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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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
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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (17,929)

Search Parameters:
Keywords = polymeric

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
38 pages, 9286 KiB  
Review
CRISPR-Cas9 Gene Therapy: Non-Viral Delivery and Stimuli-Responsive Nanoformulations
by Hyunwoo Lee, Won-Yeop Rho, Yoon-Hee Kim, Hyejin Chang and Bong-Hyun Jun
Molecules 2025, 30(3), 542; https://doi.org/10.3390/molecules30030542 - 24 Jan 2025
Abstract
The CRISPR-Cas9 technology, one of the groundbreaking genome editing methods for addressing genetic disorders, has emerged as a powerful, precise, and efficient tool. However, its clinical translation remains hindered by challenges in delivery efficiency and targeting specificity. This review provides a comprehensive analysis [...] Read more.
The CRISPR-Cas9 technology, one of the groundbreaking genome editing methods for addressing genetic disorders, has emerged as a powerful, precise, and efficient tool. However, its clinical translation remains hindered by challenges in delivery efficiency and targeting specificity. This review provides a comprehensive analysis of the structural features, advantages, and potential applications of various non-viral and stimuli-responsive systems, examining recent progress to emphasize the potential to address these limitations and advance CRISPR-Cas9 therapeutics. We describe how recent reports emphasize that nonviral vectors, including lipid-based nanoparticles, extracellular vesicles, polymeric nanoparticles, gold nanoparticles, and mesoporous silica nanoparticles, can offer diverse advantages to enhance stability, cellular uptake, and biocompatibility, based on their structures and physio-chemical stability. We also summarize recent progress on stimuli-responsive nanoformulations, a type of non-viral vector, to introduce precision and control in CRISPR-Cas9 delivery. Stimuli-responsive nanoformulations are designed to respond to pH, redox states, and external triggers, facilitate controlled and targeted delivery, and minimize off-target effects. The insights in our review suggest future challenges for clinical applications of gene therapy technologies and highlight the potential of delivery systems to enhance CRISPR-Cas9’s clinical efficacy, positioning them as pivotal tools for future gene-editing therapies. Full article
16 pages, 6489 KiB  
Article
Structural Aspects and Adhesion of Polyurethane Composite Coatings for Surface Acoustic Wave Sensors
by Mauro dos Santos de Carvalho, Michael Rapp, Achim Voigt, Marian Dirschka and Udo Geckle
Coatings 2025, 15(2), 139; https://doi.org/10.3390/coatings15020139 - 24 Jan 2025
Abstract
Surface acoustic wave-based (SAW) sensors are of great interest due to their high sensibility and fast and stable responses. They can be obtained at an overall low cost and with an intuitive and easy-to-use method. The chemical sensitization of a piezoelectric transducer plays [...] Read more.
Surface acoustic wave-based (SAW) sensors are of great interest due to their high sensibility and fast and stable responses. They can be obtained at an overall low cost and with an intuitive and easy-to-use method. The chemical sensitization of a piezoelectric transducer plays a key role in defining the properties of SAW sensors. In this study, we investigate the structural and adhesion properties of a new class of coating material based on polyurethane polymeric composites. We used dark-field microscopy (DFM) and scanning electron microscopy (SEM) to observe the microstructure of polyurethane composite coatings on piezoelectric sensor elements and to analyze the effects of the chemical resistance and adhesion test (CAT) on the coating layers obtained with the polyurethane polymeric composites. The results of the microscopy showed that all polyurethane composite coatings exhibited excellent uniformity and stability after chemical adherence testing (CAT). All of the observations were correlated with the results of the ultrasonic analysis, which demonstrated the role of polyurethane as a binder to form the stable structure of the composites and, at the same time, as an adhesion promoter, increasing the chemical resistance and the adherence of the coating layer to the complex surface of the piezoelectric sensor element. Full article
19 pages, 1223 KiB  
Review
Bibliometric Analysis on Graphitic Carbon Nitride (g-C3N4) as Photocatalyst for the Remediation of Water Polluted with Contaminants of Emerging Concern
by José M. Veiga-del-Baño, Gabriel Pérez-Lucas, Pedro Andreo-Martínez and Simón Navarro
Catalysts 2025, 15(2), 115; https://doi.org/10.3390/catal15020115 - 24 Jan 2025
Abstract
Carbon nitrides are polymeric materials with a broad range of applications, including photocatalysis. Among them, graphitic carbon nitride (g-C3N4), a low-cost material, is an excellent photocatalyst under visible light irradiation owing to its features such as correct band positions, [...] Read more.
