[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Search Results (3,207)

Search Parameters:
Keywords = viscoelastic

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 68807 KiB  
Article
Structural and Viscoelastic Properties of Bacterial Cellulose Composites: Implications for Prosthetics
by Natalia Pogorelova, Daniil Parshin, Anna Lipovka, Alexey Besov, Ilya Digel and Pyotr Larionov
Polymers 2024, 16(22), 3200; https://doi.org/10.3390/polym16223200 - 18 Nov 2024
Abstract
This study investigates the morphological, mechanical, and viscoelastic properties of bacterial cellulose (BC) hydrogels synthesized by the microbial consortium Medusomyces gisevii. BC gel films were produced under static (S) or bioreactor (BioR) conditions. Additionally, an anisotropic sandwich-like composite BC film was developed [...] Read more.
This study investigates the morphological, mechanical, and viscoelastic properties of bacterial cellulose (BC) hydrogels synthesized by the microbial consortium Medusomyces gisevii. BC gel films were produced under static (S) or bioreactor (BioR) conditions. Additionally, an anisotropic sandwich-like composite BC film was developed and tested, consisting of a rehydrated (S-RDH) BC film synthesized under static conditions, placed between two BioR-derived BC layers. Sample characterization was performed using scanning electron microscopy (SEM), atomic force microscopy (AFM), rheometry, and uniaxial stretching tests. To our knowledge, this is the first study to combine uniaxial and rheological tests for BC gels. AFM and SEM revealed that the organization of BC fibrils (80±20 nm in diameter) was similar to that of collagen fibers (96±31 nm) found in human dura mater, suggesting potential implications for neurosurgical practice. Stretching tests demonstrated that the drying and rehydration of BC films resulted in a 2- to 8-fold increase in rigidity compared to other samples. This trend was consistent across both small and large deformations, regardless of direction. Mechanically, the composite (BioR+S-RDH) outperformed BC hydrogels synthesized under static and bioreactor conditions by approx. 26%. The composite material (BioR+S-RDH) exhibited greater anisotropy in the stretching tests compared to S-RDH, but less than the BioR-derived hydrogels, which had anisotropy coefficients ranging from 1.29 to 2.03. BioR+S-RDH also demonstrated the most consistent viscoelastic behavior, indicating its suitability for withstanding shear stress and potential use in prosthetic applications. These findings should provide opportunities for further research and medical applications. Full article
(This article belongs to the Special Issue Advances in Cellulose-Based Polymers and Composites, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic set-up used for the measurement of the viscoelastic characteristics of the samples (<b>A</b>); installation of a BC hydrogel sample prior to testing on rheometer (<b>B</b>).</p>
Full article ">Figure 2
<p>Sample during uniaxial stretching.</p>
Full article ">Figure 3
<p>Visualization of the differences between the sides of the BC(S) film and its cross section: SEM images of the BC film synthesized under static conditions (<b>A</b>); the denser side of the surface is that formed at the air–nutrient interface. (<b>B</b>) The porous opposite side. (<b>C</b>,<b>D</b>) Layered structure seen in the cross section of the sample.</p>
Full article ">Figure 4
<p>SEM images of lyophilized BC(BioR) film: (<b>A</b>,<b>B</b>) porous structure of a section of some samples.</p>
Full article ">Figure 5
<p>AFM photographs of the surface of BC samples synthesized under static (<b>top row</b>) and reactor (<b>bottom row</b>) conditions: 2D image (<b>top left</b>, <b>bottom left</b>); 3D image (<b>top center</b>, <b>bottom center</b>); and differential signal (<b>top right</b>, <b>bottom right</b>), respectively.</p>
Full article ">Figure 6
<p>AFM photographs of the surface of a lyophilized BC gel film synthesized under reactor conditions: 2D image (<b>A</b>,<b>B</b>); 3D image (<b>C</b>,<b>D</b>); and differential signal (<b>E</b>,<b>F</b>), respectively. Arrows indicate branching points of BC fibrils. Scan size: 5 × 5 μm (<b>left</b>); 3 × 3 μm (<b>right</b>).</p>
Full article ">Figure 7
<p>Overlapping strain–stress diagrams of uniaxial strength tests (<b>A</b>) and samples’ corresponding ultimate stress and ultimate strain values (<b>B</b>).</p>
Full article ">Figure 8
<p>Amplitude sweep data of tested BC hydrogel samples. Squared line—<math display="inline"><semantics> <mrow> <msup> <mi>G</mi> <mo>′</mo> </msup> </mrow> </semantics></math>. Triangled line—<math display="inline"><semantics> <mrow> <msup> <mi>G</mi> <mo>″</mo> </msup> </mrow> </semantics></math> data (Pa). Horizontal axis—shear strain (%).</p>
Full article ">Figure 9
<p>Frequency sweep data of tested hydrogel samples. Squared line—<math display="inline"><semantics> <mrow> <msup> <mi>G</mi> <mo>′</mo> </msup> </mrow> </semantics></math> (storage modulus). triangled line—<math display="inline"><semantics> <mrow> <msup> <mi>G</mi> <mo>″</mo> </msup> </mrow> </semantics></math> (loss modulus) data (Pa). Horizontal axis—angular frequency (1/s).</p>
Full article ">Figure A1
<p>Visualization of BC composite rupture (BioR+S-RDH) in air: (<b>A</b>) test plan, (<b>B</b>) onset of rupture, and (<b>C</b>) the rupture of the middle layer occurs in the central part of the sample.</p>
Full article ">Figure A2
<p>Full stress–strain diagrams of tensile tests.</p>
Full article ">
15 pages, 7304 KiB  
Article
Preparation and Characteristics of Ball-Milled Blueberry Peel Particles and Their Application in Ice Cream
by Li-Hua Pan, Jia-Hui Lin, Mei-Jia Li, Lei Cao, Xiao-Yu Liu, Yuan-Yuan Deng, Shui-Zhong Luo and Zhi Zheng
Foods 2024, 13(22), 3660; https://doi.org/10.3390/foods13223660 (registering DOI) - 17 Nov 2024
Viewed by 282
Abstract
Ice cream is popular but contains high amounts of saturated fats and few health-promoting ingredients. In the presence of xanthan gum (0.25%), blueberry peel particles prepared through ball-milling treatment (BMPs) were used to prepare ice cream containing camellia oil as a fat replacer. [...] Read more.
Ice cream is popular but contains high amounts of saturated fats and few health-promoting ingredients. In the presence of xanthan gum (0.25%), blueberry peel particles prepared through ball-milling treatment (BMPs) were used to prepare ice cream containing camellia oil as a fat replacer. The BMPs possessed smaller particle sizes, larger contact angles, and higher contents of anthocyanin aglycone compared with commonly milled blueberry peel particles. BMPs with the largest contact angle (66.30°) were obtained by ball-milling the blueberry peel at 15 Hz for 6 h (BMP15Hz6h). The ice cream mixes were depicted as linear viscoelastic gel-like solids, and their apparent viscosity, G′ and G′, increased with the increase in the BMP15Hz6h concentration. Ice cream with strong antioxidant activity and good freeze–thaw stability was acceptable and desirable in the presence of 0.5% BMP15Hz6h. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of ball milling frequency and ball milling time on contact angle (<b>A</b>) and microstructure (<b>B</b>) of blueberry peel particles.</p>
Full article ">Figure 2
<p>Rheological properties of ice cream mixes with different levels of BMP<sub>15Hz1.5h</sub> added. (<b>A</b>), G’, storage modulus; (<b>B</b>), apparent viscosity.</p>
Full article ">Figure 3
<p>Antioxidant activities (<b>A</b>), storage physical stability (<b>B</b>), and sensory scores (<b>C</b>,<b>D</b>) of ice cream prepared with different levels of BMP<sub>15Hz1.5h</sub>. Means with different lowercase letters in the same row are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Visuals (<b>A</b>), optical microscopy images (<b>B</b>), and confocal micrographs (<b>C</b>) of ice cream prepared with different levels of BMP<sub>15Hz1.5h</sub>. (Red and green in (<b>C</b>) represent fat globules and protein granules, respectively.)</p>
Full article ">
22 pages, 9843 KiB  
Article
Viscoelastic Polyurethane Foam Biocomposites with Enhanced Flame Retardancy
by Grzegorz Węgrzyk, Dominik Grzęda, Milena Leszczyńska, Bartosz Nędza, Katarzyna Bulanda, Mariusz Oleksy, Joanna Ryszkowska and Ugis Cabulis
Polymers 2024, 16(22), 3189; https://doi.org/10.3390/polym16223189 - 16 Nov 2024
Viewed by 510
Abstract
The growing demand for viscoelastic polyurethane foams creates a need for new sustainable raw materials that support cost-effective production while maintaining the desired material performance and fire safety standards. In this regard, our study aimed to develop viscoelastic polyurethane foam composites with reduced [...] Read more.
The growing demand for viscoelastic polyurethane foams creates a need for new sustainable raw materials that support cost-effective production while maintaining the desired material performance and fire safety standards. In this regard, our study aimed to develop viscoelastic polyurethane foam composites with reduced flammability and a high proportion of renewable raw materials. To achieve this, blackcurrant pomace, expandable graphite and a third-generation blowing agent were introduced to a viscoelastic polyurethane foam composition containing a reactive flame retardant in the formulation. The effects of the incorporated additives on the foaming process, flammability, chemical structure, cellular structure, thermal properties and physico-mechanical properties of the composites were determined. The results showed that the viscoelastic foam composite containing 30 php of blackcurrant pomace and 15 php of expandable graphite had a pHRRmax 52% lower than that of the reference material. The additional use of a blowing agent enhanced the flame-retardant effect of the materials, resulting in a 67% reduction in pHRRmax of the composite compared to the reference material. Moreover, the developed biocomposites exhibited promising limiting oxygen index values of 26–28%, compared to the 21% shown for the reference sample. Consequently, the best-performing biocomposites achieved the V-0 flammability rating according to the UL-94 standard. This study’s results indicate the composites’ high application potential due to their reduced flammability and the materials’ desirable physical and mechanical properties. Full article
(This article belongs to the Special Issue Advances in Fire-Safe Polymer Materials)
Show Figures

