[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
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 (2,530)

Search Parameters:
Keywords = glass transition

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 1279 KiB  
Article
Self-Healable, Transparent, Biodegradable, and Shape Memorable Polyurethanes Derived from Carbon Dioxide-Based Diols
by Xin Huang, TingTing Zhao, ShuanJin Wang, Dongmei Han, Sheng Huang, Hui Guo, Min Xiao and Yuezhong Meng
Molecules 2024, 29(18), 4364; https://doi.org/10.3390/molecules29184364 (registering DOI) - 13 Sep 2024
Viewed by 216
Abstract
A series of CO2-based thermoplastic polyurethanes (TPUs) were prepared using CO2-based poly(polycarbonate) diol (PPCDL), 4,4′-methylenebis (cyclohexyl isocyanate) (HMDI), and polypropylene glycol (PPG and 1,4-butanediol (BDO) as the raw materials. The mechanical, thermal, optical, and barrier properties shape memory behaviors, [...] Read more.
A series of CO2-based thermoplastic polyurethanes (TPUs) were prepared using CO2-based poly(polycarbonate) diol (PPCDL), 4,4′-methylenebis (cyclohexyl isocyanate) (HMDI), and polypropylene glycol (PPG and 1,4-butanediol (BDO) as the raw materials. The mechanical, thermal, optical, and barrier properties shape memory behaviors, while biocompatibility and degradation behaviors of the CO2-based TPUs are also systematically investigated. All the synthesized TPUs are highly transparent amorphous polymers, with one glass transition temperature at ~15–45 °C varying with hard segment content and soft segment composition. When PPG is incorporated into the soft segments, the resultant TPUs exhibit excellent self-healing and shape memory performances with the average shape fixity ratio and shape recovery ratio as high as 98.9% and 88.3%, respectively. Furthermore, the CO2-based TPUs also show superior water vapor permeability resistance, good biocompatibility, and good biodegradation properties, demonstrating their pretty competitive potential in the polyurethane industry applications. Full article
(This article belongs to the Special Issue Biodegradable Functional Copolymers)
13 pages, 5792 KiB  
Article
Modification of the Surface Crystallinity of Polyphenylene Sulfide and Polyphthalamide Treated by a Pulsed-Arc Atmospheric Pressure Plasma Jet
by Abdessadk Anagri, Sarab Ben Saïd, Cyrille Bazin, Farzaneh Arefi-Khonsari and Jerome Pulpytel
Polymers 2024, 16(18), 2582; https://doi.org/10.3390/polym16182582 (registering DOI) - 12 Sep 2024
Viewed by 286
Abstract
Atmospheric plasma jets generated from air or nitrogen using commercial sources with relatively high energy densities are commonly used for industrial applications related to surface treatments, especially to increase the wettability of polymers or to deposit thin films. The heat fluxes to which [...] Read more.
Atmospheric plasma jets generated from air or nitrogen using commercial sources with relatively high energy densities are commonly used for industrial applications related to surface treatments, especially to increase the wettability of polymers or to deposit thin films. The heat fluxes to which the substrates are subjected are typically in the order of 100–300 W/cm2, depending on the treatment conditions. The temperature rise in the treated polymer substrates can have critical consequences, such as a change in the surface crystallinity or even the surface degradation of the materials. In this work, we report the phase transitions of two semicrystalline industrial-grade polymer resins reinforced with glass fibers, namely polyphenylene sulfide (PPS) and polyphthalamide (PPA), subjected to plasma treatments, as well as the modeling of the associated heat transfer phenomena using COMSOL Multiphysics. Depending on the treatment time, the surface of PPS becomes more amorphous, while PPA becomes more crystalline. These results show that the thermal history of the materials must be considered when implementing surface engineering by this type of plasma discharge. Full article
(This article belongs to the Special Issue Plasma Processing of Polymers, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photographs of (<b>a</b>) the free plasma jet and (<b>b</b>) the plasma jet impinging on a substrate; (<b>c</b>) flow regions of an impinging jet (Adapted with permission from [<a href="#B17-polymers-16-02582" class="html-bibr">17</a>], Elsevier, 2006).</p>
Full article ">Figure 2
<p>Evolution of the thermocouple temperature as a function of (<b>a</b>) <span class="html-italic">T<sub>jet</sub></span> with <span class="html-italic">h</span> = 2000 W/m<sup>2</sup>.K and (<b>b</b>) <span class="html-italic">h</span> with <span class="html-italic">T<sub>jet</sub></span> = 793 K.</p>
Full article ">Figure 3
<p>Substrate meshing, plasma torch trajectory, and heat transfer mechanisms.</p>
Full article ">Figure 4
<p>The surface temperature of PPS at <span class="html-italic">t</span> = 9 s. The first 4 lines of the torch pattern are also shown.</p>
Full article ">Figure 5
<p>The maximum temperature on each line as a function of the line speed.</p>
Full article ">Figure 6
<p>The temperature profile on a line as a function of time. The maximum temperature observed on each line corresponds to the center of the plasma jet. The line speed was 5 m/min.</p>
Full article ">Figure 7
<p>(<b>a</b>) Temperature profile along the depth of the material at a given time and (<b>b</b>) along the cut line at the same time. The line speed was 5 m/min.</p>
Full article ">Figure 8
<p>(<b>a</b>) Deconvolution of the WAXD diffractograms of untreated PPS using 3 Gaussian components at 2θ = 18.6°, 19.9° and 20.5°. The baseline was corrected by interpolation and <span class="html-italic">R</span><sup>2</sup> = 0.994; (<b>b</b>) diffractograms of PPS as a function of line speed.</p>
Full article ">Figure 9
<p>(<b>a</b>) Deconvolution of the WAXD diffractograms of untreated PPA using 5 Gaussian components at 2θ = 18.4° and 22.4°, assigned to α1 and α2 crystal, respectively, and 20.3° assigned to the γ crystal phase at 2θ = 20.2 and 24.9, assigned to the amorphous phase and glass fiber, respectively. The baseline was corrected by interpolation and <span class="html-italic">R</span><sup>2</sup> = 0.997; (<b>b</b>) diffractograms of PPA as a function of line speed.</p>
Full article ">Figure 10
<p>Evolution of the percentage of amorphous phase in PPS and PPA treated by air or nitrogen plasma as a function of line speed.</p>
Full article ">Figure 11
<p>DSC curve at a heating rate of 10 K/min for PPA composite; the first and second heat scans.</p>
Full article ">Figure 12
<p>DSC curve at a heating rate of 10 K/min for PPS composite; the first and second heat scans.</p>
Full article ">
15 pages, 7673 KiB  
Article
Tensile Deformation Mechanism of an In Situ Formed Ti-Based Bulk Metallic Glass Composites
by Haiyun Wang, Na Chen, Huanwu Cheng, Yangwei Wang and Denghui Zhao
Materials 2024, 17(18), 4486; https://doi.org/10.3390/ma17184486 - 12 Sep 2024
Viewed by 211
Abstract
Ti-based bulk metallic glass composites (BMGMCs) containing an in situ formed metastable β phase normally exhibit enhanced plasticity attributed to induced phase transformation or twinning. However, the underlying deformation micromechanism remains controversial. This study investigates a novel deformation mechanism of Ti-based BMGMCs with [...] Read more.