Carbon nitrides are polymeric materials with a broad range of applications, including photocatalysis. Among them, graphitic carbon nitride (g-C3N4), a low-cost material, is an excellent photocatalyst under visible light irradiation owing to its features such as correct band positions, high stability and non-toxicity. g-C3N4 is a metal-free material that is easily synthesized by polymerizing nitrogen-rich compounds and is an efficient heterogeneous catalyst for many reaction procedures due to its distinctive electronic structure and the benefits of the mesoporous texture. In addition, in situ or post-modification of g-C3N4 can further improve catalytic performance or expand its application for remediating environmental pollution. Water pollution from organic compounds such as pesticides and pharmaceuticals is increasing dramatically and is becoming a serious problem around the world. These pollutants enter water supplies in a variety of ways, including industrial and hospital wastewater, agricultural runoff, and chemical use. To solve this problem, photocatalysis is a promising technology. Without the use of other oxidative chemicals, g-C3N4 uses renewable solar energy to transform harmful pollutants into harmless products. As a result, much recent research has focused on the photocatalytic activity of g-C3N4 for wastewater treatment. For this reason, the main objective of this paper is to contribute a chronological overview of the bibliometrics on g-C3N4 for the removal of pesticides and pharmaceuticals from water using the tools BibExcel, Bibliometrix and R-Studio IDE. A bibliometric analysis was performed using the Science Citation Index Expanded (WoS©) database to analyze the scientific literature published in the field over the last 10 years. The results were used to identify limitations and guide future research. Full article
15 pages, 3593 KiB  
Article
Enhancing the Fluorescence and Antimicrobial Performance of Carbon Dots via Hypochlorite Treatment
by Spyridon Gavalas, Mohammed S. Beg, Ella N. Gibbons and Antonios Kelarakis
Nanomaterials 2025, 15(3), 184; https://doi.org/10.3390/nano15030184 - 24 Jan 2025
Abstract
This paper presents a simple, post-synthesis treatment of carbon dots (C-dots) that relies on the oxidizing activity of sodium hypochlorite to induce surface oxidation, etching and pronounced structural rearrangements. The thus treated C-dots (ox-C-dots) exhibit up to six-fold enhancement in quantum yield compared [...] Read more.
This paper presents a simple, post-synthesis treatment of carbon dots (C-dots) that relies on the oxidizing activity of sodium hypochlorite to induce surface oxidation, etching and pronounced structural rearrangements. The thus treated C-dots (ox-C-dots) exhibit up to six-fold enhancement in quantum yield compared to non-oxidised analogues, while maintaining low levels of cytotoxicity against HeLa and U87 cell lines. In addition, we demonstrate that a range of polymeric materials (polyurethane sponge, polyvinylidene fluoride membrane, polyester fabric) impregnated with ox-C-dots show advanced antifungal properties against Talaromyces pinophilus, while their untreated counterparts fail to do so. Full article
Show Figures

Figure 1

Figure 1
<p>TEM images and corresponding size distribution histograms (n = 50) of (<b>a</b>) C-dots and (<b>b</b>) ox3-C-dots.</p>
Full article ">Figure 2
<p>Deconvolution of C1s XPS spectra of (<b>a</b>) C-dots and (<b>b</b>) ox3-C-dots.</p>
Full article ">Figure 3
<p>FTIR spectra of C-dots, ox3-C-dots and ox7-C-dots.</p>
Full article ">Figure 4
<p>Zeta potential of 0.05 mg/mL aqueous dispersions of C-dots (black squares) and ox-3-C-dots (red circles).</p>
Full article ">Figure 5
<p>PL spectra of 0.1 mg/mL aqueous dispersions of (<b>a</b>) C-dots, (<b>b</b>) ox1-C-dots, (<b>c</b>) ox2-C-dots, (<b>d</b>) ox3-C-dots and (<b>e</b>) ox7-C-dots at λ<sub>ex</sub> indicated.</p>
Full article ">Figure 6
<p>QY of C-dots and ox-C-dots as a function of the concentration of NaClO used for the treatment of C-dots as detailed in <a href="#nanomaterials-15-00184-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 7
<p>Photos of aqueous dispersions of ox-C-dots under (<b>a</b>) UV radiation and (<b>b</b>) daylight compared to the original (untreated) C-dot dispersion. All photos were taken 24 h following the addition of NaClO.</p>
Full article ">Figure 8
