant-based patties based on extruded TVP with different levels of cold-swelling proteins (0–90%). Error bars represent standard deviation. Different letters imply significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 11
<p>The internal structure of hydrated whole pieces of extruded TVP captured in the horizontal (perpendicular to flow exiting the extruder) and longitudinal directions (parallel to the flow) for products with different cold-swelling-protein concentrations (0–90% CS).</p>
Full article ">Figure 11 Cont.
<p>The internal structure of hydrated whole pieces of extruded TVP captured in the horizontal (perpendicular to flow exiting the extruder) and longitudinal directions (parallel to the flow) for products with different cold-swelling-protein concentrations (0–90% CS).</p>
Full article ">Figure 12
<p>The average overall-liking scores from the consumer study for plant-based meat patties with different levels of cold-swelling proteins (% CS) and beef hamburger. Different letters imply significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
16 pages, 3908 KiB  
Article
Superparamagnetic Nanocrystals Clustered Using Poly(ethylene glycol)-Crosslinked Amphiphilic Copolymers for the Diagnosis of Liver Cancer
by Ling Jiang, Jiaying Chi, Jiahui Wang, Shaobin Fang, Tingting Peng, Guilan Quan, Daojun Liu, Zhongjie Huang and Chao Lu
Pharmaceutics 2023, 15(9), 2205; https://doi.org/10.3390/pharmaceutics15092205 - 25 Aug 2023
Viewed by 1162
Abstract
Superparamagnetic iron oxide (SPIO) nanocrystals have been extensively studied as theranostic nanoparticles to increase transverse (T2) relaxivity and enhance contrast in magnetic resonance imaging (MRI). To improve the blood circulation time and enhance the diagnostic sensitivity of MRI contrast agents, we [...] Read more.
Superparamagnetic iron oxide (SPIO) nanocrystals have been extensively studied as theranostic nanoparticles to increase transverse (T2) relaxivity and enhance contrast in magnetic resonance imaging (MRI). To improve the blood circulation time and enhance the diagnostic sensitivity of MRI contrast agents, we developed an amphiphilic copolymer, PCPZL, to effectively encapsulate SPIO nanocrystals. PCPZL was synthesized by crosslinking a polyethylene glycol (PEG)-based homobifunctional linker with a hydrophobic star-like poly(ε-benzyloxycarbonyl-L-lysine) segment. Consequently, it could self-assemble into shell-crosslinked micelles with enhanced colloidal stability in bloodstream circulation. Notably, PCPZL could effectively load SPIO nanocrystals with a high loading capacity of 66.0 ± 0.9%, forming SPIO nanoclusters with a diameter of approximately 100 nm, a high cluster density, and an impressive T2 relaxivity value 5.5 times higher than that of Resovist®. In vivo MRI measurements highlighted the rapid accumulation and contrast effects of SPIO-loaded PCPZL micelles in the livers of both healthy mice and nude mice with an orthotopic hepatocellular carcinoma tumor model. Moreover, the magnetic micelles remarkably enhanced the relative MRI signal difference between the tumor and normal liver tissues. Overall, our findings demonstrate that PCPZL significantly improves the stability and magnetic properties of SPIO nanocrystals, making SPIO-loaded PCPZL micelles promising MRI contrast agents for diagnosing liver diseases and cancers. Full article
(This article belongs to the Special Issue Polymeric Micelles for Drug Delivery and Cancer Therapy)
Show Figures

Figure 1

Figure 1
<p>Schematic of the preparation of SPIO-loaded PCPZL micelles. (<b>A</b>) Synthesis route of PCPZL copolymer. (<b>B</b>) Schematic representation of shell-crosslinked magnetic micelles self-assembled from PCPZL copolymers and hydrophobic SPIO nanocrystals.</p>