Ti-based bulk metallic glass composites (BMGMCs) containing an in situ formed metastable β phase normally exhibit enhanced plasticity attributed to induced phase transformation or twinning. However, the underlying deformation micromechanism remains controversial. This study investigates a novel deformation mechanism of Ti-based BMGMCs with a composition of Ti42.3Zr28Cu8.3Nb4.7Ni1.7Be15 (at%). The microstructures after tension were analyzed using advanced electron microscopy. The dendrites were homogeneously distributed in the glassy matrix with a volume fraction of 55 ± 2% and a size of 1~5 μm. The BMGMCs deformed in a serrated manner with a fracture strength (σf) of ~1710 MPa and a fracture strain of ~7.1%, accompanying strain hardening. The plastic deformation beyond yielding was achieved by a synergistic action, which includes shear banding, localized amorphization and a localized BCC (β-Ti) to HCP (α-Ti) structural transition. The localized amorphization was caused by high local strain rates during shear band extension from the amorphous matrix to the crystalline reinforcements. The localized structural transition from BCC to HCP resulted from accumulating concentrated stress during deformation. The synergistic action enriches our understanding of the deformation mechanism of Ti-based BMGMCs and also sheds light on material design and performance improvement. Full article
(This article belongs to the Special Issue Synthesis, Sintering, and Characterization of Composites)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Back scattered SEM image showing dendritic reinforcements were uniformly distributed in the glass matrix; (<b>b</b>) XRD pattern of the as-cast Ti-based BMGMCs showing crystalline peaks were overlapped with amorphous hump.</p>
Full article ">Figure 2
<p>TEM and HRTEM images of the as-cast Ti-BMGMCs: (<b>a</b>) bright field TEM image, selected area electron diffraction (SAED) pattern of (<b>b</b>) β-Ti dendrites, (<b>c</b>) glass matrix, HRTEM images of (<b>d</b>) the interface between β-Ti dendritic phase and glassy phase, (<b>e</b>) β-Ti dendrites and (<b>f</b>) glass matrix.</p>
Full article ">Figure 3
<p>(<b>a</b>) Room-temperature compressive stress–strain curve; and (<b>b</b>) room-temperature tensile stress–strain curve and its corresponding work-hardening rate–strain curve of the Ti-based amorphous matrix composites.</p>
Full article ">Figure 4
<p>(<b>a</b>) XRD pattern of the Ti-based BMGMCs before and after tension deformation; SEM images of (<b>b</b>) fracture surface of the specimen after tension in low magnification, (<b>c</b>) fracture surface morphology of dimples and (<b>d</b>) fracture surface morphology of river-like pattern in high magnification.</p>
Full article ">Figure 5
<p>SEM images of tensile deformed Ti-based BMGMCs at room temperature: (<b>a</b>,<b>b</b>) lateral side images in low magnification; (<b>c</b>) profuse shear band near fracture; (<b>d</b>) detailed shear banding in high magnification; (<b>e</b>) deformation feature in dual phase region; (<b>f</b>) magnified image of Zone I showing multiple shear bands; (<b>g</b>) magnified image of Zone II showing shear steps; (<b>h</b>) interactions of shear bands and slip bands due to severe plastic deformation.</p>
Full article ">Figure 6
<p>TEM micrographs in low magnification of the tension-deformed Ti-based BMGMCs: (<b>a</b>) deformed microstructures at low magnification; (<b>b</b>) SAED pattern from amorphous phase; (<b>c</b>) SAED pattern from dendritic crystalline phase; (<b>d</b>) slip step formed in the interface between dendrites and glass matrix.</p>
Full article ">Figure 7
<p>(<b>a</b>) HRTEM images showing slip step formed in the interface between the amorphous phase and dendritic phase; (<b>b</b>) enlarged area from the white square in (<b>a</b>,<b>c</b>) FFTs of red rectangle regions in (<b>b</b>), which follow the directions indicated by the arrows; (<b>d</b>) the curves of true stress vs. time and true strain vs. time; (<b>e</b>) the plot of shear strain rate <math display="inline"><semantics> <mover accent="true"> <mi>γ</mi> <mo>˙</mo> </mover> </semantics></math> vs. time.</p>
Full article ">Figure 8
<p>(<b>a</b>) HRTEM image of tension-fractured Ti-based BMGMCs showing microstructural characteristics in the crystalline region close to slip step; (<b>b</b>) enlarged image of the red square in (<b>a</b>), which shows atomic microstructure of the dendrite phase close to slip step; (<b>c</b>) FFTs of Zone I, Zone II and Zone III indicated by red squares in (<b>b</b>), respectively.</p>
Full article ">Figure 9
<p>(<b>a</b>) An HRTEM image of the region showing a BCC to HCP transition and inset shows a corresponding FFT pattern; (<b>b</b>) IFFT of (<b>a</b>) shows a clear transition from BCC to HCP; (<b>c</b>) a schematic diagram of the transition from BCC to HCP.</p>
Full article ">Figure 10
<p>Schematic illustrations of the microstructural evolution for Ti-based BMGMCs during different tension deformation stages: (<b>a</b>,<b>b</b>) elastic stage, during which dendrites deformed initially, leading to interfacial stress concentration; (<b>c</b>–<b>f</b>) plastic stage, both dendrites and glassy matrix deformed, shear bands formed and propagated, leading to the final failure.</p>
Full article ">
23 pages, 5813 KiB  
Article
Raman Spectroscopy and Electrical Transport in 30Li2O• (67−x) B2O3•(x) SiO2•3Al2O3 Glasses
by Amrit P. Kafle, David McKeown, Winnie Wong-Ng, Meznh Alsubaie, Manar Alenezi, Ian L. Pegg and Biprodas Dutta
Electron. Mater. 2024, 5(3), 166-188; https://doi.org/10.3390/electronicmat5030012 - 12 Sep 2024
Viewed by 306
Abstract
We have investigated the influence of the relative proportions of glass formers in a series of lithium alumino-borosilicate glasses with respect to electrical conductivity (σ) and glass transition temperature (Tg) as functions of glass structure, as determined using Raman spectroscopy. [...] Read more.
We have investigated the influence of the relative proportions of glass formers in a series of lithium alumino-borosilicate glasses with respect to electrical conductivity (σ) and glass transition temperature (Tg) as functions of glass structure, as determined using Raman spectroscopy. The ternary lithium alumino-borate glass exhibits the highest σ and lowest Tg among all the compositions of the glass series, 30Li2O•3Al2O3• (67−x) B2O3xSiO2. However, as B2O3 is replaced by SiO2, a shallow minimum in σ, as well as a shallow maximum in Tg, are observed near x = 27, where the Raman spectra indicate that isolated diborate/tetraborate/orthoborate groups are being progressively replaced by danburite/reedmergnerite-like borosilicate network units. Overall, as the glasses become silica-rich, σ is minimized, while Tg is maximized. In general, these findings show correlations among Tg (sensitive to network polymerization), σ (proportional to ionic mobility), and the different borate and silicate glass structural units as determined using Raman spectroscopy. Full article
Show Figures

Figure 1

Figure 1
<p>Li<sub>2</sub>O-B<sub>2</sub>O<sub>3</sub>-SiO<sub>2</sub> ternary diagram with the LABS series glass compositions (green squares on red line), with some R (pink) and K (blue) ratio values displayed.</p>
Full article ">Figure 2
<p>DTA plots, arrow showing <span class="html-italic">T</span><sub>g</sub>, for some glasses of the series. Endothermic changes are shown as a drop in heat flow.</p>
Full article ">Figure 3
<p>Electrical conductivity (as log σ) with respect to LABS glass compositions, with guiding lines to aid visualization.</p>
Full article ">Figure 4
<p>Arrhenius plots with error bars showing variation in log σ in S/cm, expressed as units of electrical conductivity, along with the inverse of temperature in K<sup>−1</sup> (legends are based on in the order of decrease in conductivity from top to bottom, not on composition across the glass series).</p>
Full article ">Figure 5
<p>Glass transition temperature (<span class="html-italic">T</span><sub>g</sub>) and activation energy from 50 °C to 170 °C with respect to glass composition.</p>
Full article ">Figure 6
<p>Parallel polarized reduced Raman spectra for the glass series. Plots are offset for clarity.</p>
Full article ">Figure 7
<p>(<b>a</b>) Simple polyborate species, (<b>b</b>) complex polyborate species, (<b>c</b>) silicate species, and (<b>d</b>) borosilicate species.</p>
Full article ">Figure 8
<p>(<b>a</b>) Parallel polarized (blue) and cross-polarized (red) reduced Raman spectra for 67B glass. Some vibrational assignments with corresponding structural units are indicated. (<b>b</b>) Parallel polarized Raman spectrum of 67B glass and associated Gaussian component fit by the program IGOR [<a href="#B65-electronicmat-05-00012" class="html-bibr">65</a>]. The data are black points, the fitted individual Gaussian components are plotted in blue, and the sum of all Gaussian-fitted components are plotted in red. Residual intensities are the differences between observed data and the sum of the Gaussian components. Structural unit assignments are indicated for some of the major fitted Gaussian components.</p>
Full article ">Figure 9
<p>(<b>a</b>) Parallel polarized (blue) and cross-polarized (red) reduced Raman spectra of 30B37Si glass. Some vibrational assignments with corresponding structural units are indicated. (<b>b</b>) Parallel polarized Raman spectrum of 30B37Si glass and associated Gaussian component fit by the program IGOR [<a href="#B65-electronicmat-05-00012" class="html-bibr">65</a>]. Conventions from <a href="#electronicmat-05-00012-f008" class="html-fig">Figure 8</a>b are followed.</p>
Full article ">Figure 10
<p>(<b>a</b>) Parallel polarized (blue) and cross-polarized (red) reduced Raman spectra of 67Si glass. Some vibrational assignments with corresponding structural units are indicated. (<b>b</b>) Parallel polarized Raman spectrum of 67Si glass and associated Gaussian component fit by the program IGOR [<a href="#B65-electronicmat-05-00012" class="html-bibr">65</a>]. Conventions from <a href="#electronicmat-05-00012-f008" class="html-fig">Figure 8</a>b are followed.</p>
Full article ">Figure 11
<p>Major Gaussian component areas versus glass composition for important Raman spectral features. Danburite ring assignments are indicated for 30B37Si and 20B47Si glasses only. Each plotted Gaussian component is labeled with its corresponding glass structural unit assignment.</p>
Full article ">Figure 12
<p>Parallel polarized reduced Raman spectra of 67B, 30B37Si, and 67Si glass with major structural unit assignments.</p>
Full article ">
16 pages, 2608 KiB  
Article
Size Effects in Climatic Aging of Epoxy Basalt Fiber Reinforcement Bar
by Anna A. Gavrilieva, Oleg V. Startsev, Mikhail P. Lebedev, Anatoly S. Krotov, Anatoly K. Kychkin and Irina G. Lukachevskaya
Polymers 2024, 16(18), 2550; https://doi.org/10.3390/polym16182550 - 10 Sep 2024
Viewed by 257
Abstract
The purpose of this study was to obtain information on the influence of the size factor on the climatic aging of circular fiber plastics produced by pultrusion. The kinetics of moisture transfer was obtained in humidification and drying modes at 60 °C in [...] Read more.