<p>MTT assay test to assess the viability of HeLa (<b>a</b>) and U87 cells (<b>b</b>) following their 24 h incubation with C-dots (black bars) and ox3-C-dots (red bars).</p>
Full article ">Figure 9
<p>Photos of the Petri dishes containing <span class="html-italic">T. Pinophilus</span> cultures in the presence of polyester fabric (<b>a</b>) and PVDF membrane (<b>b</b>) impregnated with (i) water (control), (ii) C-dots and (iii) ox3-C-dots.</p>
Full article ">
13 pages, 12021 KiB  
Article
Production of Monodisperse Oil-in-Water Droplets and Polymeric Microspheres Below 20 μm Using a PDMS-Based Step Emulsification Device
by Naotomo Tottori, Seungman Choi and Takasi Nisisako
Micromachines 2025, 16(2), 132; https://doi.org/10.3390/mi16020132 - 24 Jan 2025
Viewed by 99
Abstract
Step emulsification (SE) is renowned for its robustness in generating monodisperse emulsion droplets at arrayed nozzles. However, few studies have explored poly(dimethylsiloxane) (PDMS)-based SE devices for producing monodisperse oil-in-water (O/W) droplets and polymeric microspheres with diameters below 20 µm—materials with broad applicability. In [...] Read more.
Step emulsification (SE) is renowned for its robustness in generating monodisperse emulsion droplets at arrayed nozzles. However, few studies have explored poly(dimethylsiloxane) (PDMS)-based SE devices for producing monodisperse oil-in-water (O/W) droplets and polymeric microspheres with diameters below 20 µm—materials with broad applicability. In this study, we present a PDMS-based microfluidic SE device designed to achieve this goal. Two devices with 264 nozzles each were fabricated, featuring straight and triangular nozzle configurations, both with a height of 4 µm and a minimum width of 10 µm. The devices were rendered hydrophilic via oxygen plasma treatment. A photocurable acrylate monomer served as the dispersed phase, while an aqueous polyvinyl alcohol solution acted as the continuous phase. The straight nozzles produced polydisperse droplets with diameters exceeding 30 µm and coefficient-of-variation (CV) values above 10%. In contrast, the triangular nozzles, with an opening width of 38 µm, consistently generated monodisperse droplets with diameters below 20 µm, CVs below 4%, and a maximum throughput of 0.5 mL h−1. Off-chip photopolymerization of these droplets yielded monodisperse acrylic microspheres. The low-cost, disposable, and scalable PDMS-based SE device offers significant potential for applications spanning from laboratory-scale research to industrial-scale particle manufacturing. Full article
(This article belongs to the Special Issue Recent Advances in Droplet Microfluidics)
Show Figures

Figure 1

Figure 1
<p>Polydimethylsiloxane (PDMS)-based step-emulsification (SE) devices for generating oil-in-water (O/W) droplets with diameters below 20 µm. (<b>a</b>) Schematic representation of the overall channel layout, showing: (1) a central channel for introducing the dispersed oil phase, (2) two side channels for supplying the continuous aqueous phase and collecting the produced droplets, and (3) two arrays of 132 shallow SE nozzles (264 nozzles in total). (<b>b</b>) Schematic cross-sectional view of the nozzle and channels, highlighting their respective heights. (<b>c</b>) Top-view illustrations of nozzle configurations, depicting straight nozzles with an upstream plateau (left) and triangular nozzles (right) along with their geometric parameters.</p>
Full article ">Figure 2
<p>Scanning electron microscopy (SEM) images of (<b>a</b>) master molds and (<b>b</b>) microchannels replicated in PDMS chips for the SE devices with straight nozzles (<b>left</b>) and triangular nozzles (<b>right</b>). Scale bar: 200 μm.</p>
Full article ">Figure 3
<p>Generation of polydisperse O/W droplets in the straight-nozzle device. (<b>a</b>) Droplet formation at a dispersed phase flow rate (<span class="html-italic">Q</span><sub>d</sub>) of 0.1 mL h<sup>−1</sup> and a continuous phase flow rate (<span class="html-italic">Q</span><sub>c</sub>) of 1.0 mL h<sup>−1</sup>. (<b>b</b>) Droplet formation under the same <span class="html-italic">Q</span><sub>d</sub> but at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup>. Arrows indicate the continuous phase flow direction. Scale bars: 100 μm.</p>
Full article ">Figure 4