Full article ">Figure 2
<p>Structural characterization of various copolymers. <sup>1</sup>H NMR spectrum of PZL2 (<b>A</b>) and PCPZL2 (<b>B</b>) in DMSO-<span class="html-italic">d</span><sub>6.</sub> (<b>C</b>) GPC chromatograms of various copolymers.</p>
Full article ">Figure 3
<p>Structural characterization of SPIO-loaded micelles. (<b>A</b>) Photograph of various SPIO-loaded micelle solutions. (<b>B</b>) SPIO loading capacity and (<b>C</b>) SPIO loading efficiency of various micelles. (<b>D</b>) TEM image of SPIO-loaded PCPZL2 micelles. Size distribution of (<b>E</b>) SPIO nanocrystals and (<b>F</b>) SPIO nanoclusters inside the SPIO-loaded PCPZL2 micelles with Gauss fitting. (<b>G</b>) Hydrodynamic diameter of SPIO-loaded PCPZL2 micelles determined using DLS. (<b>H</b>) Schematic of the structural composition of SPIO-loaded PCPZL2 micelles. (<b>I</b>) Fluorescence spectrum of pyrene in PCPZL2 aqueous solutions with a constant pyrene concentration (6 × 10<sup>−7</sup> mol L<sup>−1</sup>) and various concentrations of PCPZL2. (<b>J</b>) Plot of <span class="html-italic">I</span><sub>339</sub>/<span class="html-italic">I</span><sub>334</sub> of pyrene versus logarithm concentration of PCPZL2.</p>
Full article ">Figure 4
<p>Magnetization and T<sub>2</sub> relaxivity measurement of SPIO-loaded micelles. (<b>A</b>) Hysteresis loops of the SPIO nanocrystals and SPIO-loaded PCPZL2 micelles. (<b>B</b>) T<sub>2</sub> relaxation rate (1/T<sub>2</sub>) as a function of the iron concentration (mM) for SPIO-loaded PCPZL2 micelles, and the slope indicates the specific r<sub>2</sub>. (<b>C</b>) T<sub>2</sub>-weighted MRI images of SPIO-loaded PCPZL2 micelles at different iron concentrations.</p>
Full article ">Figure 5
<p>In vivo T<sub>2</sub>-weighted MRI performed on the liver of mice. (<b>A</b>) Schematic of mice intravenously injected with SPIO-loaded PCPZL2 micelles. (<b>B</b>) Time-dependence of the relative signal intensity in the livers, muscles, and bladders of mice before and after tail vein administration of SPIO-loaded PCPZL2 micelles. (<b>C</b>) Mouse T<sub>2</sub>-weighted images of mice at different time points before and after administration of SPIO-loaded micelles.</p>
Full article ">Figure 6
<p>In vivo T<sub>2</sub>-weighted MRI of nude mice bearing human hepatocellular carcinoma. (<b>A</b>) T<sub>2</sub>-weighted images of mice at different time points before and after tail vein administration of SPIO-loaded PCPZL2 micelles. The blue, green, and red circles indicate the regions of interest for the tumor, liver, and muscle, respectively, selected for the signal intensity measurement. (<b>B</b>) Mean values of signal intensity collected before and after micelle administration. (<b>C</b>) Time-dependence of the relative signal difference between tumor and normal liver tissues at different time points.</p>
Full article ">Scheme 1
<p>Schematic illustration of the structural differences between SPIO-loaded micelles based on the PCPZL and conventional linear amphiphilic copolymers to diagnose liver disease or cancer.</p>
Full article ">
25 pages, 6341 KiB  
Article
Fast Self-Healing at Room Temperature in Diels–Alder Elastomers
by Ali Safaei, Joost Brancart, Zhanwei Wang, Sogol Yazdani, Bram Vanderborght, Guy Van Assche and Seppe Terryn
Polymers 2023, 15(17), 3527; https://doi.org/10.3390/polym15173527 - 24 Aug 2023
Cited by 6 | Viewed by 2061
Abstract
Despite being primarily categorized as non-autonomous self-healing polymers, we demonstrate the ability of Diels–Alder polymers to heal macroscopic damages at room temperature, resulting in complete restoration of their mechanical properties within a few hours. Moreover, we observe immediate partial recovery, occurring mere minutes [...] Read more.