The purpose of this study was to obtain information on the influence of the size factor on the climatic aging of circular fiber plastics produced by pultrusion. The kinetics of moisture transfer was obtained in humidification and drying modes at 60 °C in samples of epoxy basalt fiber reinforcement bars: after 28 months of exposure in the extremely cold climate of Yakutsk and 30 months of exposure in the moderately warm climate of Gelendzhik. It was shown that the 2D Langmuir model adequately describes the kinetics. The diffusion coefficients in the reinforcement direction for bars with diameters of 6, 8, 10, 16 and 20 mm turned out to be significantly higher than in the radial direction. To clarify the aging mechanism of the bars and the tensile, compressive and bending strength, the coefficient of linear thermal expansion and the glass transition temperature of the epoxy matrix of the bars with a diameter of 6, 8 and 10 mm after 51 months of exposure in Yakutsk and 54 months of exposure in Gelendzhik were measured. It was shown that after climatic exposure, the deformability of the bars decreased with increasing diameter of the bar; the glass transition temperature increased more significantly in the bar with a smaller diameter. In 6 mm diameter bars, the compressive and bending strength limits decreased by 10–25 % due to the plasticizing effect of moisture. With the same depth of moisture penetration into the volume of the samples, its effect on the strength of thin bars was significant, and for thick bars, it was insignificant. An increase in the glass transition temperature by 6 °C, associated with the additional curing of the polymer matrix, occurred in the surface layer of the epoxy basalt fiber reinforcement bars and was revealed in bars with a smaller diameter. Full article
(This article belongs to the Special Issue Advanced Epoxy-Based Materials, 5th Edition)
Show Figures

Figure 1

Figure 1
<p>A set of BFRP bar samples with a nominal diameter 6, 8, 10, 16 and 20 mm to determine moisture transfer characteristics.</p>
Full article ">Figure 2
<p>Temperature dependences of the relative thermal expansion of three parallel BFRP 6 samples after 54 months of exposure in Yakutsk.</p>
Full article ">Figure 3
<p>Average CLTE values of BFRP bar samples after 54 months of exposure in Yakutsk: (1) BFRP 6; (2) BFRP 8; (3) BFRP 10.</p>
Full article ">Figure 4
<p>(<b>a</b>) A typical example of temperature dependences of the dynamic loss modulus of BFRP samples after 54 month of exposure in Yakutsk; (<b>b</b>) A typical example of temperature dependences of the dynamic loss modulus of BFRP samples after 64 month of exposure in Gelendzhik.</p>
Full article ">Figure 5
<p>Kinetic curves of the absorption–desorption cycle of BFRP bars samples unexposed, after 28 months of exposure in Yakutsk and after 30 months of exposure in Gelendzhic at RH 98% and 60 °C: (<b>a</b>) D = 20 mm, (<b>b</b>) D = 16 mm, (<b>c</b>) D = 10 mm, (<b>d</b>) D = 8 mm, (<b>e</b>) D = 6 mm. Solid curves from Equations (<a href="#FD3-polymers-16-02550" class="html-disp-formula">3</a>) and (<a href="#FD5-polymers-16-02550" class="html-disp-formula">5</a>).</p>
Full article ">Figure 5 Cont.
<p>Kinetic curves of the absorption–desorption cycle of BFRP bars samples unexposed, after 28 months of exposure in Yakutsk and after 30 months of exposure in Gelendzhic at RH 98% and 60 °C: (<b>a</b>) D = 20 mm, (<b>b</b>) D = 16 mm, (<b>c</b>) D = 10 mm, (<b>d</b>) D = 8 mm, (<b>e</b>) D = 6 mm. Solid curves from Equations (<a href="#FD3-polymers-16-02550" class="html-disp-formula">3</a>) and (<a href="#FD5-polymers-16-02550" class="html-disp-formula">5</a>).</p>
Full article ">Figure 6
<p>Moisture equilibrium content during absorption of BFRP bar samples that are (<b>a</b>) not exposed, (<b>b</b>) exposed in Yakutsk for 28 months, (<b>c</b>) exposed in Gelendzhik for 30 months. Moisture equilibrium content during desorption of BFRP bar samples that are (<b>d</b>) not exposed, (<b>e</b>) exposed in Yakutsk for 28 months, (<b>f</b>) exposed in Gelendzhik for 30 months.</p>
Full article ">Figure 7
<p>The proportion of pseudo-equilibrium in the total equilibrium content of moisture during absorption of BFRP bar samples that are (<b>a</b>) not exposed, (<b>b</b>) exposed in Yakutsk for 28 months, (<b>c</b>) exposed in Gelendzhik for 30 months.</p>
Full article ">Figure 8
<p>Comparative analysis of diffusion coefficients during absorption of BFRP bar samples that are (<b>a</b>) not exposed, (<b>b</b>) exposed in Yakutsk for 28 months, (<b>c</b>) exposed in Gelendzhik for 30 months.</p>
Full article ">
16 pages, 2783 KiB  
Article
Development of Solid-State Lithium-Ion Batteries (LIBs) to Increase Ionic Conductivity through Interactions between Solid Electrolytes and Anode and Cathode Electrodes
by Majid Monajjemi and Fatemeh Mollaamin
Energies 2024, 17(18), 4530; https://doi.org/10.3390/en17184530 - 10 Sep 2024
Viewed by 478
Abstract
Although in general ions are not able to migrate in the solid-state position due to rigid skeletal structure, in some solid electrolytes with a low energy barrier and high ionic conductivities, these ion transition can occur. In this work, we considered several solid [...] Read more.
Although in general ions are not able to migrate in the solid-state position due to rigid skeletal structure, in some solid electrolytes with a low energy barrier and high ionic conductivities, these ion transition can occur. In this work, we considered several solid electrolytes including lithium phosphorus oxy-nitride (LIPON), a lithium super-ionic conductor (SILICON), and thio-LISICON. For the fabrication and characterization of the solid electrolyte’s fabrication, we used a single-step ball milling (SSBM) procedure. Through this research on all-solid-state rechargeable lithium-ion batteries, our target is to discuss solving several problems in solid LIBs that have recently escalated due to raised concerns relating to safety hazards such as solvent leakage and the flammability of the liquid electrolytes used for commercial LIBs. Through this research, we tested the conductivity amounts of various substrates containing amorphous glass, SSBM, and glass-ceramic samples. Obviously, the SSBM glass-ceramics increased the conductivity, and we also found that the values for conductivity attained by SSBM were higher than those values for glass-ceramics. Using an SSBM technique, silicon nanoparticles were used as an anode material and it was found that the charge and discharge curves in the battery cell cycled between 0.009 and 1.45 V versus Li+/Li at a current density of 210 mA g−1 at room temperature. Since high resistance causes degradation between the cathode material (LiCoO2) and the solid electrolyte, we added GeS2 and SiS2 to the Li2S-P2S5 system to obtain higher conductivities and better stability of the electrode–electrolyte interface. Full article
(This article belongs to the Section D2: Electrochem: Batteries, Fuel Cells, Capacitors)
Show Figures

Figure 1

Figure 1
<p>XRD of crystals glass showing a general phase diagram of the 75% Li<sub>2</sub>S with %25P<sub>2</sub>S<sub>5</sub>.</p>
Full article ">Figure 2
<p>(<b>A</b>) Anode electrode containing acetylene black. (<b>B</b>) Multi-wall carbon nanotubes in anode. (<b>C</b>) Solid-state lithium battery including titanium.</p>
Full article ">Figure 3
<p>Titanium test with Li<sub>2</sub>S–GeS<sub>2</sub>–P<sub>2</sub>S<sub>5</sub> SSE, (<b>A</b>) solid electrolyte battery, and (<b>B</b>) conductivity tester.</p>
Full article ">Figure 4
<p>XRD patterns for all tested electrolytes, (a) Li<sub>2</sub>S, (b) P<sub>2</sub>S<sub>5</sub>, (c) x = GeS<sub>2</sub>, (d) x = 64, (f) x = 76.</p>
Full article ">Figure 5
<p>(a) Conductivity map for Li<sub>(4–x)</sub>Ge<sub>(1–x)</sub>P<sub>x</sub>S<sub>4</sub>; (b) glass, (c) glass-ceramic, (d) SSBM glass-ceramic.</p>
Full article ">Figure 6
<p>Comparison showing superior performance of MWCNT as a conductive additive for all solid–state lithium batteries over acetylene black in different voltages and three temperatures.</p>
Full article ">Figure 7
<p>Charge–discharge curves of all-solid-state cell fabricated with n-Si as anode material by different voltage ranges. (<b>A</b>–<b>F</b>) compare the cycling performance of solid state and liquid electrolyte systems using nano-Si. (<b>G</b>,<b>H</b>) exhibit the charge-discharge performance of ion cells, tested with lithium phosphorus oxy-nitride (LIPON) and lithium super-ionic conductor (SILICON) solid electrolyte, respectively.</p>
Full article ">Figure 7 Cont.
<p>Charge–discharge curves of all-solid-state cell fabricated with n-Si as anode material by different voltage ranges. (<b>A</b>–<b>F</b>) compare the cycling performance of solid state and liquid electrolyte systems using nano-Si. (<b>G</b>,<b>H</b>) exhibit the charge-discharge performance of ion cells, tested with lithium phosphorus oxy-nitride (LIPON) and lithium super-ionic conductor (SILICON) solid electrolyte, respectively.</p>
Full article ">Scheme 1
<p>Operating LIBs.</p>
Full article ">
18 pages, 9602 KiB  
Article
Investigation on the Curing and Thermal Properties of Epoxy/Amine/Phthalonitrile Blend
by Cong Peng, Tao Luo, Zhanjun Wu and Shichao Li
Materials 2024, 17(17), 4411; https://doi.org/10.3390/ma17174411 - 7 Sep 2024
Viewed by 366
Abstract
The bisphenol A-type phthalonitrile (BAPH) was blended with the classic epoxy system E51/DDS to prepare the epoxy/phthalonitrile thermoset. The curing kinetics were investigated by differential scanning calorimetry (DSC) using the isoconversional principle, and the average activation energy (Eα) of the E51/DDS [...] Read more.