<p>O/W droplet generation using the device with triangular nozzles. (<b>a</b>) Optical micrograph of the nozzle arrays operating at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>d</sub> = 0.1 mL h<sup>−1</sup>. Scale bar: 200 µm. (<b>b</b>) Magnified views of the nozzles in (<b>a</b>), demonstrating operation in the ‘small drop’ (SD) mode. (<b>c</b>) Magnified view of the nozzles operating at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>d</sub> = 0.7 mL h<sup>−1</sup>, showing two nozzles on the left operating in the SD mode, while the remaining nozzles operate in the ‘large drop’ (LD) mode. Scale bars: 20 µm.</p>
Full article ">Figure 5
<p>Micrographs and size distributions of O/W droplets collected from the device with triangular nozzles. Droplets were produced at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>d</sub> = (<b>a</b>) 0.1 mL h<sup>−1</sup>, (<b>b</b>) 0.7 mL h<sup>−1</sup>. Scale bars: 20 μm.</p>
Full article ">Figure 6
<p>Evolution of the average droplet diameter (<span class="html-italic">D</span><sub>avg</sub>) and CV values across the SD and LD regimes for O/W droplets generated using the device with triangular nozzles, with <span class="html-italic">Q</span><sub>d</sub> varied from 0.1 to 1.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>c</sub> fixed at 5.0 mL h<sup>−1</sup>.</p>
Full article ">Figure 7
<p>Monodisperse polymeric microspheres obtained through off-chip photopolymerization. (<b>a</b>) SEM images of polymeric microspheres derived from O/W droplets generated by the device with edge-shaped nozzles at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>d</sub> = 0.1 mL h<sup>−1</sup>. (<b>b</b>) Size distribution of the microspheres. Scale bar: 10 µm.</p>
Full article ">
20 pages, 3010 KiB  
Article
Synthesis of Acrylic–Urethane Hybrid Polymer Dispersions and Investigations on Their Properties as Binders in Leather Finishing
by Selime Keskin, Catalina N. Cheaburu-Yilmaz, Aylin Altinisik Tagac, Raluca Nicoleta Darie-Nita and Onur Yilmaz
Polymers 2025, 17(3), 308; https://doi.org/10.3390/polym17030308 - 24 Jan 2025
Viewed by 121
Abstract
This study investigates the synthesis and application of acrylic–urethane hybrid polymer dispersions as advanced binders for leather finishing. Two polymerization techniques—seeded emulsion and miniemulsion—were used to produce hybrid polymer dispersions by varying the ratios of polyurethane (PU) and acrylic (AC). The synthesized dispersions, [...] Read more.
This study investigates the synthesis and application of acrylic–urethane hybrid polymer dispersions as advanced binders for leather finishing. Two polymerization techniques—seeded emulsion and miniemulsion—were used to produce hybrid polymer dispersions by varying the ratios of polyurethane (PU) and acrylic (AC). The synthesized dispersions, i.e., the hybrid polyurethanes, showed stable, uniform particle sizes, inferring good compatibility and interaction between the PU and AC phases, as confirmed by particle sizes, FTIR, and DSC analyses. The performance of the coating on leather surfaces was assessed by using standard physical tests, including rubbing fastness, flexing endurance, water spot resistance, and grain strength. The results showed that the hybrid polymers outperformed their individual PU and AC counterparts, particularly in terms of abrasion resistance and mechanical integrity. Of the two polymerization techniques, the seeded emulsion hybrids exhibited superior coating properties, providing greater resistance to cracking and abrasion under stress, improved grain strength, and better color retention during rubbing tests. These findings highlight the potential of acrylic–urethane hybrids, particularly those prepared via seeded emulsion polymerization, to address the limitations of traditional binders in high-performance leather applications. Full article
(This article belongs to the Section Polymer Composites and Nanocomposites)
Show Figures

Figure 1

Figure 1
<p>Particle size distribution curves of the latexes: (<b>a</b>) neat PU and AC latexes, (<b>b</b>) seeded hybrid latexes (S), and (<b>c</b>) hybrid miniemulsions (M).</p>
Full article ">Figure 2
<p>FTIR spectra of the PU, prepolymer, and PU monomers.</p>
Full article ">Figure 3
<p>FTIR spectra of the PU, AC, and PU–AC hybrid polymers.</p>
Full article ">Figure 4
<p>DSC thermograms of PU, AC, and hybrid polymers.</p>
Full article ">Figure 5