Figure 1

Figure 1
<p>SEM images of the fillers and foams.</p>
Full article ">Figure 2
<p>Pore size distribution in foams.</p>
Full article ">Figure 3
<p>TG and DTG thermograms of blackcurrant pomace.</p>
Full article ">Figure 4
<p>TG and DTG thermograms of expandable graphite.</p>
Full article ">Figure 5
<p>TG and DTG thermograms of foams.</p>
Full article ">Figure 6
<p>Representative heat release rate curves obtained for analysed materials.</p>
Full article ">Figure 7
<p>Images after conducting UL-94 and TGA test.</p>
Full article ">Figure 8
<p>DSC curves (C1) of the developed materials.</p>
Full article ">Figure 9
<p>DSC curves (C2) of the developed materials.</p>
Full article ">Figure 10
<p>Spectra FTIR of blackcurrant pomace.</p>
Full article ">Figure 11
<p>Spectra FTIR of SOLKANE<sup>®</sup> 365/227.</p>
Full article ">Figure 12
<p>FTIR spectra of the analysed polyurethane materials.</p>
Full article ">Figure 13
<p>Strain–compression correlation charts for the analysed foams.</p>
Full article ">
19 pages, 6697 KiB  
Article
Enhanced Rheological and Structural Properties of the Exopolysaccharide from Rhizobium leguminosarum VF39 Through NTG Mutagenesis
by Kyungho Kim, Eunkyung Oh, Sohyun Park, Jae-pil Jeong, Sobin Jeon, Sujin Lee, Younghyun Shin and Seunho Jung
Polymers 2024, 16(22), 3179; https://doi.org/10.3390/polym16223179 - 15 Nov 2024
Viewed by 337
Abstract
Microbial exopolysaccharides (EPSs) are biopolymer materials with advantages such as biodegradability, biocompatibility, ease of mass production, and reproducibility. The EPS that was isolated from Rhizobium leguminosarum bv. viciae VF39 is an anionic polysaccharide with a backbone structure consisting of one galactose, five glucose [...] Read more.
Microbial exopolysaccharides (EPSs) are biopolymer materials with advantages such as biodegradability, biocompatibility, ease of mass production, and reproducibility. The EPS that was isolated from Rhizobium leguminosarum bv. viciae VF39 is an anionic polysaccharide with a backbone structure consisting of one galactose, five glucose molecules, and two glucuronic acids, along with 3-hydroxybutanoyl, acetyl, and pyruvyl functional groups. Through N-methyl-N′-nitro-N-nitrosoguanidine (NTG) mutagenesis, we isolated and purified a mutant EPS from VF39, VF39 #54, which demonstrated enhanced physicochemical and rheological properties compared to the wild-type VF39. The EPS structure of the VF39 #54 mutant strain showed a loss of glucuronic acid and 3-hydroxybutanoyl groups compared to the wild-type, as confirmed by FT-IR, NMR analysis, and uronic acid assays. The molecular weight of the VF39 #54 EPS was 250% higher than that of the wild-type. It also exhibited improved viscoelasticity and thermal stability. In the DSC and TGA analyses, VF39 #54 had a higher endothermic peak (172 °C) compared to the wild-type (142 °C), and its thermal decomposition point was 260 °C, surpassing the wild-type’s value of 222 °C. Additionally, the VF39 #54 EPS maintained a similar viscosity to the wild-type in various pH, temperature, and metal salt conditions, while also exhibiting a higher overall viscosity. The cytotoxicity test using HEK-293 cells confirmed that the VF39 #54 EPS was non-toxic. Due to its high viscoelastic properties, the VF39 #54 EPS shows potential for use in products such as thickeners, texture enhancers, and stabilizers. Furthermore, its thermal stability and biocompatibility make it a promising candidate for applications in food, pharmaceuticals, and cosmetic formulations. Additionally, its ability to maintain viscosity under varying environmental conditions highlights its suitability for industrial processes that require consistent performance. Full article
(This article belongs to the Section Biobased and Biodegradable Polymers)
Show Figures

Figure 1

Figure 1
<p>Scheme of VF39 NTG mutagenesis. NTG was dissolved in distilled water with 10% acetone for activation.</p>
Full article ">Figure 2
<p>Cultured mutant strain colonies from the VF39 wild-type after NTG mutagenesis.</p>
Full article ">Figure 3
<p>Structure of the VF39 EPS (<b>a</b>) and #54 EPS (<b>b</b>). The attached functional groups, 3-hydroxybutanoyl, pyruvate, and acetyl, are highlighted with red circles.</p>
Full article ">Figure 4
<p>FT-IR spectra of the VF39 EPS and VF39 #54 EPS (<b>a</b>) and expanded spectra of wavelength 2000~600 cm<sup>−1</sup> (<b>b</b>).</p>
Full article ">Figure 5
<p>The <sup>1</sup>H NMR of the VF39 and #54 EPSs. The <sup>1</sup>H NMR spectra was measured using a 10 mg/mL concentration of sample solutions dissolved with D<sub>2</sub>O (99%) at 60 °C.</p>
Full article ">Figure 6
<p><sup>13</sup>C NMR methyl peak of the VF39 and VF39 #54 EPSs. The methyl peaks of acetyl (2,2′), pyruvate (3,3′), and 3-hydroxybutanoyl (4) are shown in the VF39 EPS spectrum. The VF39 #54 EPS shows only the acetyl (2) and pyruvyl (3) peaks.</p>
Full article ">Figure 7
<p>HSQC functional group spectra of the VF39 and VF39 #54 EPSs with D<sub>2</sub>O (99%). Peaks of NMR data revealed 3-hydroxylbutanoyl (1,2), pyruvyl (3,3′), and acetyl (4,4′) groups of EPS samples.</p>
Full article ">Figure 8
<p>Glucuronic acid concentration assay with the <span class="html-italic">m</span>-hydroxydiphenyl method. No glucuronic acid was detected in the VF39 #54 EPS.</p>
Full article ">Figure 9
<p>(<b>a</b>) TGA and (<b>b</b>) DSC curves of the VF39 and VF39 #54 EPSs.</p>
Full article ">Figure 10
<p>Viscosity of the VF39 wild-type and #54 EPSs with a 2% concentration of solution. Measurements were performed at 25 °C.</p>
Full article ">Figure 11
<p>Viscosity of the VF39 wild-type and #54 with different concentrations (<b>a</b>), temperature (<b>b</b>), pH (<b>c</b>), and metal salts (<b>d</b>).</p>
Full article ">Figure 12
<p>Rheological measurements of (<b>a</b>) the VF39 wild-type EPS and (<b>b</b>) the VF39 #54 EPS with heating and cooling from 25 °C to 75 °C at 2 wt% concentration with shear rate of 10 s<sup>−1</sup>.</p>
Full article ">Figure 13
<p>Rheological measurements with (<b>a</b>) frequency sweep test and (<b>b</b>) amplitude sweep test with a 2 wt% concentration at 25 °C.</p>
Full article ">Figure 14
<p>Gelation of the VF39 and VF39 #54 EPSs with metal chelation.</p>
Full article ">Figure 15
<p>Antioxidant activity of the VF39 EPS samples with DPPH radical scavenging activity test.</p>
Full article ">Figure 16
<p>Cell cytotoxicity test of the VF39 EPS samples with HEK293 cells. DMSO was used as a negative control.</p>
Full article ">
34 pages, 3272 KiB  
Review
Analysis of Tire-Road Interaction: A Literature Review
by Haniyeh Fathi, Zeinab El-Sayegh, Jing Ren and Moustafa El-Gindy
Machines 2024, 12(11), 812; https://doi.org/10.3390/machines12110812 - 14 Nov 2024
Viewed by 364
Abstract
This paper presents a comprehensive literature review of the most popular and recent work on passenger and truck tires. Previous papers discuss a huge amount of work on the modeling of passenger car tires using finite element analysis. In addition, recent works on [...] Read more.
This paper presents a comprehensive literature review of the most popular and recent work on passenger and truck tires. Previous papers discuss a huge amount of work on the modeling of passenger car tires using finite element analysis. In addition, recent works on tire–road interaction and the validation of tires using experimental measurements have been described. Moreover, the history of the tire-road contact algorithms is explained. In addition, friction modeling that is implemented in tire–road interaction applications are discussed. Also, a summary of current state-of-the-art research work definitions and requirements of the tread rubber compound are covered from previous studies using various literature reviews and hyper-viscoelastic material models that are implemented for the tread top and the tread base rubber compound. Furthermore, the effect of tire temperature from previous works is presented here. Finally, this literature review also highlights the shortcomings of recent research work and describes the areas lacking in the literature. Full article
(This article belongs to the Section Vehicle Engineering)
Show Figures