The bisphenol A-type phthalonitrile (BAPH) was blended with the classic epoxy system E51/DDS to prepare the epoxy/phthalonitrile thermoset. The curing kinetics were investigated by differential scanning calorimetry (DSC) using the isoconversional principle, and the average activation energy (Eα) of the E51/DDS curing reaction was found to decrease from 87 kJ/mol to 68.6 kJ/mol. Combining the results of the rheological study, the promoting effect of phthalonitrile on the crosslink of epoxy/amine is confirmed. The curing reaction of the blended resin was characterized using FTIR, and the results showed that BAPH could react with DDS. The thermal behaviors of the thermosets were investigated via DMA and TGA. The glass transition temperature (Tg) is found to increase from 181 °C to 195 °C. The char yield increases from 16% to 59.6% at 800 °C in a N2 atmosphere, which is higher than the calculated value based on the proportional principle. The AFM phase images show that there is no phase separation in the cured thermoset. The results imply that the cured epoxy/amine/phthalonitrile blend is probably a kind of copolymer. The real-time TG-MS indicated that the pyrolysis of the thermoset can be divided into two relatively independent stages, which can be assigned to the cleavage of the E51/DDS network, and the phthalocyanine/triazine/isoindoline, respectively. Full article
(This article belongs to the Special Issue Advanced Resin Composites: From Synthesis to Application)
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of the raw materials and products. (<b>a</b>) The BPA and BPPH resin before and after curing.(<b>b</b>) DDS and BAPH/DDS blend before and after curing.</p>
Full article ">Figure 2
<p><sup>1</sup>HNMR (<b>a</b>) and <sup>13</sup>CNMR (<b>b</b>) spectra of the BAPH monomer .</p>
Full article ">Figure 3
<p>Complex viscosity (ƞ*) as a function of (<b>a</b>) temperature and (<b>b</b>) time for EDSX specimens.</p>
Full article ">Figure 4
<p>DSC curves of the curing process of various resin blends (<b>a</b>) and the peak fitting results of the DSC curve of EDPH2 (<b>b</b>).</p>
Full article ">Figure 5
<p>DSC curves of EDPH2 with different heating rates (<b>a</b>–<b>d</b>), the fitted DSC curves of epoxy/DDS curing reaction (<b>e</b>), conversion rate α as function of temperature (<b>f</b>), linear fitting plots at various conversion rates (<b>g</b>), and the values of E<sub>α</sub> depending on α for the crosslink reaction (<b>h</b>).</p>
Full article ">Figure 6
<p>TG curves (<b>a</b>,<b>c</b>) and DTG curves (<b>b</b>,<b>d</b>) of the E51/DDS/BAPH thermosets in N<sub>2</sub> and air atmospheres.</p>
Full article ">Figure 7
<p>Photographs of char residues before (<b>a</b>) and after (<b>b</b>) decomposition in tube furnace under 800 °C, N<sub>2</sub> atmosphere.</p>
Full article ">Figure 8
<p>FTIR spectra of the pyrolysis gas products in N<sub>2</sub> atmosphere for EDPH2 (<b>a</b>), the FTIR absorbance intensity of CH<sub>4</sub>, and aromatic compounds (<b>b</b>).</p>
Full article ">Figure 9
<p>The MS spectra of pyrolysis products of BAPH2 at 380 °C (<b>a</b>), 440 °C (<b>b</b>), and 670 °C (<b>c</b>).</p>
Full article ">Figure 10
<p>DMA results of the cured thermosets.</p>
Full article ">Figure 11
<p>AFM height images (<b>a</b>,<b>b</b>) and phase images (<b>c</b>,<b>d</b>) of the cured ED and EDPH2. The scan size is 10 μm × 10 μm for all images.</p>
Full article ">Scheme 1
<p>The synthesis route of BAPH and the molecular formula of the main stuff. <sup>1</sup>HNMR (400 MHz, DMSO-d6), δ (ppm): 1.7 (s, 3H), 7.12 to 7.14 (d, 2H), 7.36 to 7.38 (d, 2H), 7.36 to 7.37 (d, 1H), 7.78 (d, 1H), 8.09 to 8.10 (d, 1H). <sup>13</sup>CNMR (400 MHz, DMSO-d6), δ (ppm): 161.53, 152.12, 147.92, 129.17, 123.14, 122.36, 120.28, 117.16, 116.38, 115.87, 42.61, 31.03.</p>
Full article ">Scheme 2
<p>Possible curing mechanism of phthalonitrile-containing amine.</p>
Full article ">Scheme 3
<p>The pyrolysis process of EDPH2.</p>
Full article ">
18 pages, 8055 KiB  
Article
Study on the Factors Affecting the Self-Healing Performance of Graphene-Modified Asphalt Based on Molecular Dynamics Simulation
by Fei Guo, Xiaoyu Li, Ziran Wang, Yijun Chen and Jinchao Yue
Polymers 2024, 16(17), 2482; https://doi.org/10.3390/polym16172482 - 30 Aug 2024
Viewed by 360
Abstract
To comprehensively understand the impact of various environmental factors on the self-healing process of graphene-modified asphalt, this study employs molecular dynamics simulation methods to investigate the effects of aging degree (unaged, short-term aged, long-term aged), asphalt type (base asphalt, graphene-modified asphalt), healing temperature [...] Read more.
To comprehensively understand the impact of various environmental factors on the self-healing process of graphene-modified asphalt, this study employs molecular dynamics simulation methods to investigate the effects of aging degree (unaged, short-term aged, long-term aged), asphalt type (base asphalt, graphene-modified asphalt), healing temperature (20 °C, 25 °C, 30 °C), and damage degree (5 Å, 10 Å, 15 Å) on the self-healing performance of asphalt. The validity of the established asphalt molecular models was verified based on four physical quantities: density, radial distribution function analysis, glass transition temperature, and cohesive energy density. The simulated healing time for the asphalt crack model was set to 200 ps. The following conclusions were drawn based on the changes in density, mean square displacement, and diffusion coefficient during the simulated healing process under different influencing factors: Dehydrogenation and oxidation of asphalt molecules during the aging process hinder molecular migration within the asphalt crack model, resulting in poorer self-healing performance. As the service life increases, the decline in the healing performance of graphene-modified asphalt is slower than that of base asphalt, indicating that graphene-modified asphalt has stronger anti-aging properties. When the vacuum layer in the asphalt crack model is small, the changes in the diffusion coefficient are less pronounced. As the crack width increases, the influence of various factors on the diffusion coefficient of the asphalt crack model becomes more significant. When the crack width is large, the self-healing effect of asphalt is more dependent on these influencing factors. Damage degree and oxidative aging have a more significant impact on the healing ability of graphene-modified asphalt than healing temperature. Full article
(This article belongs to the Special Issue Simulation and Calculation of Polymer Composite Materials)
Show Figures

Figure 1

Figure 1
<p>Asphalt molecular structure.</p>
Full article ">Figure 2
<p>Asphalt molecular model (grey for carbon atoms, white for hydrogen atoms, yellow for sulfur atoms, red for oxygen atoms, and blue for nitrogen atoms).</p>
Full article ">Figure 3
<p>Oxygen-containing functional group model (gray for carbon atoms, white for hydrogen atoms, yellow for sulfur atoms, and red for oxygen atoms).</p>
Full article ">Figure 4
<p>Short-term aged asphalt molecular model (grey for carbon atoms, white for hydrogen atoms, yellow for sulfur atoms, red for oxygen atoms, and blue for nitrogen atoms).</p>
Full article ">Figure 5
<p>Long-term aged asphalt molecular model (grey for carbon atoms, white for hydrogen atoms, yellow for sulfur atoms, red for oxygen atoms, and blue for nitrogen atoms).</p>
Full article ">Figure 6
<p>Graphene molecule.</p>
Full article ">Figure 7
<p>Radial distribution function of the asphalt molecular model.</p>
Full article ">Figure 8
<p>The relation of specific volume of asphalt model with temperature.</p>
Full article ">Figure 9
<p>Base asphalt crack mode.</p>
Full article ">Figure 10
<p>Density of asphalt molecular models with different crack widths: (<b>a</b>) 5 Å; (<b>b</b>) 10 Å; (<b>c</b>) 15 Å.</p>
Full article ">Figure 11
<p>Density of matrix asphalt molecular models at different temperatures (the dashed lines indicate the boundary between the two stages of the asphalt crack model).</p>
Full article ">Figure 12
<p>Mean square displacement of asphalt molecular models at different temperatures.</p>
Full article ">Figure 13
<p>The MSD of asphalt molecular models with different crack widths at 25 °C is analyzed.</p>
Full article ">
10 pages, 3459 KiB  
Article
Prediction of Glass Transition Temperature of Polymers Using Simple Machine Learning
by Jaka Fajar Fatriansyah, Baiq Diffa Pakarti Linuwih, Yossi Andreano, Intan Septia Sari, Andreas Federico, Muhammad Anis, Siti Norasmah Surip and Mariatti Jaafar
Polymers 2024, 16(17), 2464; https://doi.org/10.3390/polym16172464 - 29 Aug 2024
Viewed by 823
Abstract
Polymer materials have garnered significant attention due to their exceptional mechanical properties and diverse industrial applications. Understanding the glass transition temperature (Tg) of polymers is critical to prevent operational failures at specific temperatures. Traditional methods for measuring Tg, [...] Read more.