Despite being primarily categorized as non-autonomous self-healing polymers, we demonstrate the ability of Diels–Alder polymers to heal macroscopic damages at room temperature, resulting in complete restoration of their mechanical properties within a few hours. Moreover, we observe immediate partial recovery, occurring mere minutes after reuniting the fractured surfaces. This fast room-temperature healing is accomplished by employing an off-stoichiometric maleimide-to-furan ratio in the polymer network. Through an extensive investigation of seven Diels–Alder polymers, the influence of crosslink density on self-healing, thermal, and (thermo-)mechanical performance was thoroughly examined. Crosslink density variations were achieved by adjusting the molecular weight of the monomers or utilizing the off-stoichiometric maleimide-to-furan ratio. Quasistatic tensile testing, dynamic mechanical analysis, dynamic rheometry, differential scanning calorimetry, and thermogravimetric analysis were employed to evaluate the individual effects of these parameters on material performance. While lowering the crosslink density in the polymer network via decreasing the off-stoichiometric ratio demonstrated the greatest acceleration of healing, it also led to a slight decrease in (dynamic) mechanical performance. On the other hand, reducing crosslink density using longer monomers resulted in faster healing, albeit to a lesser extent, while maintaining the (dynamic) mechanical performance. Full article
(This article belongs to the Section Smart and Functional Polymers)
Show Figures

Figure 1

Figure 1
<p>Qualitative illustration of the fast room temperature self-healing in a DPBM-FT5000-r0.5 sample: (<b>a</b>) A cylindrical sample with a diameter of 10 mm and length of 60 mm (<b>b</b>–<b>d</b>) is cut in half using a razor blade. (<b>e</b>) The fracture surfaces are manually reconnected within seconds. (<b>f</b>) The reconnected sample is left to heal at room temperature (25 °C) for only 3 min. (<b>g</b>) Subsequently, it is positioned vertically. (<b>h</b>) After 3 min of self-healing at ambient conditions, the sample can hold a weight of 200 g. The video can be found at the following link: <a href="https://www.youtube.com/watch?v=a-bCetuPKjs" target="_blank">https://www.youtube.com/watch?v=a-bCetuPKjs</a> (accessible since 20 July 2023).</p>
Full article ">Figure 2
<p>Monomers of the Diels–Alder polymer networks: (<b>a</b>) The furfuryl glycidyl ether (FGE). (<b>b</b>) Chemical structure of the diamine Jeffamines D4000 (x = 49.0) and D2000. (<b>c</b>) Chemical structure of the triamine Jeffamines T5000 (x + y + z = 96.8, n = 0) and T3000 (x + y + z = 54.2, n = 0). (<b>d</b>) The 1,1′-(methylenedi-4,1-phenylene) bismaleimide (DPBM). (<b>e</b>) Hydroquinone (1,4-benzenediol).</p>
Full article ">Figure 3
<p>The Diels–Alder crosslink density (<span class="html-italic">[DA]<sub>eq</sub></span>) can be varied using two parameters that were selected in this research: (<b>a</b>) By decreasing the molar mass of the furan-containing compound in the synthesis (<span class="html-italic">M<sub>Jeffamine</sub></span>), the crosslink density of the resulting network is reduced. (<b>b</b>) By decreasing the off-stoichiometric maleimide-to-furan ratio (<span class="html-italic">r</span>), a larger excess of furan is created, resulting in a decrease in crosslink density in the resulting network.</p>
Full article ">Figure 4
<p>Preparation and testing of the healed tensile samples: (<b>a</b>) 60 × 30 × 1 mm<sup>3</sup> sheets are cut in half using a razor blade. Immediately after cutting, the fracture surfaces are brought back into contact. (<b>b</b>) The rectangular 5 × 30 × 1 mm<sup>3</sup> are cut. (<b>c</b>) The tensile samples are left to heal at room temperature. (<b>d</b>) Upon fracturing a sample that was not completely healed due to insufficient healing time, the sample fractures at the location of the scar. (<b>e</b>) Given enough time at room temperature, the fracture occurs at a location that is different from that of the scar. This illustrates full recovery of the mechanical properties. (<b>f</b>–<b>h</b>) Microscopic images of the samples prior to damage (<b>f</b>), after damage (<b>g</b>), and seconds after recontact (<b>h</b>).</p>
Full article ">Figure 5