Figure 1

Figure 1
<p>Tire Axis Terminology (<b>a</b>) SAE standard and (<b>b</b>) ISO standard [<a href="#B2-machines-12-00812" class="html-bibr">2</a>].</p>
Full article ">Figure 2
<p>Two-dimensional passenger car finite element model [<a href="#B11-machines-12-00812" class="html-bibr">11</a>].</p>
Full article ">Figure 3
<p>One-wheel system with (<b>a</b>) lumped friction, and (<b>b</b>) distributed friction [<a href="#B31-machines-12-00812" class="html-bibr">31</a>].</p>
Full article ">Figure 4
<p>A physical-based Radial- Inter radial model of the tire in the Spring-Damper forms [<a href="#B35-machines-12-00812" class="html-bibr">35</a>].</p>
Full article ">Figure 5
<p>Brush tire model, stretching of the small volumes of rubber in the tire contact patch [<a href="#B37-machines-12-00812" class="html-bibr">37</a>].</p>
Full article ">Figure 6
<p>In-plane and out-of-plane rigid ring tire models on hard surface [<a href="#B38-machines-12-00812" class="html-bibr">38</a>].</p>
Full article ">Figure 7
<p>Distribution of the friction coefficient in tire contact patch: (<b>a</b>) Deployed of Savkoor’s model and Effect of temperature-dependent material properties for tire compound on tire footprint; (<b>b</b>) Modeling without thermal coupling; (<b>c</b>) Modeling with thermo-mechanical coupling [<a href="#B40-machines-12-00812" class="html-bibr">40</a>].</p>
Full article ">Figure 8
<p>Effect of internal temperature on the rolling resistance coefficient [<a href="#B2-machines-12-00812" class="html-bibr">2</a>].</p>
Full article ">Figure 9
<p>Variation in the rolling resistance coefficient with shoulder temperature for a car tire [<a href="#B2-machines-12-00812" class="html-bibr">2</a>].</p>
Full article ">Figure 10
<p>Temperature distributions of a patterned tire model at 140 km/h [<a href="#B59-machines-12-00812" class="html-bibr">59</a>].</p>
Full article ">Figure 11
<p>Rolling resistance versus tire shoulder temperature at various speeds [<a href="#B66-machines-12-00812" class="html-bibr">66</a>].</p>
Full article ">
17 pages, 9244 KiB  
Article
Accuracy of Dynamic Modulus Models of Asphalt Mixtures Containing Reclaimed Asphalt (RA)
by Majda Belhaj, Jan Valentin and Nicola Baldo
Appl. Sci. 2024, 14(22), 10505; https://doi.org/10.3390/app142210505 - 14 Nov 2024
Viewed by 326
Abstract
The dynamic modulus (∣E*∣) is a fundamental mechanical parameter for studying the performance of hot mix asphalt and simulating its viscoelastic behaviour under different loading and thermal conditions. It is a primary tool to replicate road surface behaviour under vehicle [...] Read more.
The dynamic modulus (∣E*∣) is a fundamental mechanical parameter for studying the performance of hot mix asphalt and simulating its viscoelastic behaviour under different loading and thermal conditions. It is a primary tool to replicate road surface behaviour under vehicle traffic loading and temperature variations. Though, laboratory testing to determine this parameter is time-consuming and costly. Several predictive models have been developed to estimate the dynamic modulus, ranging from rheological to empirical regression models. This research was dedicated to studying two predictive models for determining the master curve of the dynamic modulus of hot mix asphalt used in a regular pavement binder course containing different reclaimed asphalt contents (0%, 30%, 40%, and 50%). Laboratory experiments were conducted to assess their accuracy. The results show that Witczak’s sigmoid function provided the best accuracy for the master curves, while the Generalized Huet-Sayegh (2S2P1D) model showed less accurate predictions, particularly at the range of low and high frequencies. Full article
(This article belongs to the Special Issue Rheology of Binders and Asphalt Mixtures)
Show Figures

Figure 1

Figure 1
<p>Sinusoidal stress (<math display="inline"><semantics> <mrow> <mi>σ</mi> </mrow> </semantics></math>) and strain (<math display="inline"><semantics> <mrow> <mi>ε</mi> </mrow> </semantics></math>) waveform in time (<math display="inline"><semantics> <mrow> <mi>σ</mi> </mrow> </semantics></math> is the stress, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> is the amplitude of the stress, <math display="inline"><semantics> <mrow> <mi>ε</mi> </mrow> </semantics></math> is the strain, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> is the amplitude of the strain, <math display="inline"><semantics> <mrow> <mi>ω</mi> </mrow> </semantics></math> is the angular frequency of the oscillation, t is the time variable, and <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math> is the phase angle).</p>
Full article ">Figure 2
<p>Graphic representation of the dynamic modulus components including the phase angle Φ.</p>
Full article ">Figure 3
<p>Parameterization of the sigmoid function [<a href="#B18-applsci-14-10505" class="html-bibr">18</a>].</p>
Full article ">Figure 4
<p>Generalized Huet–Sayegh rheological model (2S2P1D).</p>
Full article ">Figure 5
<p>Aggregate gradation of the asphalt mixture (red dots represent ACL aggregate gradation limits according to ČSN 73 6121:2019 [<a href="#B26-applsci-14-10505" class="html-bibr">26</a>].</p>
Full article ">Figure 6
<p>Isothermal curves of asphalt mixtures with (<b>a</b>) 0% RA, (<b>b</b>) 30% RA, (<b>c</b>) 40% RA, and (<b>d</b>) 50% RA.</p>
Full article ">Figure 7
<p>Master curves of asphalt mixtures with 0% RA, 30% RA, 40% RA, and 50% RA.</p>
Full article ">Figure 8
<p>Comparison between the master curves of the experimental data and those of the sigmoid function of the asphalt mixture with 0% RA at <span class="html-italic">T<sub>R</sub></span> = 20 °C.</p>
Full article ">Figure 9
<p>Comparison between the master curves of the experimental data and those of the sigmoid function of the asphalt mixture with 30% RA at <span class="html-italic">T<sub>R</sub></span> = 20 °C.</p>
Full article ">Figure 10
<p>Comparison between the master curves of experimental data and those of the sigmoid function of the asphalt mixture with 40% RA at <span class="html-italic">T<sub>R</sub></span> = 20 °C.</p>
Full article ">Figure 11
<p>Comparison between the master curves of experimental data and those of the sigmoid function of the asphalt mixture with 50% RA at <span class="html-italic">T<sub>R</sub></span> = 20 °C.</p>
Full article ">Figure 12
<p>Comparison between measured |E*|_exp and calculated |E*|_Sigmoid model for the asphalt mixture with 0% RA (the red line represents the identity function y = x).</p>
Full article ">Figure 13
<p>Comparison between measured |E*|_exp and calculated |E*|_Sigmoid model for the asphalt mixture with 30% RA (the red line represents the identity function y = x).</p>
Full article ">Figure 14
<p>Comparison between measured |E*|_exp and calculated |E*|_Sigmoid model for the asphalt mixture with 40% RA (the red line represents the identity function y = x).</p>
Full article ">Figure 15
<p>Comparison between measured |E*|_exp and calculated |E*|_Sigmoid model for the asphalt mixture with 50% RA (the red line represents the identity function y = x).</p>
Full article ">Figure 16
<p>Graphical representation of Huet–Sayegh model parameters in (<b>a</b>) Cole–Cole diagram and (<b>b</b>) Black diagram [<a href="#B30-applsci-14-10505" class="html-bibr">30</a>].</p>
Full article ">Figure 17
<p>Comparison between the master curves of the experimental data and those of the Generalized Huet–Sayegh (2S2P1D) model for asphalt mixture with 0% RA at <span class="html-italic">T<sub>R</sub></span> = 20 °C.</p>
Full article ">Figure 18
<p>Comparison between the master curves of the experimental data and those of the Generalized Huet–Sayegh (2S2P1D) model for asphalt mixture with 30% RA at <span class="html-italic">T<sub>R</sub></span> = 20 °C.</p>
Full article ">Figure 19
<p>Comparison between the master curves of the experimental data and those of the Generalized Huet–Sayegh (2S2P1D) model for asphalt mixture with 40% RA at <span class="html-italic">T<sub>R</sub></span> = 20 °C.</p>
Full article ">Figure 20
<p>Comparison between the master curves of the experimental data and those of the Generalized Huet–Sayegh (2S2P1D) model for asphalt mixture with 50% RA at <span class="html-italic">T<sub>R</sub></span> = 20 °C.</p>
Full article ">Figure 21
<p>Comparison between measured |E*|_exp and calculated |E*|_2S2P1D model for the asphalt mixture with 0% RA (the green dotted line represents the trend line of the comparison data, and the red line represents the identity function y = x).</p>
Full article ">Figure 22
<p>Comparison between measured |E*|_exp and calculated |E*|_2S2P1D model for the asphalt mixture with 30% RA (the green dotted line represents the trend line of the comparison data, and the red line represents the identity function y = x).</p>
Full article ">Figure 23
<p>Comparison between measured |E*|_exp and calculated |E*|_2S2P1D model for the asphalt mixture with 40% RA (the green dotted line represents the trend line of the comparison data, and the red line represents the identity function y = x).</p>
Full article ">Figure 24
<p>Comparison between measured |E*|_exp and calculated |E*|_2S2P1D model for the asphalt mixture with 50% RA (the green dotted line represents the trend line of the comparison data, and the red line represents the identity function y = x).</p>
Full article ">
19 pages, 5984 KiB  
Article
A Rapid Prediction of Suppressed Vibration in Composite Bridges Equipped with Constrained Layer Damping
by Quanmin Liu, Weiwang Fu, Lizhong Song, Kui Gao and Peipei Xu
Buildings 2024, 14(11), 3621; https://doi.org/10.3390/buildings14113621 - 14 Nov 2024
Viewed by 256
Abstract
The vibration characteristics of a composite bridge with constrained layer damping (CLD) were investigated using the wave and finite element method (WFEM), and the effects of the material and geometrical parameters of the CLD on the vibration reduction in the bridge were analyzed. [...] Read more.
The vibration characteristics of a composite bridge with constrained layer damping (CLD) were investigated using the wave and finite element method (WFEM), and the effects of the material and geometrical parameters of the CLD on the vibration reduction in the bridge were analyzed. Firstly, a numerical model for the dynamic response of a composite steel–concrete bridge using WFEM. The calculated acceleration of the bridge under the wheel–rail force obtained using this model was in good agreement with that obtained using the conventional finite element method and field measurements. Second, a segment model of the bridge with a CLD was established. The equation of motion based on the WFEM was solved to determine the dynamic response of the bridge induced by running trains. Finally, the effects of the covering area and CLD parameters on the vibration mitigation of steel–concrete bridges were analyzed. The results show that a reduction of 5–10 dB of the acceleration level of steel members in the full frequency range can be achieved by installing the CLD. A lower shear modulus of the viscoelastic core is beneficial for low-frequency vibration reduction in the bridge. However, a higher shear modulus of the damping layer is required for vibration mitigation in the high-frequency range. The vibration reduction in the composite bridge was more sensitive to the thickness of the constraining layer than to that of the damping layer. Full article
(This article belongs to the Section Building Structures)
Show Figures