Polymer materials have garnered significant attention due to their exceptional mechanical properties and diverse industrial applications. Understanding the glass transition temperature (Tg) of polymers is critical to prevent operational failures at specific temperatures. Traditional methods for measuring Tg, such as differential scanning calorimetry (DSC) and dynamic mechanical analysis, while accurate, are often time-consuming, costly, and susceptible to inaccuracies due to random and uncertain factors. To address these limitations, the aim of the present study is to investigate the potential of Simplified Molecular Input Line Entry System (SMILES) as descriptors in simple machine learning models to predict Tg efficiently and reliably. Five models were utilized: k-nearest neighbors (KNNs), support vector regression (SVR), extreme gradient boosting (XGBoost), artificial neural network (ANN), and recurrent neural network (RNN). SMILES descriptors were converted into numerical data using either One Hot Encoding (OHE) or Natural Language Processing (NLP). The study found that SMILES inputs with fewer than 200 characters were inadequate for accurately describing compound structures, while inputs exceeding 200 characters diminished model performance due to the curse of dimensionality. The ANN model achieved the highest R2 value of 0.79; however, the XGB model, with an R2 value of 0.774, exhibited the highest stability and shorter training times compared to other models, making it the preferred choice for Tg prediction. The efficiency of the OHE method over NLP was demonstrated by faster training times across the KNN, SVR, XGB, and ANN models. Validation of new polymer data showed the XGB model’s robustness, with an average prediction deviation of 9.76 from actual Tg values. These findings underscore the importance of optimizing SMILES conversion methods and model parameters to enhance prediction reliability. Future research should focus on improving model accuracy and generalizability by incorporating additional features and advanced techniques. This study contributes to the development of efficient and reliable predictive models for polymer properties, facilitating the design and application of new polymer materials. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of the relationship between SMILES character length and model performance.</p>
Full article ">Figure 2
<p>Performance and data distribution of the KNN (<b>a</b>), SVR (<b>b</b>), XGBoost (<b>c</b>), ANN (<b>d</b>), and RNN (<b>e</b>) models trained with the optimal parameters. The blue dots represent the testing data points, while the red line denotes a condition where the true value is the same as predicted value.</p>
Full article ">Figure 2 Cont.
<p>Performance and data distribution of the KNN (<b>a</b>), SVR (<b>b</b>), XGBoost (<b>c</b>), ANN (<b>d</b>), and RNN (<b>e</b>) models trained with the optimal parameters. The blue dots represent the testing data points, while the red line denotes a condition where the true value is the same as predicted value.</p>
Full article ">Figure 3
<p>Comparison of model performance stability over 10 iterations using K-fold method.</p>
Full article ">Figure 4
<p>The <span class="html-italic">T<sub>g</sub></span> polymer data distribution.</p>
Full article ">
12 pages, 16643 KiB  
Article
Structural Relaxation and Delayed Yielding in Cyclically Sheared Cu-Zr Metallic Glasses
by Nikolai V. Priezjev
Metals 2024, 14(9), 984; https://doi.org/10.3390/met14090984 - 29 Aug 2024
Viewed by 319
Abstract
The yielding transition, structural relaxation, and mechanical properties of metallic glasses subjected to repeated loading are examined using molecular dynamics simulations. We consider a poorly annealed Cu-Zr amorphous alloy periodically deformed in a wide range of strain amplitudes at room temperature. It is [...] Read more.
The yielding transition, structural relaxation, and mechanical properties of metallic glasses subjected to repeated loading are examined using molecular dynamics simulations. We consider a poorly annealed Cu-Zr amorphous alloy periodically deformed in a wide range of strain amplitudes at room temperature. It is found that low-amplitude cyclic loading leads to a logarithmic decay of the potential energy, and lower energy states are attained when the strain amplitude approaches a critical point from below. Moreover, the potential energy after several thousand loading cycles is a linear function of the peak value of the stress overshoot during startup continuous shear deformation of the annealed sample. We show that the process of structural relaxation involves collective, irreversible rearrangements of groups of atoms whose spatial extent is most pronounced at the initial stage of loading and at higher strain amplitudes. At the critical amplitude, the glass becomes mechanically annealed for a number of transient cycles and then yields via the formation of a shear band. The yielding transition is clearly marked by abrupt changes in the potential energy, storage modulus, and fraction of atoms with large nonaffine displacements. Full article
Show Figures

Figure 1

Figure 1
<p>The potential energy at the end of each cycle versus the cycle number for the indicated strain amplitudes (<math display="inline"><semantics> <msub> <mi>γ</mi> <mn>0</mn> </msub> </semantics></math>). The period of oscillation is <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>1.0</mn> <mspace width="0.166667em"/> <mi>ns</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>A linear-log plot of the same data as in <a href="#metals-14-00984-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>The variation of potential energy when strain is zero as a function of the number of cycles for strain amplitudes of <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.055</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>0.056</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>0.057</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>0.060</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mn>0.061</mn> </mrow> </semantics></math>. The data for <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.055</mn> </mrow> </semantics></math> are the same as in <a href="#metals-14-00984-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 4
<p>The storage modulus (<math display="inline"><semantics> <msup> <mi>G</mi> <mo>′</mo> </msup> </semantics></math>, in units of GPa) as a function of the cycle number for the indicated values of the strain amplitude (<math display="inline"><semantics> <msub> <mi>γ</mi> <mn>0</mn> </msub> </semantics></math>). The period of oscillation is <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>1.0</mn> <mspace width="0.166667em"/> <mi>ns</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Shear stress (<math display="inline"><semantics> <msub> <mi>σ</mi> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </msub> </semantics></math>, in units of GPa), versus strain for glasses deformed at a constant strain rate of <math display="inline"><semantics> <mrow> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> <mspace width="0.166667em"/> <msup> <mi>ps</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math> after 4000 cycles at the indicated <math display="inline"><semantics> <msub> <mi>γ</mi> <mn>0</mn> </msub> </semantics></math>. The lower violet curve represents data before cyclic loading.</p>
Full article ">Figure 6
<p>The shear modulus (<span class="html-italic">G</span>, in GPa) and the peak value of the stress overshoot (<math display="inline"><semantics> <msub> <mi>σ</mi> <mi>Y</mi> </msub> </semantics></math>, in GPa) as functions of the strain amplitude (<math display="inline"><semantics> <msub> <mi>γ</mi> <mn>0</mn> </msub> </semantics></math>) and the potential energy after 4000 loading cycles (<span class="html-italic">U</span>). The horizontal dashed lines indicate <span class="html-italic">G</span> and <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>Y</mi> </msub> </semantics></math> for the steadily sheared glass before cyclic loading was applied. The straight red line in the panel (<b>b</b>) is the best fit to the data (<math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math> = 0.94).</p>
Full article ">Figure 7
<p>The storage modulus (<math display="inline"><semantics> <msup> <mi>G</mi> <mo>′</mo> </msup> </semantics></math> in units of GPa) versus the cycle number for strain amplitudes <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>0</mn> </msub> <mo>⩾</mo> <mn>0.055</mn> </mrow> </semantics></math>. The data for <math display="inline"><semantics> <msup> <mi>G</mi> <mo>′</mo> </msup> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.055</mn> </mrow> </semantics></math> are the same as in <a href="#metals-14-00984-f004" class="html-fig">Figure 4</a>. The oscillation period is <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>1.0</mn> <mspace width="0.166667em"/> <mi>ns</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>The average of the nonaffine quantity (<math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>[</mo> <mrow> <mo>(</mo> <mi>n</mi> <mo>−</mo> <mn>1</mn> <mo>)</mo> </mrow> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>]</mo> </mrow> </mrow> </semantics></math>, in units of <math display="inline"><semantics> <msup> <mo>Å</mo> <mn>2</mn> </msup> </semantics></math>) as a function of the number of cycles for the indicated strain amplitudes. The oscillation period is <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>1.0</mn> <mspace width="0.166667em"/> <mi>ns</mi> </mrow> </semantics></math>. The inset shows the fraction of atoms with <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>[</mo> <mrow> <mo>(</mo> <mi>n</mi> <mo>−</mo> <mn>1</mn> <mo>)</mo> </mrow> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>]</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math> versus the cycle number for the same strain amplitudes.</p>
Full article ">Figure 9
<p>The average of <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>[</mo> <mrow> <mo>(</mo> <mi>n</mi> <mo>−</mo> <mn>1</mn> <mo>)</mo> </mrow> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>]</mo> </mrow> </mrow> </semantics></math> (in units of <math display="inline"><semantics> <msup> <mo>Å</mo> <mn>2</mn> </msup> </semantics></math>) versus the number of cycles for strain amplitudes of <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>0</mn> </msub> <mo>⩾</mo> <mn>0.055</mn> </mrow> </semantics></math>. The data for the strain amplitude of <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.055</mn> </mrow> </semantics></math> (the black curve) are the same as in <a href="#metals-14-00984-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 10
<p>The fraction of atoms with large nonaffine displacements (<math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>[</mo> <mrow> <mo>(</mo> <mi>n</mi> <mo>−</mo> <mn>1</mn> <mo>)</mo> </mrow> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>]</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>) as a function of the number of cycles for the tabulated strain amplitudes. The oscillation period is <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>1.0</mn> <mspace width="0.166667em"/> <mi>ns</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Selected configurations of atoms in <math display="inline"><semantics> <mrow> <msub> <mi>Cu</mi> <mn>50</mn> </msub> <msub> <mi>Zr</mi> <mn>50</mn> </msub> </mrow> </semantics></math> glass loaded at a strain amplitude of <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.055</mn> </mrow> </semantics></math>. The nonaffine quantity is (<b>a</b>) <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mn>50</mn> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mn>500</mn> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mn>1000</mn> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>, and (<b>d</b>) <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mn>3000</mn> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>. Cu and Zr atoms are not drawn to scale. The legend in panel (<b>d</b>) indicates the magnitude of the nonaffine quantity.</p>
Full article ">Figure 12
<p>Snapshots of atomic configurations of the binary glass periodically deformed at the critical strain amplitude of <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.056</mn> </mrow> </semantics></math>. The nonaffine measure is (<b>a</b>) <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mn>300</mn> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mn>600</mn> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mn>660</mn> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>, and (<b>d</b>) <math display="inline"><semantics> <mrow> <msup> <mi>D</mi> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mn>800</mn> <mspace width="0.166667em"/> <mi>T</mi> <mo>,</mo> <mi>T</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <mn>0.49</mn> <mspace width="0.166667em"/> <msup> <mo>Å</mo> <mn>2</mn> </msup> </mrow> </semantics></math>. Atoms are not depicted to scale. The color code is the same as in <a href="#metals-14-00984-f011" class="html-fig">Figure 11</a>.</p>
Full article ">
17 pages, 7389 KiB  
Article
Enhancing Polylactic Acid Properties with Graphene Nanoplatelets and Carbon Black Nanoparticles: A Study of the Electrical and Mechanical Characterization of 3D-Printed and Injection-Molded Samples
by Salvador Giner-Grau, Carlos Lazaro-Hdez, Javier Pascual, Octavio Fenollar and Teodomiro Boronat
Polymers 2024, 16(17), 2449; https://doi.org/10.3390/polym16172449 - 29 Aug 2024
Viewed by 386
Abstract
This study investigates the enhancement of polylactic acid (PLA) properties through the incorporation of graphene nanoplatelets (GNPs) and carbon black (CB) for applications in 3D printing and injection molding. The research reveals that GNPs and CB improve the electrical conductivity of PLA, although [...] Read more.