<p>Uniaxial tensile test until fracture with a fixed strain ramp of 1%.s<sup>−1</sup> for the triamine-based elastomers (<b>a</b>) and the diamine-based elastomers (<b>b</b>). In (<b>c</b>), the resulting toughness’s are presented, which were calculated by integrating the stress–strain curve in the total strain window until fracture.</p>
Full article ">Figure 6
<p>Healing tests based on uniaxial tensile testing of healed samples and reaction simulations for the triamine-based samples: (<b>a</b>,<b>c</b>) Evolution of the stress as function of strain during the uniaxial tensile test with fixed strain ramp of 1%.s<sup>−1</sup>. The pristine, undamaged samples are compared with samples that were healed at different times. (<b>b</b>,<b>d</b>) The resulting healing efficiencies are based on the recovery of Young’s modulus, fracture stress, fracture strain, and toughness. (<b>e</b>,<b>f</b>) Comparison of the increase in healing efficiency (<b>e</b>) and fracture stress (<b>f</b>) as a function of time for the two triamine-based networks with different crosslink densities. (<b>g</b>,<b>h</b>) Comparison of the increase in simulated conversion <span class="html-italic">x</span> (<b>g</b>), Diels–Alder concentration <span class="html-italic">[DA],</span> and maleimide concentration <span class="html-italic">[M]</span> (<b>h</b>) as a function of time for the two triamine-based networks with different crosslink densities.</p>
Full article ">Figure 7
<p>Healing tests based on uniaxial tensile testing of healed samples and reaction simulations for the diamine-based Diels–Alder polymers: (<b>a</b>,<b>c</b>) Evolution of the stress as function of strain during the uniaxial tensile test with fixed strain ramp of 1%.s<sup>−1</sup>. The pristine, undamaged samples are compared with samples that were healed at different times. (<b>b</b>,<b>d</b>) The resulting healing efficiencies based on the recovery of Young’s modulus, fracture stress, fracture strain and toughness. (<b>e</b>,<b>f</b>) Comparison of the increase in healing efficiency (<b>e</b>) and fracture stress (<b>f</b>) as a function of time for the two diamine-based networks with different crosslink densities. (<b>g</b>,<b>h</b>) Comparison of the increase in simulated conversion <span class="html-italic">x</span> (<b>g</b>) and Diels–Alder concentration <span class="html-italic">[DA]</span> (<b>f</b>) as a function of time for the two diamine-based networks with different crosslink densities.</p>
Full article ">Figure 8
<p>Visco-elastic behavior as a function of temperature of the triamine-based Diels–Alder polymers: (<b>a</b>–<b>c</b>) Dynamic mechanical analysis (DMA) as a function of temperature, presented by the storage modulus (<b>a</b>), loss modulus (<b>b</b>), and Tan delta (<b>c</b>). (<b>d</b>–<b>f</b>) Dynamic rheometry as a function of temperature, presented by the storage modulus (<b>d</b>), loss modulus (<b>e</b>), and phase angle (<b>f</b>). To illustrate the degelation temperature, the loss angle as a function of temperature is plotted for different frequencies 3.1 Hz, 1.8 Hz, 1 Hz, 0.6 Hz, and 0.3 Hz. (<b>g</b>–<b>j</b>) Simulation of the equilibrium conversion x<sub>eq</sub> (<b>g</b>), the Diels–Alder concentration <span class="html-italic">[DA</span><sub>eq</sub><span class="html-italic">]</span> (<b>h</b>), the maleimide concentration <span class="html-italic">[M</span><sub>eq</sub><span class="html-italic">]</span> (<b>i</b>), and the furan concentration <span class="html-italic">[F</span><sub>eq</sub><span class="html-italic">]</span> as a function of temperature.</p>
Full article ">Figure 9
<p>Visco-elastic behavior as a function of temperature of the diamine-based Diels–Alder polymers: (<b>a</b>–<b>c</b>) Dynamic mechanical analysis (DMA) as a function of temperature, presented by the storage modulus (<b>a</b>), loss modulus (<b>b</b>), and loss angle (<b>c</b>). (<b>d</b>–<b>f</b>) Dynamic rheometry as a function of temperature, presented by the storage modulus (<b>d</b>), loss modulus (<b>e</b>), and loss angle (<b>f</b>). To illustrate the degelation temperature, the loss angle as a function of temperature is plotted for different frequencies 3.1 Hz, 1.8 Hz, 1 Hz, 0.6 Hz, and 0.3 Hz. (<b>g</b>–<b>j</b>) Simulation of the equilibrium conversion x<sub>eq</sub> (<b>g</b>), the Diels–Alder concentration <span class="html-italic">[DA</span><sub>eq</sub><span class="html-italic">]</span> (<b>h</b>), the maleimide concentration <span class="html-italic">[M</span><sub>eq</sub><span class="html-italic">]</span> (<b>i</b>), and the furan concentration <span class="html-italic">[F</span><sub>eq</sub><span class="html-italic">]</span> as a function of temperature.</p>