Figure 1

Figure 1
<p>Segments of waveguide structure.</p>
Full article ">Figure 2
<p>Excitation diagram of an infinite waveguide structure.</p>
Full article ">Figure 3
<p>Waves amplitudes in a finite waveguide structure.</p>
Full article ">Figure 4
<p>Cross-section dimensions (unit: mm).</p>
Full article ">Figure 5
<p>WFE model of composite bridge.</p>
Full article ">Figure 6
<p>Dispersion curves of steel−concrete composite bridge.</p>
Full article ">Figure 7
<p>Typical wave modes at 25 Hz. The solid line means the wave mode and the dot line is the initial shape. (<b>a</b>) wave A; (<b>b</b>) wave B; (<b>c</b>) wave C; (<b>d</b>) wave D; (<b>e</b>) wave E; (<b>f</b>) wave F; and (<b>g</b>) wave G.</p>
Full article ">Figure 8
<p>Layout of CLD.</p>
Full article ">Figure 9
<p>Material loss factor and real shear modulus of viscoelastic layer at 20 °C.</p>
Full article ">Figure 10
<p>WFE model of composite bridge with CLD.</p>
Full article ">Figure 11
<p>Measured and computed acceleration levels of bottom flange of bridge without CLD.</p>
Full article ">Figure 12
<p>Measured and computed acceleration levels of bottom flange of bridge with CLD.</p>
Full article ">Figure 13
<p>Acceleration levels of bottom flange of bridge with and without CLD.</p>
Full article ">Figure 14
<p>Arrangement of CLD on web of bridge. (<b>a</b>) Laying area ratio of 50%; (<b>b</b>) Laying area ratio of 80%; (<b>c</b>) Laying area ratio of 100%.</p>
Full article ">Figure 15
<p>Vibration response of bridge at various laying area ratios.</p>
Full article ">Figure 16
<p>Vibration responses of bridge under various shear moduli.</p>
Full article ">Figure 17
<p>Vibration response of bridge with damping layer of various thicknesses.</p>
Full article ">Figure 18
<p>Vibration response of the bridge under different materials of constraining layer.</p>
Full article ">Figure 19
<p>Vibration response of bridge with constraining layer of various thicknesses.</p>
Full article ">
14 pages, 6126 KiB  
Article
Numerical Simulation of Nuclear Power Plant Pile Foundation Damage Under Earthquake Action
by Feng Xia, Wenhao Qi, Liping Jing, Zhan Wang and Xinyu Lu
Buildings 2024, 14(11), 3617; https://doi.org/10.3390/buildings14113617 - 14 Nov 2024
Viewed by 277
Abstract
This study investigates the pile foundation of a nuclear power plant situated on medium-soft soil. It employs an improved viscoelastic artificial boundary unit to accurately simulate the boundary conditions of the calculation area. The research utilizes a constitutive model of concrete damage plasticity [...] Read more.
This study investigates the pile foundation of a nuclear power plant situated on medium-soft soil. It employs an improved viscoelastic artificial boundary unit to accurately simulate the boundary conditions of the calculation area. The research utilizes a constitutive model of concrete damage plasticity for the pile foundation and an equivalent linearized model for the soil layer. Through large-scale shaking table experiments and numerical simulations, we explore the internal force distribution within the nuclear power structure’s pile foundation and assess the extent of the damage. The results indicate that damage primarily occurs in the medium-soft ground, concentrating in the upper part of the pile and affecting the entire cross-section. Subsequent numerical analyses were conducted after reinforcing the soil layer around the top of the pile. The findings demonstrate that this reinforcement leads to a more uniform and rational distribution of internal forces along the pile, significantly reducing damage. Notably, there is no severe damage extending across the entire cross-section after reinforcement. This outcome highlights the potential for improving the force distribution in the pile foundations of nuclear power structures through appropriate soil layer reinforcement. The insights gained from this study provide valuable guidance for the seismic design of nuclear power structures. Full article
(This article belongs to the Section Building Structures)
Show Figures