This study investigates the enhancement of polylactic acid (PLA) properties through the incorporation of graphene nanoplatelets (GNPs) and carbon black (CB) for applications in 3D printing and injection molding. The research reveals that GNPs and CB improve the electrical conductivity of PLA, although conductivity remains within the insulating range, even with up to 10% wt of nanoadditives. Mechanical characterization shows that nanoparticle addition decreases tensile strength due to stress concentration effects, while dispersants like polyethylene glycol enhance ductility and flexibility. This study compares the properties of materials processed by injection molding and 3D printing, noting that injection molding yields isotropic properties, resulting in better mechanical properties. Thermal analysis indicates that GNPs and CB influence the crystallization behavior of PLA with small changes in the melting behavior. Dynamic Mechanical Thermal Analysis (DMTA) results show how the glass transition temperature and crystallization behavior fluctuate. Overall, the incorporation of nanoadditives into PLA holds potential for enhanced performance in specific applications, though achieving optimal conductivity, mechanical strength, and thermal properties requires careful optimization of nanoparticle type, concentration, and dispersion methods. Full article
(This article belongs to the Special Issue Additive Manufacturing of (Bio)Polymeric Materials, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Conductivity test of PLA samples by injection and additive manufacturing in terms of resistance (R). (<b>a</b>) Conductivity of 3D printed samples. (<b>b</b>) Conductivity of injected samples.</p>
Full article ">Figure 2
<p>Mechanical properties in terms of maximum tensile strength (σ<sub>m</sub>), elongation at break (ε<sub>b</sub>) and tensile modulus (E<sub>t</sub>), shore D hardness (H<sub>D</sub>), and impact strength (a<sub>cU</sub>). (<b>a</b>) Tensile strength, elongation at break and tensile modulus of 3D printed samples. (<b>b</b>) Tensile strength, elongation at break and tensile modulus of injected samples. (<b>c</b>) Shore D hardness and impact strength of 3D printed samples. (<b>d</b>) Shore D hardness and impact strength of injected samples.</p>
Full article ">Figure 3
<p>DSC thermograms of the compounded formulations employed: (<b>a</b>) cooling process and (<b>b</b>) second heating process.</p>
Full article ">Figure 4
<p>FESEM images of the injection-molded samples taken at ×500 increments: (<b>a</b>) PLA-I, (<b>b</b>) 5G-I, (<b>c</b>) 10G-I, (<b>d</b>) 5C-I, and (<b>e</b>) 10C-I.</p>
Full article ">Figure 5
<p>FESEM images of the 3D-printed samples taken at ×500 increments: (<b>a</b>) PLA-3D, (<b>b</b>) 5G-3D, (<b>c</b>) 10G-3D, (<b>d</b>) 5C-3D, and (<b>e</b>) 10C-3D.</p>
Full article ">Figure 6
<p>FESEM images of the 3D-printed samples taken at ×5000 increments: (<b>a</b>) 5G-3D, (<b>b</b>) 5G-I, (<b>c</b>) 10G-3D, (<b>d</b>) 10G-I, (<b>e</b>) 5C-3D, (<b>f</b>) 5C-I, (<b>g</b>) 10C-3D, and (<b>h</b>) 10C-I.</p>
Full article ">Figure 7
<p>DMTA curves for the additive-manufactured and injection-molded samples in terms of storage modulus (G’) and damping factor (tan (δ)). (<b>a</b>) the G’ of the 3D printed samples. (<b>b</b>) the G’ of the injected samples. (<b>c</b>) the tan(δ) of the 3D printed samples. (<b>d</b>) the tan(δ) of the injected samples.</p>
Full article ">
16 pages, 8399 KiB  
Article
The Hydrogen Bonding in the Hard Domains of the Siloxane Polyurea Copolymer Elastomers
by Ming Bao, Tianyu Liu, Ying Tao and Xiuyuan Ni
Polymers 2024, 16(17), 2438; https://doi.org/10.3390/polym16172438 - 28 Aug 2024
Viewed by 355
Abstract
For probing the structure–property relationships of the polyurea elastomers, we synthesize the siloxane polyurea copolymer elastomer by using two aminopropyl-terminated polysiloxane monomers with low and high number-average molecular weight (Mn), i.e., L-30D and H-130D. To study the influence of the [...] Read more.
For probing the structure–property relationships of the polyurea elastomers, we synthesize the siloxane polyurea copolymer elastomer by using two aminopropyl-terminated polysiloxane monomers with low and high number-average molecular weight (Mn), i.e., L-30D and H-130D. To study the influence of the copolymer structures on the film properties, these films are analyzed to obtain the tensile performance, UV-vis spectra, cross-sectional topographies, and glass transition temperature (Tg). The two synthetic thermoplastic elastomer films are characterized by transparency, ductility, and the Tg of the hard domains, depending on the reacting compositions. Furthermore, the film elasticity behavior is studied by the strain recovery and cyclic tensile test, and then, the linear fitting of the tensile data is used to describe the film elasticity based on the Mooney–Rivlin model. Moreover, the temperature-dependent infrared (IR) spectra during heating and cooling are conducted to study the strength and recovery rate of the hydrogen bonding, respectively, and their influence on the film performance is further analyzed; the calculated Mn of the hard segment chains is correlated to the macroscopic recovery rate of the hydrogen bonding. These results can add deep insight to the structure–property relationships of the siloxane polyurea copolymer. Full article
(This article belongs to the Special Issue Advances in Functional Rubber and Elastomer Composites II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The FTIR spectra of the siloxane polyurea copolymer using L-30D (<b>a</b>) and H-130D (<b>b</b>).</p>
Full article ">Figure 2
<p>The tensile stress–strain curves of the copolymer films using L-30D (<b>a</b>) and H-130D (<b>b</b>).</p>
Full article ">Figure 3
<p>The loss factor (tan δ) versus temperature of the films using L-30D (<b>a</b>) and H-130D (<b>b</b>).</p>
Full article ">Figure 4
<p>The UV-vis absorption spectra and photos of the films using L-30D (<b>a</b>) and H-130D (<b>b</b>). The characters covered by the films mean Fuand University in Chinese.</p>
Full article ">Figure 5
<p>The cross-sectional topography of SEM for the films using L-30D, the IPDI contents: (<b>a</b>) 11%, (<b>b</b>) 13%, (<b>c</b>) 18%.</p>
Full article ">Figure 6
<p>The cross-sectional topography of SEM for the films using H-130D, the IPDI contents: (<b>a</b>) 11%, (<b>b</b>) 13%, (<b>c</b>) 18%.</p>
Full article ">Figure 7
<p>The cyclic tensile curves of the samples: (<b>a</b>) L-30D-11%, (<b>b</b>) L-30D-18%, (<b>c</b>) H-130D-11%, and (<b>d</b>) H-130D-18%.</p>
Full article ">Figure 8
<p>The scatters and fitted lines (dashes) of [<span class="html-italic">f</span><sup>∗</sup>] versus <span class="html-italic">λ</span><sup>−1</sup> for the copolymer films using H-130D (<b>a</b>) and the films by the oligomers of different <span class="html-italic">M<sub>n</sub></span> (<b>b</b>).</p>
Full article ">Figure 9
<p>The FTIR spectra of the carbonyl during heating for the films using H-130D (<b>a</b>) and L-30D (<b>b</b>) at the IPDI content of 18%. (<b>c</b>) The fitting example of carbonyl peaks in IR spectra for the films. (<b>d</b>) The area percentage of free carbonyl peaks at rising temperatures for these samples.</p>
Full article ">Figure 10
<p>The FTIR spectra of the carbonyls during cooling for the films using H-130D, the IPDI contents: (<b>a</b>) 11%; (<b>b</b>) 13%. The recovery percentage curves (<b>c</b>) and the recovery rate (<b>d</b>) of the carbonyl peaks for the films using various IPDI contents.</p>
Full article ">Figure 11
<p>The FTIR spectra of the carbonyls during cooling for the films using L-30D (<b>a</b>) and H-130D (<b>b</b>). The recovery percentage curves (<b>c</b>) and the recovery rate (<b>d</b>) for the carbonyl peaks of these films.</p>
Full article ">Scheme 1
<p>The schematic representation of the synthesis of siloxane polyurea copolymers using aminopropyl-terminated PDMS.</p>
Full article ">
22 pages, 10345 KiB  
Article
A Precise Prediction of the Chemical and Thermal Shrinkage during Curing of an Epoxy Resin
by Jesper K. Jørgensen, Vincent K. Maes, Lars P. Mikkelsen and Tom L. Andersen
Polymers 2024, 16(17), 2435; https://doi.org/10.3390/polym16172435 - 28 Aug 2024
Viewed by 461
Abstract
A precise prediction of the cure-induced shrinkage of an epoxy resin is performed using a finite element simulation procedure for the material behaviour. A series of experiments investigating the cure shrinkage of the resin system has shown a variation in the measured cure-induced [...] Read more.