<p>Images of the finished leather samples after the wet rubbing test.</p>
Full article ">Scheme 1
<p>Synthesis of waterborne PU.</p>
Full article ">Scheme 2
<p>Illustration of the PU–AC hybrid synthesis via seeded emulsion polymerization.</p>
Full article ">Scheme 3
<p>Illustration of the PU–AC hybrid synthesis via miniemulsion polymerization.</p>
Full article ">
16 pages, 4204 KiB  
Article
Anti-PTK7 Monoclonal Antibodies Suppresses Oncogenic Phenotypes in Cellular and Xenograft Models of Triple-Negative Breast Cancer
by Min Ho Kim, Mi Kyung Park, Han Na Park, Seung Min Ham, Ho Lee and Seung-Taek Lee
Cells 2025, 14(3), 181; https://doi.org/10.3390/cells14030181 - 24 Jan 2025
Viewed by 128
Abstract
Protein tyrosine kinase 7 (PTK7), a catalytically defective receptor protein tyrosine kinase, is frequently upregulated in various cancers, including triple-negative breast cancer (TNBC), and is associated with poor clinical outcomes. Analysis of The Cancer Genome Atlas (TCGA) data confirmed that PTK7 mRNA expression [...] Read more.
Protein tyrosine kinase 7 (PTK7), a catalytically defective receptor protein tyrosine kinase, is frequently upregulated in various cancers, including triple-negative breast cancer (TNBC), and is associated with poor clinical outcomes. Analysis of The Cancer Genome Atlas (TCGA) data confirmed that PTK7 mRNA expression is significantly higher in TNBC tumor tissues compared with adjacent normal tissues and non-TNBC breast cancer subtypes. Kaplan–Meier survival analysis demonstrated a strong correlation between high PTK7 expression and worse relapse-free survival in TNBC patients (HR = 1.46, p = 0.015). In vitro, anti-PTK7 monoclonal antibodies (mAbs) significantly reduced proliferation, wound healing, migration, and invasion in TNBC MDA-MB-231 cells. Ki-67 immunofluorescence assays revealed substantial decreases in cell proliferation following treatment with PTK7 mAbs (32-m, 43-m, 50-m, and 52-m). Moreover, actin polymerization, a critical process in cell migration and invasion, was markedly impaired upon PTK7 mAb treatment. In vivo, PTK7 mAbs significantly reduced tumor volume and weight in a TNBC xenograft mouse model compared with controls. Treated tumors exhibited decreased expression of Ki-67 and vimentin, indicating reduced proliferation and epithelial-to-mesenchymal transition. These findings highlight PTK7 as a promising therapeutic target in TNBC and demonstrate the potent anti-cancer effects of PTK7-neutralizing mAbs both in vitro and in vivo. Further exploration of PTK7-targeted therapies, including humanized mAbs and antibody-drug conjugates, is warranted to advance treatment strategies for PTK7-positive TNBC. Full article
Show Figures

Figure 1

Figure 1
<p>PTK7 mRNA expression in BC tumor and adjacent normal tissues and in non-TNBC and TNBC tumor tissues. Expression levels of PTK7 mRNA in BC and adjacent normal tissues (<b>A</b>) and in non-TNBC and TNBC tissues (<b>B</b>) were analyzed using TCGA data through the Breast Cancer Integrative Platform (<a href="http://www.omicsnet.org/bcancer/" target="_blank">http://www.omicsnet.org/bcancer/</a> accessed on 6 January 2024). Bars represent the median expression levels in each group. Statistical analysis was performed using the Mann–Whitney U test.</p>
Full article ">Figure 2
<p>Survival analysis based on PTK7 mRNA expression in BC and TNBC cohorts. The prognostic significance of PTK7 mRNA expression in BC and TNBC was evaluated using the Kaplan–Meier Plot database (<a href="http://kmplot.com/analysis/" target="_blank">http://kmplot.com/analysis/</a> accessed on 6 January 2024). OS and RFS were analyzed in all BC patients, while DMFS was assessed in patients with lymph node-positive cancer. The number of patients with low and high PTK7 expression used in each analysis is indicated. Statistical significance was determined using log-rank tests. A hazard ratio (HR) below 1 indicates better survival probabilities for the low-PTK7-expression group compared with the high-PTK7-expression group.</p>
Full article ">Figure 3