Figure 1

Figure 1
<p>Group pile arrangement and pile number diagram.</p>
Full article ">Figure 2
<p>Numerical modeling and material parameters.</p>
Full article ">Figure 3
<p>Pile Foundation Layout: (<b>a</b>) schematic diagram of pile foundation arrangement; (<b>b</b>) finite element diagram of pile foundation; (<b>c</b>) overall model of nuclear power plant; (<b>d</b>) pile foundation mode.</p>
Full article ">Figure 4
<p>Soil <math display="inline"><semantics> <mrow> <mi mathvariant="normal">G</mi> <mo>/</mo> <msub> <mrow> <mi mathvariant="normal">G</mi> </mrow> <mrow> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">x</mi> </mrow> </msub> <mo stretchy="false">~</mo> <msub> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> <mrow> <mi>d</mi> </mrow> </msub> <mo>,</mo> <mi mathvariant="sans-serif">λ</mi> <mo stretchy="false">~</mo> <msub> <mrow> <mi mathvariant="sans-serif">γ</mi> </mrow> <mrow> <mi>d</mi> </mrow> </msub> </mrow> </semantics></math> relationship curve ((<b>a</b>) silty clay, (<b>b</b>) silt).</p>
Full article ">Figure 5
<p>Stress–strain curve of pile material.</p>
Full article ">Figure 6
<p>Seismic acceleration time history curves in three directions.</p>
Full article ">Figure 7
<p>Comparison of site seismic response analysis results in X direction.</p>
Full article ">Figure 8
<p>Internal forces of pile.</p>
Full article ">Figure 9
<p>Pile compressive damage distribution. (<b>a</b>) Distribution of pile group body pressure damage; (<b>b</b>) single pile body pressure damage distribution.</p>
Full article ">Figure 10
<p>Soil treatment diagram.</p>
Full article ">Figure 11
<p>Internal forces of pile after soil reinforcement.</p>
Full article ">Figure 12
<p>Pile pressure damage distribution after soil layer treatment. (<b>a</b>) Pressure damage distribution of pile group after soil reinforcement; (<b>b</b>) single pile body pressure damage distribution.</p>
Full article ">
16 pages, 5445 KiB  
Article
Effect of Part Size, Displacement Rate, and Aging on Compressive Properties of Elastomeric Parts of Different Unit Cell Topologies Formed by Vat Photopolymerization Additive Manufacturing
by Lindsey B. Bezek, Sushan Nakarmi, Alexander C. Pantea, Jeffery A. Leiding, Nitin P. Daphalapurkar and Kwan-Soo Lee
Polymers 2024, 16(22), 3166; https://doi.org/10.3390/polym16223166 - 13 Nov 2024
Viewed by 281
Abstract
Due to its ability to achieve geometric complexity at high resolution and low length scales, additive manufacturing (AM) has increasingly been used for fabricating cellular structures (e.g., foams and lattices) for a variety of applications. Specifically, elastomeric cellular structures offer tunability of compliance [...] Read more.
Due to its ability to achieve geometric complexity at high resolution and low length scales, additive manufacturing (AM) has increasingly been used for fabricating cellular structures (e.g., foams and lattices) for a variety of applications. Specifically, elastomeric cellular structures offer tunability of compliance as well as energy absorption and dissipation characteristics. However, there are limited data available on compression properties for printed elastomeric cellular structures of different designs and testing parameters. In this work, the authors evaluate how unit cell topology, part size, the rate of compression, and aging affect the compressive response of polyurethane-based simple cubic, body-centered, and gyroid structures formed by vat photopolymerization AM. Finite element simulations incorporating hyperelastic and viscoelastic models were used to describe the data, and the simulated results compared well with the experimental data. Of the designs tested, only the parts with the body-centered unit cell exhibited differences in stress–strain responses at different part sizes. Of the compression rates tested, the highest displacement rate (1000 mm/min) often caused stiffer compressive behavior, indicating deviation from the quasi-static assumption and approaching the intermediate rate response. The cellular structures did not change in compression properties across five weeks of aging time, which is desirable for cushioning applications. This work advances knowledge on the structure–property relationships of printed elastomeric cellular materials, which will enable more predictable compressive properties that can be traced to specific unit cell designs. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Digital renderings of the simple cubic (SC), body-centered (BC), and gyroid (G) cellular structures with 5 mm unit cells patterned into cubes with side lengths made up of 3 unit cells; (<b>b</b>) samples of printed SC, BC, and G cellular structures with 5 mm sized unit cells and side lengths made up of 3, 6, and 12 unit cells.</p>
Full article ">Figure 2
<p>Evolution of compression response for the (<b>a</b>) SC, (<b>b</b>) BC, and (<b>c</b>) G structures with 12 × 12 × 12 unit cells strained to 80%. For individual structures, top rows are snapshots from experiments and the bottom rows are corresponding snapshots from simulations. The SC structure experiences global structural buckling, while the BC and G parts deform more uniformly according to the unit cell design.</p>
Full article ">Figure 3
<p>Compressive stress–strain profiles for (<b>a</b>) SC, (<b>b</b>) BC, and (<b>c</b>) G structures of different part sizes. The data are replotted separately for (<b>d</b>–<b>f</b>) 3 × 3 × 3, (<b>g</b>–<b>i</b>) 6 × 6 × 6, and (<b>j</b>–<b>l</b>) 12 × 12 × 12 part sizes for SC, BC, and G structures, respectively. Simulated results are indicated using ‘Model’ and plotted using black lines. Experimental results (n = 3) are plotted with red lines.</p>
Full article ">Figure 3 Cont.
<p>Compressive stress–strain profiles for (<b>a</b>) SC, (<b>b</b>) BC, and (<b>c</b>) G structures of different part sizes. The data are replotted separately for (<b>d</b>–<b>f</b>) 3 × 3 × 3, (<b>g</b>–<b>i</b>) 6 × 6 × 6, and (<b>j</b>–<b>l</b>) 12 × 12 × 12 part sizes for SC, BC, and G structures, respectively. Simulated results are indicated using ‘Model’ and plotted using black lines. Experimental results (n = 3) are plotted with red lines.</p>
Full article ">Figure 4
<p>Compressive stress–strain profiles for (<b>a</b>) SC, (<b>b</b>) BC, (<b>c</b>) G, and (<b>d</b>) S structures when subjected to varying displacement rates. n = 3.</p>
Full article ">Figure 5
<p>Compressive strains of the cellular structures tested up to five weeks after printing. Error bars represent one sample standard deviation across three samples.</p>
Full article ">Figure A1
<p>(<b>a</b>) Schoen I-graph-wrapped package (IWP) cellular structure; image reproduced from Nakarmi et. al with permission from the authors [<a href="#B29-polymers-16-03166" class="html-bibr">29</a>]; (<b>b</b>) comparison of compressive strains at 0.1 MPa stress for the IWP cellular structure that were subjected to different post-processing conditions: either 3 days or 20 min total in isopropyl alcohol (IPA) and with or without a 5 min UV cure submerged in DI water. Error bars represent one sample standard deviation across three samples.</p>
Full article ">
17 pages, 6130 KiB  
Article
Impact of Optimal Silane Concentration on the Rheological Properties and 3D Printing Performance of Al2O3-Acrylate Composite Slurries
by Kook-Hyun Ryu, Ung-Soo Kim, Jin-Ho Kim, Jung-Hoon Choi and Kyu-Sung Han
Materials 2024, 17(22), 5541; https://doi.org/10.3390/ma17225541 - 13 Nov 2024
Viewed by 265
Abstract
In this study, 3-trimethoxy-silylpropane-1-thiol (MPTMS) was used as a surface modifier for Al2O3 powder to systematically analyze the effects of MPTMS concentration on the rheological properties, photocuring characteristics, and 3D printing performance of photocurable composite slurries. MPTMS concentration significantly influenced [...] Read more.
In this study, 3-trimethoxy-silylpropane-1-thiol (MPTMS) was used as a surface modifier for Al2O3 powder to systematically analyze the effects of MPTMS concentration on the rheological properties, photocuring characteristics, and 3D printing performance of photocurable composite slurries. MPTMS concentration significantly influenced the rheological behavior of the slurry. Slurries containing 2 wt.% and 5 wt.% MPTMS exhibited a wide linear viscoelastic range (LVR). However, at concentrations of 10 wt.% and 20 wt.%, the LVR range narrowed, which led to reduced dispersion stability. In dispersion stability tests, the slurry with 2 wt.% MPTMS showed the most stable dispersion, while the 5 wt.% MPTMS concentration exhibited the highest photocuring rate. In 3D printing experiments, the 5 wt.% MPTMS concentration resulted in the most stable printed structures, whereas printing failures occurred with the 2 wt.% concentration. At 10 wt.% and 20 wt.%, internal cracking was observed, leading to structural defects. In conclusion, MPTMS forms silane bonds on the Al2O3 surface, significantly impacting the stability, rheological properties, and printing quality of Al2O3-acrylate composite slurries. An MPTMS concentration of 5 wt.% was found to be optimal, contributing to the formation of stable and robust structures. Full article
(This article belongs to the Special Issue Advanced Additive Manufacturing Processing of Ceramic Materials)
Show Figures