Full article ">Figure 10
<p>Differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA) for the diamine and triamine-based Diels–Alder polymers: (<b>a</b>) DSC of the triamine-based polymers. (<b>b</b>) DSC of the diamine-based polymers. (<b>c</b>) TGA of the triamine-based polymers. (<b>d</b>) TGA of the diamine-based polymers.</p>
Full article ">Figure 11
<p>(<b>a</b>) Uniaxial tensile test until fracture with a fixed strain ramp of 1%.s<sup>−1</sup> for the triamine-based elastomers with varying maleimide-to-furan ratio <span class="html-italic">r</span>. In (<b>b</b>), the resulting toughness’s are presented, which were calculated by integrating the stress–strain curve in the total strain window until fracture.</p>
Full article ">Figure 12
<p>Healing tests based on uniaxial tensile testing of healed samples and reaction simulations for the triamine-based samples: (<b>a</b>,<b>c</b>,<b>e</b>) Evolution of the stress as a function of strain during the uniaxial tensile test with fixed strain ramp of 1%.s<sup>−1</sup>. The pristine, undamaged samples are compared with samples that were healed at different times. (<b>b</b>,<b>d</b>,<b>f</b>) The resulting healing efficiencies are based on the recovery of Young’s modulus, fracture stress, fracture strain, and toughness. (<b>g</b>,<b>h</b>) Comparison of the increase in healing efficiency (<b>g</b>) and fracture stress (<b>h</b>) as a function of time for the two triamine-based networks with different crosslink densities. (<b>i</b>,<b>j</b>) Comparison of the increase in simulated conversion <span class="html-italic">x</span> (<b>i)</b> and Diels–Alder concentration <span class="html-italic">[DA]</span> (<b>j</b>) as a function of time for the two triamine-based networks with different crosslink densities.</p>
Full article ">Figure 13
<p>Visco-elastic behavior as a function of temperature of the triamine-based Diels–Alder polymers with varying <span class="html-italic">r</span>-ratio: (<b>a</b>–<b>c</b>) Dynamic mechanical analysis (DMA) as a function of temperature, presented by the storage modulus (<b>a</b>), loss modulus (<b>b</b>), and loss angle (<b>c</b>). (<b>d</b>–<b>f</b>) Dynamic rheometry as a function of temperature, presented by the storage modulus (<b>d</b>), loss modulus (<b>e</b>), and loss angle (<b>f</b>). To illustrate the degelation temperature, the loss angle as a function of temperature is plotted for different frequencies 3.1 Hz, 1.8 Hz, 1 Hz, 0.6 Hz, and 0.3 Hz. (<b>g</b>–<b>j</b>) Simulation of the equilibrium conversion <span class="html-italic">x<sub>eq</sub></span> (<b>g</b>), the Diels–Alder concentration <span class="html-italic">[DA</span><sub>eq</sub><span class="html-italic">]</span> (<b>h</b>), the maleimide concentration <span class="html-italic">[M</span><sub>eq</sub><span class="html-italic">]</span> (<b>i</b>), and the furan concentration <span class="html-italic">[F</span><sub>eq</sub><span class="html-italic">]</span> as a function of temperature.</p>
Full article ">Figure 14
<p>Thermal analysis for the triamine-based Diels–Alder polymers with varying <span class="html-italic">r</span>-ratio. (<b>a</b>) Differential scanning calorimetry (DSC). (<b>b</b>) Thermogravimetric analysis (TGA).</p>
Full article ">
16 pages, 6073 KiB  
Article
Facile Synthesis of Polymer-Reinforced Silica Aerogel Microspheres as Robust, Hydrophobic and Recyclable Sorbents for Oil Removal from Water
by Zhiyang Zhao, Jian Ren, Wei Liu, Wenqian Yan, Kunmeng Zhu, Yong Kong, Xing Jiang and Xiaodong Shen
Polymers 2023, 15(17), 3526; https://doi.org/10.3390/polym15173526 - 24 Aug 2023
Cited by 5 | Viewed by 1259
Abstract
With the rapid development of industry and the acceleration of urbanization, oil pollution has caused serious damage to water, and its treatment has always been a research hotspot. Compared with traditional adsorption materials, aerogel has the advantages of light weight, large adsorption capacity [...] Read more.