<p>Effect of anti-PTK7 mAbs on proliferation in MDA-MB-231 cells. MDA-MB-231 cells were incubated in DMEM without or with 5% FBS and treated with anti-PTK7 mAbs (32-m, 43-m, 50-m, and 52-m; 10 μg/mL) for 24 h. Proliferating cells were detected by staining with anti-Ki-67 and Rhodamine Red-X-conjugated anti-mouse IgG antibodies, with nuclear staining performed with Hoechst 33258. (<b>A</b>) Representative images were captured using confocal fluorescence microscopy at 200× magnification. (<b>B</b>) Quantification of Ki-67 staining was performed using ImageJ software (version 1.53), normalized to nuclear staining. Each value represents the mean ± SD of three independent experiments, with statistical significance denoted as *** <span class="html-italic">p</span> &lt; 0.001 compared with the FBS-treated control group. The scale bar represents 100 μm.</p>
Full article ">Figure 4
<p>Effect of anti-PTK7 mAbs on wound healing, migration, and invasion in MDA-MB-231 cells. (<b>A</b>) Wound healing was evaluated 24 h post-scratch in monolayer culture after treatment with anti-PTK7 mAbs. Representative micrographs (100× magnification) were captured at 0 h and 24 h post-antibody application. The wound-healing area was quantified using ImageJ software. The graph shows the relative wound area in anti-PTK7 mAb-treated groups normalized to the FBS-treated control group. The scale bar represents 1 mm. (<b>B</b>) Chemotactic migration was measured 24 h post-antibody treatment using a Transwell apparatus. (<b>C</b>) Chemotactic invasion was measured 48 h post-antibody treatment using a Transwell apparatus. Migrated and invaded cells on the bottom of the Transwell were stained with crystal violet for visualization. Representative micrographs (40× magnification) were captured at 24 h for migration and 48 h for invasion. Stained cells were solubilized in 1% sodium dodecyl sulfate, and absorbance was measured at 600 nm. Graphs show the relative absorbance of the anti-PTK7 mAbs-treated groups, normalized to the FBS-treated control group. The scale bar represents 500 μm. Each value represents the mean ± SD of three independent experiments: ** <span class="html-italic">p</span> &lt; 0.01 or *** <span class="html-italic">p</span> &lt; 0.001 vs. FBS-treated control.</p>
Full article ">Figure 5
<p>Effect of anti-PTK7 mAbs on actin polymerization in MDA-MB-231 cells. MDA-MB-231 cells were cultured in DMEM supplemented with or without 5% FBS and treated with anti-PTK7 mAbs (10 μg/mL) for 24 h. Actin filaments visualized using FITC-phalloidin are shown alongside nuclei staining (Hoechst 33258). Images were captured via confocal fluorescence microscopy at 200× magnification. The scale bar represents 100 μm.</p>
Full article ">Figure 6
<p>Inhibition of tumorigenesis by anti-PTK7 mAbs in the TNBC xenograft model. The TNBC xenograft mouse model was established by subcutaneous injection of MDA-MB-231 cells into the dorsal regions of nude mice. PTK7 mAbs (10 mg/kg) were administered intravenously via tail veins twice a week for three weeks (indicated by red vertical arrows). Mice were euthanized 2 weeks after the final treatment, and tumors were collected for analysis. (<b>A</b>) Tumor volume was monitored throughout the in vivo experiment. (<b>B</b>) Images of tumors excised from euthanized mice are shown. (<b>C</b>) Tumor weight data are presented graphically. Each value represents the mean ± SD of three mice: * <span class="html-italic">p</span> &lt; 0.05 vs. the vehicle control group.</p>
Full article ">Figure 7
<p>H&amp;E and IHC staining of Ki-67 and vimentin in xenograft tumor sections. (<b>A</b>) Representative images of H&amp;E and IHC staining for Ki-67 and vimentin are shown. Ki-67 (<b>B</b>) and vimentin (<b>C</b>) levels in IHC images were quantitated using ImageJ software. Each value represents the mean ± SD of three mice. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 vs. vehicle control group; n.s.: not significant. The scale bar represents 100 μm.</p>
Full article ">
20 pages, 9784 KiB  
Article
Transparent Poly(amide-imide)s with Low Coefficient of Thermal Expansion from Trifluoromethylated Trimellitic Anhydride
by Seong Jong Kim, SeongUk Jeong, Taejoon Byun, Jun Sung Kim, Haeshin Lee and Sang Youl Kim
Polymers 2025, 17(3), 309; https://doi.org/10.3390/polym17030309 - 24 Jan 2025
Viewed by 138
Abstract
Making transparent aromatic polymers with high Tg and low thermal expansion behavior, like glass, is challenging. We report transparent and soluble poly(amide-imide)s (PAIs) with high dimensional stability synthesized from the new monomer, trifluoromethylated trimellitic anhydride. Insertion of trifluoromethyl (CF3) groups [...] Read more.