A precise prediction of the cure-induced shrinkage of an epoxy resin is performed using a finite element simulation procedure for the material behaviour. A series of experiments investigating the cure shrinkage of the resin system has shown a variation in the measured cure-induced strains. The observed variation results from the thermal history during the pre-cure. A proposed complex thermal expansion model and a conventional chemical shrinkage model are utilised to predict the cure shrinkage observed with finite element simulations. The thermal expansion model is fitted to measured data and considers material effects such as the glass transition temperature and the evolution of the expansion with the degree of cure. The simulations accurately capture the exothermal heat release from the resin and the cure-induced strains across various temperature profiles. The simulations follow the experimentally observed behaviour. The simulation predictions achieve good accuracy with 2–6% discrepancy compared with the experimentally measured shrinkage over a wide range of cure profiles. Demonstrating that the proposed complex thermal expansion model affects the potential to minimise the shrinkage of the studied epoxy resin. A recommendation of material parameters necessary to accurately determine cure shrinkage is listed. These parameters are required to predict cure shrinkage, allow for possible minimisation, and optimise cure profiles for the investigated resin system. Furthermore, in a study where the resin movement is restrained and therefore able to build up residual stresses, these parameters can describe the cure contribution of the residual stresses in a component. Full article
(This article belongs to the Special Issue Modeling and Simulation of Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>The thermal expansion model involving the transition for the difference between the instantaneous cure temperature <span class="html-italic">T</span> and the glass transition temperature, <math display="inline"><semantics> <msub> <mi>T</mi> <mi>g</mi> </msub> </semantics></math>.</p>
Full article ">Figure 2
<p>The experimental setup consists of a thin polymer bag. The specimen size is approx 150 × 150 mm and has an average thickness of 4 mm. A fibre optic sensor with an FBG and thermocouple is placed near the middle of the thickness.</p>
Full article ">Figure 3
<p>Cure experiment <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>70</mn> <mi>S</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math>—strain measured with the optic FBG and temperature developing in the oven <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>o</mi> <mi>v</mi> <mi>e</mi> <mi>n</mi> </mrow> </msub> </semantics></math>, the temperature recorded inside the resin <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>r</mi> <mi>e</mi> <mi>s</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </semantics></math> by the thermocouple, as well as the degree of cure <span class="html-italic">X</span> and <math display="inline"><semantics> <msub> <mi>T</mi> <mi>g</mi> </msub> </semantics></math> predicted based on the recorded resin temperature. The blue dotted lines represent zero strain. the green dotted line represents a level of 100% cure.</p>
Full article ">Figure 4
<p>The degree of cure at the load transfer point, <math display="inline"><semantics> <msub> <mi>X</mi> <mi>σ</mi> </msub> </semantics></math> recorded for the experiments in <a href="#polymers-16-02435-t002" class="html-table">Table 2</a>. As a function of the difference in temperature between the load transfer point and room temperature, <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>T</mi> </mrow> </semantics></math>. Colour-mapped according to the degree of cure at the end of pre-cure.</p>
Full article ">Figure 5
<p>The final cure-induced strains measured at <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>r</mi> <mi>o</mi> <mi>o</mi> <mi>m</mi> </mrow> </msub> </semantics></math> as function of <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>T</mi> </mrow> </semantics></math> and colour-mapped according to <math display="inline"><semantics> <msub> <mi>X</mi> <mrow> <mi>p</mi> <mi>c</mi> <mi>e</mi> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 6
<p>The relationship between <math display="inline"><semantics> <msubsup> <mi>ε</mi> <mrow> <mi>C</mi> <mi>I</mi> </mrow> <mrow> <msup> <mn>21</mn> <mo>∘</mo> </msup> <mi mathvariant="normal">C</mi> </mrow> </msubsup> </semantics></math> and <math display="inline"><semantics> <msub> <mi>X</mi> <mrow> <mi>p</mi> <mi>c</mi> <mi>e</mi> </mrow> </msub> </semantics></math>, demonstrating the effect of pre-cure length on the measured strain.</p>
Full article ">Figure 7
<p>(<b>a</b>) The strain measured for case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>30</mn> <mi>S</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>30</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math>, demonstrating the effect of the length of pre-cure on these similar cases. (<b>b</b>) Strain measured for the cases <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>L</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>L</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> to show the effect pre-cure length.</p>
Full article ">Figure 8
<p>Fitting and extrapolation to determine the volumetric shrinkage based on strain measurements at pre-cure of <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math>. The blue line represents the strain measured, similar to <a href="#polymers-16-02435-f003" class="html-fig">Figure 3</a>. The dashed orange line represents the fitted behaviour. The black dot represents the load transfer point, corresponding to <math display="inline"><semantics> <msub> <mi>X</mi> <mi>σ</mi> </msub> </semantics></math>.</p>
Full article ">Figure 9
<p>Strain measured based on a reheated <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> at both 1 K/min an 3 K/min after fully curing the specimen for 4 h at 100 °C.</p>
Full article ">Figure 10
<p>The thermal expansion evolution found by the derivative of the strain measured in <a href="#polymers-16-02435-f009" class="html-fig">Figure 9</a> for data measured at 1 K/min.</p>
Full article ">Figure 11
<p>Evolution of thermal expansion for <math display="inline"><semantics> <mrow> <msup> <mi>T</mi> <mo>*</mo> </msup> <mo>&gt;</mo> <msubsup> <mi>T</mi> <mn>5</mn> <mo>′</mo> </msubsup> </mrow> </semantics></math> as function of <span class="html-italic">X</span>, for heating and cooldown.</p>
Full article ">Figure 12
<p>A cutout of the experimental setup showing the thermocouple and FBG. It illustrates how the 1D thermomechanical model is built to simulate the cure behaviour of the resin through the thickness.</p>
Full article ">Figure 13
<p>The predicted and measured values of temperature and strain over time for case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>70</mn> <mi>S</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> as well as predicted <span class="html-italic">X</span> and <math display="inline"><semantics> <msub> <mi>T</mi> <mi>g</mi> </msub> </semantics></math> by the simulation. The final values of the predicted and measured strain are shown in the plot.</p>
Full article ">Figure 14
<p>The predicted and measured values of strain as function of temperature for case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>70</mn> <mi>S</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 15
<p>(<b>a</b>) The predicted and measured temperature in the resin as well as oven temperature as a function of the degree of cure. (<b>b</b>) The predicted and measured values of strain as a function of the degree of cure for case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>70</mn> <mi>S</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>Comparison between the cure-induced strain from the experiments and simulations plotted together and linear and second-order tendencies plotted together. Circles indicate the measured shrinkages and the crosses represent the simulated shrinkage.</p>
Full article ">Figure 17
<p>The comparison of pre-cure effects between the experiments with solid lines and simulations with dashed lines. The colours for the cases refer to the same colour bar for <math display="inline"><semantics> <msub> <mi>X</mi> <mrow> <mi>p</mi> <mi>c</mi> <mi>e</mi> </mrow> </msub> </semantics></math> in <a href="#polymers-16-02435-f016" class="html-fig">Figure 16</a>. (<b>a</b>) The experiments and simulations of case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>30</mn> <mi>S</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>30</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math>. (<b>b</b>) The experiments and simulations of case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>L</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>L</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure A1
<p>(<b>a</b>) Case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>40</mn> <mi>L</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>L</mi> </msub> </mrow> </semantics></math> strain over time. (<b>b</b>) Case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>L</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>L</mi> </msub> </mrow> </semantics></math> strain over time.</p>
Full article ">Figure A2
<p>(<b>a</b>) Case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> strain over time. (<b>b</b>) Case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>70</mn> <mi>S</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> strain over time.</p>
Full article ">Figure A3
<p>(<b>a</b>) Case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>60</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> strain over time. (<b>b</b>) Case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>60</mn> <mi>L</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> strain over time.</p>
Full article ">Figure A4
<p>(<b>a</b>) Case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>30</mn> <mi>S</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> strain over time. (<b>b</b>) Case <math display="inline"><semantics> <mrow> <mrow> <mo>[</mo> <msub> <mn>50</mn> <mi>S</mi> </msub> <msub> <mn>30</mn> <mi>M</mi> </msub> <mo>]</mo> </mrow> <msub> <mn>80</mn> <mi>M</mi> </msub> </mrow> </semantics></math> strain over time.</p>
Full article ">
23 pages, 5976 KiB  
Article
Structure–Glass Transition Relationships in Non-Isocyanate Polyhydroxyurethanes
by Konstantinos N. Raftopoulos, Izabela Łukaszewska, Sebastian Lalik, Paulina Zając, Artur Bukowczan, Edyta Hebda, Monika Marzec and Krzysztof Pielichowski
Molecules 2024, 29(17), 4057; https://doi.org/10.3390/molecules29174057 - 27 Aug 2024
Viewed by 474
Abstract
The molecular dynamics, with an emphasis on the calorimetric and dynamic glass transitions, of non-isocyanate polyhydroxyurethanes (PHUs) produced by the equimolar polyaddition of polyether-based dicyclic carbonates (P-CCs) and various short diamines was studied. The diamine component consisted of a short aliphatic diamine (1,4-diaminobutane, [...] Read more.