With the rapid development of industry and the acceleration of urbanization, oil pollution has caused serious damage to water, and its treatment has always been a research hotspot. Compared with traditional adsorption materials, aerogel has the advantages of light weight, large adsorption capacity and high selective adsorption, features that render it ideal as a high-performance sorbent for water treatment. The objective of this research was to develop novel hydrophobic polymer-reinforced silica aerogel microspheres (RSAMs) with water glass as the precursor, aminopropyltriethoxysilane as the modifier, and styrene as the crosslinker for oil removal from water. The effects of drying method and polymerization time on the structure and oil adsorption capacity were investigated. The drying method influenced the microstructure and pore structure in a noteworthy manner, and it also significantly depended on the polymerization time. More crosslinking time led to more volume shrinkage, thus resulting in a larger apparent density, lower pore volume, narrower pore size distribution and more compact network. Notably, the hydrophobicity increased with the increase in crosslinking time. After polymerization for 24 h, the RSAMs possessed the highest water contact angle of 126°. Owing to their excellent hydrophobicity, the RSAMs via supercritical CO2 drying exhibited significant oil and organic liquid adsorption capabilities ranging from 6.3 to 18.6 g/g, higher than their state-of-the-art counterparts. Moreover, their robust mechanical properties ensured excellent reusability and recyclability, allowing for multiple adsorption–desorption cycles without significant degradation in performance. The novel sorbent preparation method is facile and inspiring, and the resulting RSAMs are exceptional in capacity, efficiency, stability and regenerability. Full article
(This article belongs to the Special Issue Status and Progress of Soluble Polymers II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The mechanism of the polymer-reinforced silica aerogel microsphere sol–gel reaction process. (<b>a</b>) Facile preparation process. (<b>b</b>) Structure of pure, amino, and reinforced aerogels. (<b>c</b>) Mechanism of polystyrene crosslinking to form the robust “golden shield”. (<b>d</b>) Typical photos of different stages of the polymerization reaction from ASGMs to RSAMs.</p>
Full article ">Figure 2
<p>Photographs of aerogel samples in water and the water contact angle (WCA) of hydrophobic samples. (<b>a</b>) PSAMs. (<b>b</b>) ASAMs. (<b>c</b>) RSAMs under different crosslinking times via VD method. (<b>d</b>) RSAMs under different crosslinking times via SCD method. (<b>e</b>) WCA of R-V-24. (<b>f</b>) WCA of R-S-24.</p>
Full article ">Figure 3
<p>FTIR spectra of the RSAMs obtained by different drying methods. (<b>a</b>) VD. (<b>b</b>) SCD.</p>
Full article ">Figure 4
<p>TGA curves of the SAMs obtained by different drying methods. (<b>a</b>) VD. (<b>b</b>) SCD.</p>
Full article ">Figure 5
<p>N<sub>2</sub> adsorption/desorption isotherms and pore size distribution curves of RSAMs obtained by different drying methods. (<b>a</b>,<b>b</b>) VD. (<b>c</b>,<b>d</b>) SCD.</p>
Full article ">Figure 6
<p>Microstructure of RSAMs obtained through different drying method under varying crosslinking times. (<b>a</b>–<b>d</b>) SEM images under vacuum drying for (<b>a</b>) 3 h, (<b>b</b>) 6 h, (<b>c</b>) 12 h, and (<b>d</b>) 24 h crosslinking. (<b>e</b>,<b>f</b>) SEM images under supercritical CO<sub>2</sub> drying for (<b>e</b>) 3 h, (<b>f</b>) 6 h, (<b>g</b>) 12 h, and (<b>h</b>) 24 h crosslinking. (<b>i</b>) Schematic diagram of the formation of the polystyrene-reinforced silica aerogel skeleton.</p>
Full article ">Figure 7
<p>Mechanical performance of RSAMs. (<b>a</b>) Test device for spherical materials using 5-point method. (<b>b</b>) Compressive curves of PSAMs and ASAMs. (<b>c</b>) Compressive curves via VD method and the enlarged linear area. (<b>d</b>) Compressive curves via SCD method and the enlarged linear area.</p>