Figure 1

Figure 1
<p>Thermogravimetric analysis of Al<sub>2</sub>O<sub>3</sub> powders treated with various concentrations of SCA.</p>
Full article ">Figure 2
<p>Weight loss of Al<sub>2</sub>O<sub>3</sub> powders treated with various concentrations of SCA. The confidence level was 99.9%.</p>
Full article ">Figure 3
<p>FT-IR spectra of as-received and SCA-treated Al<sub>2</sub>O<sub>3</sub> powders.</p>
Full article ">Figure 4
<p>Results of the amplitude sweep test for the Al<sub>2</sub>O<sub>3</sub> slurries with different SCA concentrations; (<b>a</b>) 2 wt.%, (<b>b</b>) 5 wt.%, (<b>c</b>) 10 wt.%, and (<b>d</b>) 20 wt.%. All the measurements were conducted at 25 °C and a frequency of 10 rad/s.</p>
Full article ">Figure 5
<p>Influence of SCA concentration on the rheological properties of photocurable Al<sub>2</sub>O<sub>3</sub> slurries; (<b>a</b>) tan δ value as a function of oscillation frequency, and (<b>b</b>) shear stress of slurries with varying shear rates.</p>
Full article ">Figure 6
<p>Backscattering profile of photocurable Al<sub>2</sub>O<sub>3</sub> slurries with different SCA concentrations; (<b>a</b>) 2 wt.%, (<b>b</b>) 5 wt.%, (<b>c</b>) 10 wt.%, and (<b>d</b>) 20 wt.%.</p>
Full article ">Figure 7
<p>TSI curves of photocurable Al<sub>2</sub>O<sub>3</sub> slurries with different SCA concentrations. The confidence level was 95%.</p>
Full article ">Figure 8
<p>Heat flow and conversion of photocurable Al<sub>2</sub>O<sub>3</sub> slurries with different SCA concentrations; (<b>a</b>) SCA 2 wt.%, (<b>b</b>) SCA 5 wt.%, (<b>c</b>) SCA 10 wt.%, and (<b>d</b>) SCA 20 wt.%.</p>
Full article ">Figure 8 Cont.
<p>Heat flow and conversion of photocurable Al<sub>2</sub>O<sub>3</sub> slurries with different SCA concentrations; (<b>a</b>) SCA 2 wt.%, (<b>b</b>) SCA 5 wt.%, (<b>c</b>) SCA 10 wt.%, and (<b>d</b>) SCA 20 wt.%.</p>
Full article ">Figure 9
<p>Printing results of photocurable Al<sub>2</sub>O<sub>3</sub> slurry with 2 wt.% SCA concentration under different exposure conditions; (<b>a</b>) surface image facing build platform and (<b>b</b>) surface image facing tray.</p>
Full article ">Figure 10
<p>Printing results of photocurable Al<sub>2</sub>O<sub>3</sub> slurry with 5 wt.% SCA concentration under different exposure conditions; (<b>a</b>) surface image facing build platform, (<b>b</b>) surface image facing tray, and (<b>c</b>) image of printed samples.</p>
Full article ">Figure 11
<p>Printing results of photocurable Al<sub>2</sub>O<sub>3</sub> slurry with 10 wt.% SCA concentration under different exposure conditions; (<b>a</b>) surface image facing build platform, (<b>b</b>) surface image facing tray, and (<b>c</b>) image of printed samples.</p>
Full article ">Figure 12
<p>Printing results of photocurable Al<sub>2</sub>O<sub>3</sub> slurry with 20 wt.% SCA concentration under different exposure conditions; (<b>a</b>) surface image facing build platform, (<b>b</b>) surface image facing tray, and (<b>c</b>) image of printed samples.</p>
Full article ">Figure 13
<p>X-ray CT cross-sectional images of printed Al<sub>2</sub>O<sub>3</sub>-acrylate composites with different SCA concentrations; (<b>a</b>) 2 wt.%, (<b>b</b>) 5 wt.%, (<b>c</b>) 10 wt.%, and (<b>d</b>) 20 wt.%.</p>
Full article ">Figure 13 Cont.
<p>X-ray CT cross-sectional images of printed Al<sub>2</sub>O<sub>3</sub>-acrylate composites with different SCA concentrations; (<b>a</b>) 2 wt.%, (<b>b</b>) 5 wt.%, (<b>c</b>) 10 wt.%, and (<b>d</b>) 20 wt.%.</p>
Full article ">
21 pages, 2634 KiB  
Article
Effect of the Ratio of Protein to Water on the Weak Gel Nonlinear Viscoelastic Behavior of Fish Myofibrillar Protein Paste from Alaska Pollock
by Timilehin Martins Oyinloye and Won Byong Yoon
Gels 2024, 10(11), 737; https://doi.org/10.3390/gels10110737 - 13 Nov 2024
Viewed by 333
Abstract
The linear and nonlinear rheological behaviors of fish myofibrillar protein (FMP) paste with 75%, 82%, and 90% moisture content were evaluated using small-amplitude oscillatory shear (SAOS) and large-amplitude oscillatory shear (LAOS) tests. SAOS revealed pastes with 75% and 82% moisture exhibited solid-like behavior, [...] Read more.
The linear and nonlinear rheological behaviors of fish myofibrillar protein (FMP) paste with 75%, 82%, and 90% moisture content were evaluated using small-amplitude oscillatory shear (SAOS) and large-amplitude oscillatory shear (LAOS) tests. SAOS revealed pastes with 75% and 82% moisture exhibited solid-like behavior, characterized by higher storage modulus (G′) than loss modulus (G″), indicative of weak gel properties with a strong protein interaction. In contrast, the 90% moisture content showed more viscous behavior due to weakened protein–protein entanglements. The frequency exponent (n′ and n″) from the power law equation varied slightly (0.24 to 0.36), indicating limited sensitivity to changes in deformation rate during SAOS. LAOS tests revealed significant structural changes, with Lissajous–Bowditch curves revealing early nonlinearities at 10% strain for 90% moisture content. Decomposed Chebyshev coefficients (e3/e1, v3/v1, S, and T) indicated strain stiffening at lower strains for the 75% and 82% moisture pastes (i.e., < 50% strain for 75% and < 10% strain for 82%), transitioning to strain thinning at higher strains. Additionally, numerical model confirmed the predictability of the 3D printing process from the nonlinear rheological data, confirmed the suitability of the 75% and 82% moisture pastes for applications requiring structural integrity. These insights are essential for optimizing processing conditions in industrial applications. The findings suggest that the 75% and 82% moisture pastes are suitable for applications requiring structural integrity, while the 90% moisture paste is ideal for flow-based processes. These insights are essential for optimizing processing conditions in industrial applications. Full article
(This article belongs to the Special Issue Food Gel-Based Systems: Gel-Forming and Food Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Strain sweep analysis for FMP paste with different moisture contents, obtained at frequency 1 Hz.</p>
Full article ">Figure 2
<p>Rheological properties of the FMP paste according to moisture content: (<b>a</b>) 75%, (<b>b</b>) 82%, (<b>c</b>) 90% (G′, G″, phase angle), (<b>d</b>) combined G′ and G″ for all moisture contents, and (<b>e</b>) combined phase angle for all moisture contents.</p>
Full article ">Figure 3
<p>Damping factor (G″/G′) for FMP paste at different frequencies (1 rad/s, 40 rad/s, and 100 rad/s) and moisture contents (75%, 82%, and 90%) (<b>a</b>), and power law constants (n′ and n″) for FMP paste at different moisture contents (<b>b</b>). Different lowercase letters indicate significant differences within the same measurement factor (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Stress vs. strain analysis of FMP paste.</p>
Full article ">Figure 5
<p>Elastic Lissajous–Bowditch plots of FMP paste at frequency of 1 rad/s and different strains of 5%, 10%, 50% 100%, 200%, and 500%. Y-axis denote shear stress (Pa), and x-axis denotes normalized strain (%).</p>
Full article ">Figure 6
<p>Viscous Lissajous–Bowditch plots of FMP paste at frequency of 1 rad/s and different strains of 5%, 10%, 50% 100%, 200%, and 500%. Y-axis denote shear stress (Pa), and x-axis denotes normalized shear rate (s<sup>−1</sup>).</p>
Full article ">Figure 7
<p>The changes in (<b>a</b>) e<sub>3</sub>/e<sub>1</sub>, (<b>b</b>) v<sub>3</sub>/v<sub>1</sub>, (<b>c</b>) S, and (<b>d</b>) T with respect to strain, γ (%), of FMP paste at different moisture contents.</p>
Full article ">Figure 8
<p>Three-dimensional-printed FMP paste at different time intervals during extrusion process. The images show the paste at (<b>a</b>) 20 s, (<b>b</b>) 100 s, (<b>c</b>) end of printing time, and (<b>d</b>) 20 min after printing.</p>
Full article ">Figure 9
<p>Simulated 3D printing process for FMP paste at different moisture contents. (<b>a</b>) Contour image of deformation in the completely printed model and (<b>b</b>) deformation value in the paste.</p>
Full article ">Figure 10
<p>Schematic diagram of 3D printed model (<b>a</b>) and die swell analysis procedure for 3D printed FMP paste with different moisture contents: (<b>b</b>) Image of extruded paste at 50 mm length, (<b>c</b>) Binary image of extruded paste, and (<b>d</b>) Image with bounding box.</p>
Full article ">
24 pages, 6049 KiB  
Review
Food Gels Based on Polysaccharide and Protein: Preparation, Formation Mechanisms, and Delivery of Bioactive Substances
by Yong Guo, Chao Ma, Yan Xu, Lianxin Du and Xin Yang
Gels 2024, 10(11), 735; https://doi.org/10.3390/gels10110735 - 13 Nov 2024
Viewed by 421
Abstract
Hydrogels have a unique three-dimensional network that can create a good environment for the loading of functional compounds; hence, they have considerable potential in the delivery of bioactive substances. Natural macromolecular substances (proteins, polysaccharides) have the features of low toxicity, degradability, and biosafety; [...] Read more.
Hydrogels have a unique three-dimensional network that can create a good environment for the loading of functional compounds; hence, they have considerable potential in the delivery of bioactive substances. Natural macromolecular substances (proteins, polysaccharides) have the features of low toxicity, degradability, and biosafety; thus, they can be employed in the manufacture of hydrogels in the food sector. With its customizable viscoelastic and porous structure, hydrogels are believed to be good bioactive material delivery vehicles, which can effectively load polyphenols, vitamins, probiotics, and other active substances to prevent their influence from the external environment, thereby improving its stability. In this research, the common raw materials, preparation methods, and applications in the delivery of bioactive elements of food gels were examined; this study aimed at presenting new ideas for the development and utilization of protein-based food gels. Full article
(This article belongs to the Special Issue Food Gels: Structure and Function)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of the gel network formed by different-sized aggregates of glutaraldehyde-cross-linked soy protein, forming a transparent gel of pea protein isolate [<a href="#B39-gels-10-00735" class="html-bibr">39</a>]. (<b>b</b>) Schematic diagram of the preparation of pea protein isolate by thermos-reversible gel formation process [<a href="#B42-gels-10-00735" class="html-bibr">42</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic representation of TGase-induced gel characterization of pea-protein/zeatin complexes [<a href="#B48-gels-10-00735" class="html-bibr">48</a>]. (<b>b</b>) Schematic representation of ultrasound-induced self-assembled modified soy protein/dextran nanogel as a delivery vehicle for riboflavin [<a href="#B50-gels-10-00735" class="html-bibr">50</a>]. (<b>c</b>) Schematic representation of soy isolate protein/cellulose nanofibrous composite gels prepared as a fat substitute [<a href="#B53-gels-10-00735" class="html-bibr">53</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic preparation of whey protein–chitosan composite hydrogels loaded with curcumin [<a href="#B117-gels-10-00735" class="html-bibr">117</a>]. (<b>b</b>) Schematic diagram of microgel preparation loaded with <span class="html-italic">Pediococcus pentosaceus</span> Li05 (the blue dots represent probiotics) [<a href="#B121-gels-10-00735" class="html-bibr">121</a>]. (<b>c</b>) Schematic preparation of bovine serum albumin–citrus pectin hydrogels providing vitamin C and SEM characterization of their microstructures [<a href="#B119-gels-10-00735" class="html-bibr">119</a>].</p>
Full article ">
15 pages, 5618 KiB  
Article
MXene/Bacterial Cellulose Hybrid Materials for Sustainable Soft Electronics
by Wojciech Guziewicz, Shreyas Srivatsa, Marcel Zambrzycki, Michał Dziadek, Piotr Szatkowski, Patryk Szymczak, Katarzyna Berent, Marianna Marciszko-Wiąckowska, Marta Radecka, Agata Kołodziejczyk and Tadeusz Uhl
Materials 2024, 17(22), 5513; https://doi.org/10.3390/ma17225513 - 12 Nov 2024
Viewed by 358
Abstract
This work evaluated bacterial cellulose (BC) as a possible biodegradable soft electronics substrate in comparison to polyethylene terephthalate (PET), while also focusing on evaluating hybrid MXene/BC material as potential flexible electronic sensor. Material characterization studies revealed that the BC material structure consists of [...] Read more.
This work evaluated bacterial cellulose (BC) as a possible biodegradable soft electronics substrate in comparison to polyethylene terephthalate (PET), while also focusing on evaluating hybrid MXene/BC material as potential flexible electronic sensor. Material characterization studies revealed that the BC material structure consists of nanofibers with diameters ranging from 70 to 140 nm, stacked layer-by-layer. BC samples produced are sensitive to post-treatment with isopropanol resulting in a change of structural and mechanical properties. The viscoelastic properties of the BC substrates have been studied experimentally in comparison with the PET film. Aged BC substrate showcased similar viscoelastic properties stability, while exhibiting better properties above 70 °C, with total storage modulus change of −15% and loss modulus change of 21%. MXenes prepared using the Minimally Intensive Layer Delamination (MILD) method were screen-printed onto BC substrates and PET films to form MXene/BC (MX/BC) and MXene/PET (MX/PET) devices. The electrical properties results showcased different resistive behavior on both BC and PET substrate samples with different impedance moduli. MX/PET presented lower sheet resistance of around 156 Ω·sq−1, while MX/BC was 2733 Ω·sq−1. Finally, the MX/BC and MX/PET devices were subjected to repeatable quasi-static load tests and the piezoresistive sensing behavior of the devices has been reported. Full article
Show Figures