The molecular dynamics, with an emphasis on the calorimetric and dynamic glass transitions, of non-isocyanate polyhydroxyurethanes (PHUs) produced by the equimolar polyaddition of polyether-based dicyclic carbonates (P-CCs) and various short diamines was studied. The diamine component consisted of a short aliphatic diamine (1,4-diaminobutane, DAB) and a more complex ‘characteristic’ diamine. The study was conducted to investigate (i) the chemical structure of the characteristic amine, (ii) its molar ratio, and (iii) the structure and molar mass of the P-CC. Infrared spectroscopy, differential scanning calorimetry, and broadband dielectric spectroscopy were employed. The P-CC, constituting the bulk of the systems, was the most crucial component for the glass transition. The characteristic amine influenced the glass transition as a result of its bulky structure, but also presumably as a result of the introduction of free volume and the formation of hydrogen bonds. The dynamic glass transition (α relaxation) trace in the Arrhenius plots showed a subtle change at a certain temperature that merits further study in the future. The charge mobility was fully coupled with the molecular mobility, as evidenced by dc conductivity being directly proportional to the characteristic frequency of α relaxation. The fluctuation in carbonyl units (β relaxation) was mildly affected by changes in their immediate environment. Full article
(This article belongs to the Special Issue Macromolecular Chemistry in Europe)
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of all materials under investigation. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to two series and that the scale is different before and after the <span class="html-italic">x</span>-axis break.</p>
Full article ">Figure 2
<p>FTIR spectra of all materials in the amide I region related to the carbonyl of the urethane group. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to two series.</p>
Full article ">Figure 3
<p>Normalized DSC curves for all materials under investigation. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to more than one series. The curves have been translated for clarity.</p>
Full article ">Figure 4
<p>Dielectric spectra of sample S4 in the four different formalisms discussed in this article. Data are shown as a representative case. (<b>a</b>) Real part of the dielectric function <span class="html-italic">ε</span>′; (<b>b</b>) imaginary part of the dielectric function <span class="html-italic">ε</span>″; (<b>c</b>) real part of the complex conductivity <span class="html-italic">σ’</span>; (<b>d</b>) “conductivity free” <span class="html-italic">ε</span>″ approximated according to the methodology of Wübbenhorst and van Turnhout [<a href="#B78-molecules-29-04057" class="html-bibr">78</a>] (details in <a href="#sec4dot5-molecules-29-04057" class="html-sec">Section 4.5</a>). Spectra are shown in steps of 10 K. Spectra at two key temperatures are highlighted, and the arrows show the progress of key features of the spectra.</p>
Full article ">Figure 5
<p>Isochronal plots of <span class="html-italic">ε</span>′ at a high frequency (10 kHz) that are not affected by electrode polarization. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to two series.</p>
Full article ">Figure 6
<p>Comparative dielectric loss spectra at –60 °C, where both secondary relaxations are visible. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to two series. For clarity, only every second point is drawn. The lines are the fits of two Cole–Cole terms.</p>
Full article ">Figure 7
<p>Arrhenius maps for secondary relaxations for all materials under investigation. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to two series.</p>
Full article ">Figure 8
<p>Characteristic parameters of the two secondary relaxations, which were dependent on the structure.</p>
Full article ">Figure 9
<p>(<b>a</b>) Derivative spectra of the representative sample S4 in steps of 10 K, in the temperature region where α relaxation was visible in the experimental window. Lines are fits of the employed model; the details are provided in the text. (<b>b</b>) A detailed example of the fitting of the model on the spectrum recorded at 65 °C.</p>
Full article ">Figure 10
<p>Arrhenius plots for all materials under investigation. Calorimetric glass transition temperatures (<span class="html-italic">T<sub>g</sub></span>) are plotted at an equivalent relaxation time of 100 s. The calorimetric <span class="html-italic">T<sub>g</sub></span> values for P74 and MBCA in panel (<b>a</b>) have been intentionally slightly translated vertically for visibility. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to two series.</p>
Full article ">Figure 11
<p>Angel plots for all materials under investigation. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to two series.</p>
Full article ">Figure 12
<p>Arrhenius plot of dc conductivity for all materials under investigation. (<b>a</b>) Series I, (<b>b</b>) series II, and (<b>c</b>) series III. Note that certain samples belong to two series.</p>
Full article ">Figure 13
<p>Correlation of the time scale of molecular dynamics and dc conductivity for all materials under investigation. In panel (<b>a</b>), dc conductivity is plotted against the maximum frequency and in panel (<b>b</b>) against the characteristic Havriliak–Negami frequency.</p>
Full article ">Figure 14
<p>FTIR spectra of the reaction mixtures at different reaction times for samples (<b>a</b>) S2, (<b>b</b>) S3, (<b>c</b>) S7, and (<b>d</b>) S8 as representative specimens.</p>
Full article ">Scheme 1
<p>Synthesis process. Abbreviations are explained in the text and in <a href="#molecules-29-04057-t001" class="html-table">Table 1</a>.</p>
Full article ">
17 pages, 17882 KiB  
Article
Evaluation of Shape Recovery Performance of Shape Memory Polymers with Carbon-Based Fillers
by Sungwoong Choi, Seongeun Jang, Seung Hwa Yoo, Gyo Woo Lee and Duyoung Choi
Polymers 2024, 16(17), 2425; https://doi.org/10.3390/polym16172425 - 27 Aug 2024
Viewed by 420
Abstract
This study focuses on enhancing the thermal properties and shape recovery performance of shape memory polymers (SMPs) through the application of carbon-based fillers. Single and mixed fillers were used to investigate their effects on the glass transition temperature (Tg), thermal conductivity, [...] Read more.
This study focuses on enhancing the thermal properties and shape recovery performance of shape memory polymers (SMPs) through the application of carbon-based fillers. Single and mixed fillers were used to investigate their effects on the glass transition temperature (Tg), thermal conductivity, and shape recovery performance. The interaction among the three-dimensional (3D) structures of mixed fillers played a crucial role in enhancing the properties of the SMP. These interactions facilitated efficient heat transfer pathways and conserved strain energy. The application of mixed fillers resulted in substantial improvements, demonstrating a remarkable 290.37% increase in thermal conductivity for SMPCs containing 60 μm carbon fiber (CF) 10 wt% + graphite 20 wt% and a 60.99% reduction in shape recovery time for SMPCs containing CF 2.5 wt% + graphite 2.5 wt%. At a content of 15 wt%, a higher graphite content compared to CF improved the thermal conductivity by 37.42% and reduced the shape recovery time by 6.98%. The findings demonstrate that the application of mixed fillers, especially those with high graphite content, is effective in improving the thermal properties and shape recovery performance of SMPs. By using mixed fillers with high graphite content, the performance of the SMP showed significant improvement in situations where fast response times were required. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of SMP.</p>
Full article ">Figure 2
<p>FE-SEM image of the fillers: (<b>a</b>) graphite; (<b>b</b>) 60 μm CF; (<b>c</b>) 100 μm CF; (<b>d</b>) mixed filler.</p>
Full article ">Figure 3
<p>Schematic diagram of SMP specimen preparation process.</p>
Full article ">Figure 4
<p>Schematic diagram of the process for testing the shape recovery performance.</p>
Full article ">Figure 5
<p>Sequential images of SMP shape recovery from fixed to original shape.</p>
Full article ">Figure 6
<p>Schematic of SMP shape recovery performance measurement.</p>
Full article ">Figure 7
<p>DSC results of SMPC according to the filler.</p>
Full article ">Figure 8
<p>Schematic diagram of the heat transfer path by plane, line, and point.</p>
Full article ">Figure 9
<p>Thermal conductivity results of single filler.</p>
Full article ">Figure 10
<p>Thermal conductivity results of mixed filler: (<b>a</b>) 60 μm CF; (<b>b</b>) 100 μm CF.</p>
Full article ">Figure 11
<p>Shape fixation ratio results of single filler.</p>
Full article ">Figure 12
<p>Shape fixation ratio results of mixed filler: (<b>a</b>) 60 μm CF; (<b>b</b>) 100 μm CF.</p>
Full article ">Figure 13
<p>Shape recovery ratio results of single filler.</p>
Full article ">Figure 14
<p>Shape recovery ratio results of mixed filler: (<b>a</b>) 60 μm CF; (<b>b</b>) 100 μm CF.</p>
Full article ">Figure 15
<p>Shape recovery time results of single filler.</p>
Full article ">Figure 16
<p>Shape recovery time results of mixed filler: (<b>a</b>) 60 μm CF; (<b>b</b>) 100 μm CF.</p>
Full article ">
Back to TopTop