Full article ">Figure 8
<p>Adsorption and recyclable performance of RSAMs. (<b>a</b>) Adsorption capacity for several oils and organic solvents. (<b>b</b>) Repeated adsorption and desorption capacity of RSAMs for pump oil. (<b>c</b>) Organics removal from water: the recycling and regeneration process using RSAMs. Step 1 is the oil–water mixture system, step 2 is the adsorption process, step 3 is the removal process, step 4 is the system after removal, step 5 is the recycled RSAMs, step 6 is the regenerated RSAMs. (<b>d</b>) Regeneration rate of RSAMs for hexane. (<b>e</b>) Comprehensive performance comparison of RSAMs with various adsorbents for oil removal.</p>
Full article ">
30 pages, 12372 KiB  
Article
Hydrophobic, Thermal Shock-and-Corrosion-Resistant XSBR Latex-Modified Lightweight Class G Cement Composites in Geothermal Well Energy Storage Systems
by Toshifumi Sugama and Tatiana Pyatina
Materials 2023, 16(17), 5792; https://doi.org/10.3390/ma16175792 - 24 Aug 2023
Cited by 2 | Viewed by 1070
Abstract
Energy losses can be significantly reduced if thermally insulating cement is used for energy storage and recovery. The thermal conductivity (TC) of the currently used cement is between 1 and 1.2 W/mK. In this study we assessed the ability of polystyrene (PS)–polybutadiene (PB)–polyacrylic [...] Read more.
Energy losses can be significantly reduced if thermally insulating cement is used for energy storage and recovery. The thermal conductivity (TC) of the currently used cement is between 1 and 1.2 W/mK. In this study we assessed the ability of polystyrene (PS)–polybutadiene (PB)–polyacrylic acid (PAA) terpolymer (cross-linked styrene–butadiene rubber, XSBR) latex to improve thermal insulating properties and thermal shock (TS) resistance of class G ordinary Portland cement (OPC) and fly ash cenosphere (FCSs) composites in the temperature range of 100–175 °C. The composites autoclaved at 100 °C were subjected to three cycles, one cycle: 175 °C heat → 25 °C water quenching). In hydrothermal and thermal (TS) environments at elevated temperatures in cement slurries the XSBR latex formed acrylic calcium complexes through acid–base reactions, and the number of such complexes increased at higher temperatures due to the XSBR degradation with formation of additional acrylic groups. As a result, these complexes offered the following five advanced properties to the OPC-based composites: (1) enhanced hydrophobicity; (2) decreased water-fillable porosity; (3) reduced TC for water-saturated composites; (4) minimized loss of compressive strength, Young’s modulus, and compressive fracture toughness after TS; and (5) abated pozzolanic activity of FCSs, which allowed FCSs to persist as thermal insulators under strongly alkaline conditions of cement slurries. Additionally, XSBR-modified slurries possessed improved workability and decreased slurry density due to the air-entraining effect of latex, which resulted in further improvement of thermal insulation performance of the modified composites. Full article
Show Figures

Figure 1

Figure 1
<p>ATR-FTIR absorption spectra for 100 °C-dried XSBR latex, and 100 °C- and 175 °C-autoclaved dried XSBR samples.</p>
Full article ">Figure 2
<p>Hydrothermal degradation and oxidation pathways of PB <span class="html-italic">trans-</span> and <span class="html-italic">cis</span>-1,4, and 1,2 vinyl units to form isolated carboxyl-terminated PB (CTPB), and non-isolated carboxyl-ended PS (CEPS) and carboxyl-ended PAA (CEPAA) as oxidation derivatives.</p>
Full article ">Figure 3
<p>ATR-FTIR spectra for 100 °C- and 175 °C-autoclaved neat OPC.</p>
Full article ">Figure 4
<p>ATR-FTIR spectra of XSBR-modified neat OPC after being autoclaved at 100 °C and 175 °C.</p>
Full article ">Figure 5