Figure 1

Figure 1
<p>Bacterial cellulose during production (<b>A</b>), bacteria (<b>B</b>), and yeast (<b>C</b>) present in dried, unsterilized material.</p>
Full article ">Figure 2
<p>MXene printed onto bacterial cellulose (<b>left</b>) and PET (<b>right</b>). Silver paste and copper wires were added for electrical and sensitivity measurements.</p>
Full article ">Figure 3
<p>(<b>A</b>) SEM image of BC fibers. (<b>B</b>) SEM cross-section of dry BC sheet. (<b>C</b>) AFM image of BC fibers. (<b>D</b>) Fiber diameter distribution chart.</p>
Full article ">Figure 4
<p>(<b>A</b>) FT–IR spectrum of BC. (<b>B</b>) XRD diffractogram of MAX precursor and obtained MXene (<b>top</b>) and XRD diffractogram of untreated BC and isopropanol treated BC (<b>bottom</b>).</p>
Full article ">Figure 5
<p>Young’s modulus, tensile strength, and elongation at break for tested samples (n = 10). Same uppercase letters mean no statistically significant difference between the results.</p>
Full article ">Figure 6
<p>Storage modulus, loss modulus and tangent of samples tested using DMA.</p>
Full article ">Figure 7
<p>(<b>A</b>) Current-voltage response of MXene layers printed on different substrates. (<b>B</b>) Nyquist spectra and (<b>C</b>) Impedance modulus and phase shift in Bode representation of MXene films on BC and PET substrates.</p>
Full article ">Figure 8
<p>Resistance measurements during cycling loading of MX/PET and MX/BC samples.</p>
Full article ">
20 pages, 3450 KiB  
Article
Rheology of Cellulosic Microfiber Suspensions Under Oscillatory and Rotational Shear for Biocomposite Applications
by Helena Cristina Vasconcelos, Henrique Carrêlo, Telmo Eleutério, Maria Gabriela Meirelles, Reşit Özmenteş and Roberto Amorim
Compounds 2024, 4(4), 688-707; https://doi.org/10.3390/compounds4040042 - 12 Nov 2024
Viewed by 277
Abstract
This study investigates the rheological behavior of cellulose microfiber suspensions derived from kahili ginger stems (Hedychium gardnerianum), an invasive species, in two adhesive matrices: a commercial water-based adhesive (Coplaseal®) and a casein-based adhesive made from non-food-grade milk, referred to [...] Read more.
This study investigates the rheological behavior of cellulose microfiber suspensions derived from kahili ginger stems (Hedychium gardnerianum), an invasive species, in two adhesive matrices: a commercial water-based adhesive (Coplaseal®) and a casein-based adhesive made from non-food-grade milk, referred to as K and S samples, respectively. Rheological analyses were performed using oscillatory and rotational shear tests conducted at 25 °C, 50 °C, and 75 °C to assess the materials’ viscoelastic properties more comprehensively. Oscillatory tests across a frequency range of 1–100 rad/s assessed the storage modulus (G′) and loss modulus (G″), while rotational shear tests evaluated apparent viscosity and shear stress across shear rates from 0.1 to 1000 s−1. Fiber-free samples consistently showed lower moduli than fiber-containing samples at all frequencies. The incorporation of fibers increased the dynamic moduli in both K and S samples, with a quasi-plateau observed at lower frequencies, suggesting solid-like behavior. This trend was consistent in all tested temperatures. As frequencies increased, the fiber network was disrupted, transitioning the samples to fluid-like behavior, with a marked increase in G′ and G″. This transition was more pronounced in K samples, especially above 10 rad/s at 25 °C and 50 °C, but less evident at 75 °C. This shift from solid-like to fluid-like behavior reflects the transition from percolation effects at low frequencies to matrix-dominated responses at high frequencies. In contrast, S samples displayed a wider frequency range for the quasi-plateau, with less pronounced moduli changes at higher frequencies. At 75 °C, the moduli of fiber-containing and fiber-free S samples nearly converged at higher frequencies, indicating similar effects of the fiber and matrix components. Both fiber-reinforced and non-reinforced suspensions exhibited pseudoplastic (shear-thinning) behavior. Fiber-containing samples exhibited higher initial viscosity, with K samples displaying greater differences between fiber-reinforced and non-reinforced systems compared to S samples, where the gap was narrower. Interestingly, S samples exhibited overall higher viscosity than K samples, implying a reduced influence of fibers on the viscosity in the S matrix. This preliminary study highlights the complex interactions between cellulosic fiber networks, adhesive matrices, and rheological conditions. The findings provide a foundation for optimizing the development of sustainable biocomposites, particularly in applications requiring precise tuning of rheological properties. Full article
Show Figures

Figure 1

Figure 1
<p>SEM micrograph of mechanically extracted fiber from the stems of the Kahili ginger plant (<span class="html-italic">Hedychium gardnerianum</span>).</p>
Full article ">Figure 2
<p>(<b>a</b>) Long <span class="html-italic">Hedychium gardnerianum</span> fibers (&gt;10 cm); (<b>b</b>) short <span class="html-italic">Hedychium gardnerianum</span> fibers (between 0.2 and 5 mm).</p>
Full article ">Figure 3
<p>Oscillatory tests for k0 and k10 at different temperatures: (<b>a</b>) 25 °C, (<b>b</b>) 50 °C, and (<b>c</b>) 75 °C. Each graph, with logarithmic scales on both axes, displays storage modulus (G′) and loss modulus (G″) as functions of angular frequency (ω) in rad/s.</p>
Full article ">Figure 4
<p>Comparison of the storage modulus (G′) and loss modulus (G″) of Kappa samples across different temperatures (25, 50, and 75 °C). Panel (<b>a</b>) shows data for Kappa samples with fibers (k10), while panel (<b>b</b>) displays data for Kappa samples without fibers (k0). Note: Error bars are not plotted here for simplicity.</p>
Full article ">Figure 5
<p>Oscillatory tests for s0 and s10 at different temperatures: (<b>a</b>) 25 °C, (<b>b</b>) 50 °C, and (<b>c</b>) 75 °C. Each graph, with logarithmic scales on both axes, displays the storage modulus (G′) and the loss modulus (G″) as functions of angular frequency (ω) in rad/s.</p>
Full article ">Figure 6
<p>Comparison of the storage modulus (G′) and loss modulus (G″) of Super samples across different temperatures (25, 50, and 75 °C). Panel (<b>a</b>) shows data for Super samples with fibers (s10), while the panel (<b>b</b>) displays data for Super samples without fibers (s0). Note: Error bars are not plotted here for simplicity.</p>
Full article ">Figure 7
<p>Continuous tests of K samples, without fibers and with 10% fibers, at different temperatures (25 °C, 50 °C, and 75 °C): Apparent viscosity (<b>a</b>); shear stress (<b>b</b>).</p>
Full article ">Figure 8
<p>End of continuous test for k25,10. Expulsion of the sample from the gap is observed.</p>
Full article ">Figure 9
<p>Post-rheological test observations of K samples at different temperatures: (<b>a</b>) Sample at 25 °C showing a liquid state without significant thickening; (<b>b</b>) sample at 50 °C with visible thick layer formation beginning to occur; (<b>c</b>) sample at 75 °C showing substantial film formation along edge zones; (<b>d</b>) another view of 75 °C sample, highlighting more pronounced thick layers at edges; (<b>e</b>) sample tested at 75 °C showing reduced volume after testing due to thick layer formation.</p>
Full article ">Figure 10
<p>Continuous tests S samples, without fibers and with 10% fibers at different temperatures (25, 50, and 75 °C): (<b>a</b>) viscosity; (<b>b</b>) shear stress.</p>
Full article ">Figure 11
<p>(<b>a</b>) Fiber-free S samples at 25 °C after rotational test with gap leak; (<b>b</b>) fiber-free S samples at 75 °C following oscillatory test; (<b>c</b>) fiber-free S samples at 50 °C after rotational testing.</p>
Full article ">Figure 12
<p>S sample with fibers after an oscillatory test at 50 °C.</p>
Full article ">
14 pages, 3225 KiB  
Article
Effects of Geometry and Supporting Silicone Layers on the Performance of Conductive Composite High-Deflection Strain Gauges
by Hailey E. Jones, Spencer A. Baker, Jadyn J. Christensen, Tyler Hutchinson, Heather A. Leany, Ulrike H. Mitchell, Anton E. Bowden and David T. Fullwood
J. Compos. Sci. 2024, 8(11), 467; https://doi.org/10.3390/jcs8110467 - 11 Nov 2024
Viewed by 581
Abstract
Piezoresistive sensors composed of nickel nanostrands, nickel-coated carbon fibers, and silicone can be used to measure large physical deflections but exhibit viscoelastic properties and creep, leading to a complex and nonlinear electrical response that is difficult to interpret. This study considers the impact [...] Read more.
Piezoresistive sensors composed of nickel nanostrands, nickel-coated carbon fibers, and silicone can be used to measure large physical deflections but exhibit viscoelastic properties and creep, leading to a complex and nonlinear electrical response that is difficult to interpret. This study considers the impact of modifying the geometry and architecture of the sensors on their mechanical and electrical performance. Varying the sensor thickness leads to potentially significant differences in conductive fiber alignment, while adding external layers of pure silicone provides elastic support for the sensors, potentially reducing their extreme viscoelastic nature. The impact of such modifications on both mechanical and electrical behavior was assessed by analyzing strain to failure, the magnitude of hysteresis with cycling, the repeatability of the electro-mechanical response, the strain level at which resistance begins to monotonically decrease, and the drift in electrical response with cycling. The results indicate that thicker single-layer sensors have less electrical drift. Sensors with a multilayered architecture exhibit several improvements in behavior, such as increasing the range of the monotonic region by approximately 52%. These improvements become more significant as the thickness of the pure silicone layers increases. Full article
(This article belongs to the Special Issue Feature Papers in Journal of Composites Science in 2024)
Show Figures

Figure 1

Figure 1
<p>Burgers’ rheological model. <span class="html-italic">k</span><sub>1</sub> is the elastic spring constant, <span class="html-italic">η<sub>1</sub></span> is the viscoplastic damper coefficient, <span class="html-italic">k</span><sub>2</sub> is the viscoelastic spring constant, and <span class="html-italic">η<sub>2</sub></span> is the viscoelastic damper coefficient.</p>
Full article ">Figure 2
<p>Average resistance at 20% strain for different sensor thicknesses of single-layer sensors.</p>
Full article ">Figure 3
<p>(<b>a</b>) Completed single-layer sensors. (<b>b</b>) Complete multilayer sensors with metal snaps.</p>
Full article ">Figure 4
<p>Multilayered sensor in Instron during testing.</p>
Full article ">Figure 5
<p>(<b>a</b>) The movement of the Instron for a multilayered calibration. (<b>b</b>) The movement of the Instron to 20 random strains for one sensor.</p>
Full article ">Figure 6
<p>Point of failure for single- and multilayered sensors at all rates of pulling.</p>
Full article ">Figure 7
<p>Stress–strain response of a sensor over cycling for a (<b>a</b>) single-layered, <b>(b</b>) thin multilayered, (<b>c</b>) medium multilayered, and (<b>d</b>) thick multilayered sensor. Darkest to lightest is the first to last cycle.</p>
Full article ">Figure 7 Cont.
<p>Stress–strain response of a sensor over cycling for a (<b>a</b>) single-layered, <b>(b</b>) thin multilayered, (<b>c</b>) medium multilayered, and (<b>d</b>) thick multilayered sensor. Darkest to lightest is the first to last cycle.</p>
Full article ">Figure 8
<p>(<b>a</b>) Average difference in normalized area between the first and second cycle. (<b>b</b>) Average difference in normalized area from cycle to cycle. This was determined for cycles two through five.</p>
Full article ">Figure 9
<p>Resistance–strain curves for (<b>a</b>) single-layered, (<b>b</b>) 3.2 mm single-layered, and (<b>c</b>) thick multilayered sensor from cycle 2 to cycle 15. The fifteenth cycle is the lightest line, and the second cycle is the darkest line. Numerals on the images indicate 1. Maximum resistance 2. Drift magnitude 3. Strain at the critical point.</p>
Full article ">Figure 10
<p>Resistance–strain curves. The second cycle is the darkest line, and the last cycle is the lightest line. Comparing 1. maximum resistance, 2. relative resistance drift magnitude, and 3. strain shift for (<b>a</b>) single-layer and (<b>b</b>) multilayer sensors, respectively. Average strain, force, and relative standard deviations at failure, disregarding rate.</p>
Full article ">Figure 11
<p>Comparing true and predicted strain. Graphs (<b>a</b>,<b>b</b>) compare known and estimated strain with normalized resistance for single- and multilayered sensors, respectively. Graphs (<b>c</b>,<b>d</b>) show an R<sup>2</sup> representation, comparing true and modeled strain for single- and multilayered sensors, respectively.</p>
Full article ">
Back to TopTop