<p>Illustration of XSBR molecular transformations in acid–base reactions between carboxyl group (proton donor acid) and Ca<sup>2+</sup> 2OH<sup>−</sup> (proton acceptor base) at hydrothermal temperatures of 100 °C and 175 °C.</p>
Full article ">Figure 6
<p>Illustration of Ca<sup>2+</sup>-coordinated chelate complex configurations derived from PB-hydrothermal degradation/oxidation at 175 °C.</p>
Full article ">Figure 7
<p>XRD patterns of crystalline phases of 175 °C-autoclaved XSBR-modified (blue) and neat OPC (red).</p>
Full article ">Figure 8
<p>TGA-DTG thermal analyses of 100 °C-dried XSBR, and 100 °C- and 175 °C-autoclaved XSBR samples.</p>
Full article ">Figure 9
<p>TGA-DTG thermal analyses of 100 °C- and 175 °C-autoclaved neat OPC samples.</p>
Full article ">Figure 10
<p>TGA-DTG thermal analyses of 100 °C- and 175 °C-autoclaved XSBR-modified OPC samples.</p>
Full article ">Figure 11
<p>Normalized heat flow curves and hydration reaction energies of OPC slurries made with 0, 5, and 15% P/C ratios at 25 °C. Illustration of suppression of ettringite formation by XSBR containing PAA-pendant chelate complexes adsorbed on cement grain surfaces.</p>
Full article ">Figure 12
<p>Normalized heat flow curves and hydration reaction energies of OPC slurries made with 0, 5, 15, and 25% P/C ratios at 85 °C.</p>
Full article ">Figure 13
<p>XRD patterns of 100 °C-autoclaved OPC composites modified (blue) and unmodified (red) with XSBR.</p>
Full article ">Figure 14
<p>XRD patterns of 175 °C-autoclaved OPC composites modified (blue) and unmodified (red) with XSBR.</p>
Full article ">Figure 15
<p>ATR-FTIR spectra of 25% P/C ratio-modified and unmodified OPC composites before and after TS tests.</p>
Full article ">Figure 16
<p>Wetting behavior and water droplet contact angle for 5, 15, and 25% P/C ratio-modified and unmodified OPC composite surfaces before and after TS tests.</p>
Full article ">Figure 17
<p>Water-fillable porosity of 5, 15, and 25% P/C ratio-modified and unmodified composites before and after TS tests.</p>
Full article ">Figure 18
<p>Thermal conductivity of XSBR-modified and unmodified OPC composites before and after TS tests.</p>
Full article ">Figure 19
<p>Changes in compressive strength as a function of P/C ratio, and strength decline rate after TS tests.</p>
Full article ">Figure 20
<p>Changes in Young’s modulus as a function of P/C ratio, and the modulus decline rate after TS tests.</p>
Full article ">Figure 21
<p>Changes in compressive fracture toughness as a function of P/C ratio, and toughness decline rate after TS tests.</p>
Full article ">Figure 22
<p>Comparison of compressive fracture toughness on the compressive stress–strain curves for 0% and 25% P/C ratio OPC composites after TS tests.</p>
Full article ">Figure 23
<p>Comparison of compressive fracture toughness for OPC/FCSs, OPC/FCSs/MGFs, OPC/FCSs/XSBR, and OPC/FCSs/MGFs/XSBR combination systems.</p>
Full article ">Figure 24
<p>Comparison of morphological features between unmodified (<b>top</b>) and XSBR-modified (<b>bottom</b>) OPC composites.</p>
Full article ">Figure 25
<p>Pozzolanic reaction products of FCSs in unmodified composite (<b>top</b>) and non-reacted FCSs in XSBR-modified OPC composite (<b>bottom</b>).</p>
Full article ">Figure 26
<p>DC-electrochemical cathodic–anodic polarization curves for 0, 5, 15, and 25% P/C ratio cement composite-coated CS surfaces and Tafel fit diagram before TS test.</p>
Full article ">Figure 27
<p>DC-electrochemical cathodic–anodic polarization curves for 0, 5, 15, and 25% P/C ratio cement composite-coated CS surfaces after TS tests.</p>
Full article ">Figure 28
<p>Corrosion current as a function of P/C ratio before and after TS tests.</p>
Full article ">Figure 29
<p>Corrosion potential as a function of P/C ratio before and after TS tests.</p>
Full article ">Figure 30
<p>Corrosion rate of CS and thickness of coating layers for 0, 5, 15, and 25% P/C ratio composite coatings before and after TS tests.</p>
Full article ">
Back to TopTop