Making transparent aromatic polymers with high Tg and low thermal expansion behavior, like glass, is challenging. We report transparent and soluble poly(amide-imide)s (PAIs) with high dimensional stability synthesized from the new monomer, trifluoromethylated trimellitic anhydride. Insertion of trifluoromethyl (CF3) groups into polymer chains enhanced solubility and the optical properties of polymers without sacrificing high thermal stability. Model reactions were utilized to study how the CF3 group in trimellitic anhydride affects the polymerization reaction with aromatic diamine monomers, and a series of new PAIs were synthesized. All the polymers were soluble in polar organic solvents and can be solution-cast into nearly colorless and flexible freestanding films. The obtained PAI films possessed high thermal stability (Td5: 437–452 °C in N2) and high transparency (84~87% transmittance at 550 nm). Interestingly, PAIs prepared in this study exhibited high thermodimensional stability with low CTE values from 9 to 26 ppm/°C. The transparent poly(amide-imide) film with low CTE value finds its application in display and optical devices that require flexible and transparent form factors. Full article
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H NMR spectra of (<b>a</b>) <b>1</b> (DMSO-<span class="html-italic">d</span><sub>6</sub>) and (<b>b</b>) <b>2</b> (Chloroform-<span class="html-italic">d</span>).</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR spectra of (<b>a</b>) <b>3</b> and (<b>b</b>) <b>4</b> (DMSO-<span class="html-italic">d</span><sub>6</sub>), and <sup>13</sup>C NMR spectrum of (<b>c</b>) <b>4</b> (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 3
<p><sup>1</sup>H NMR spectra of (<b>a</b>) <b>5</b> and (<b>b</b>) <b>6</b> (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 4
<p><sup>1</sup>H NMR spectra of model reaction for <b>8</b>, aromatic region (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 5
<p><sup>1</sup>H NMR spectra of (<b>a</b>) model compound <b>8</b> and (<b>b</b>) <b>8</b> with side products (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 6
<p><sup>1</sup>H NMR spectra of model reaction for <b>9</b>, aromatic region (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 7
<p><sup>1</sup>H NMR spectrum of the model compound <b>9</b> (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 8
<p><sup>1</sup>H NMR spectra of model reaction for <b>10</b>, aromatic region (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 9
<p><sup>1</sup>H NMR spectra of (<b>a</b>) model compound <b>10</b> and (<b>b</b>) <b>10</b> with side products (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 10
<p>(<b>a</b>) <sup>1</sup>H NMR spectrum and (<b>b</b>) <sup>13</sup>C NMR spectrum of <b>11</b> (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 11
<p>FT-IR spectra of PAIs (film).</p>
Full article ">Figure 12
<p><sup>1</sup>H NMR spectra of PAIs (DMSO-<span class="html-italic">d</span><sub>6</sub>).</p>
Full article ">Figure 13
<p>TGA curves of the PAIs in (<b>a</b>) N<sub>2</sub> and (<b>b</b>) air at a heating rate of 10 °C/min.</p>
Full article ">Figure 14
<p>(<b>a</b>) DSC curves of the PAIs (the second heating run ranging from 0 °C to 350 °C at a heating rate of 10 °C/min in N<sub>2</sub>). (<b>b</b>) TMA curves of the PAIs (the second (solid) and third (dash) heating runs ranging from 40 °C to 250 °C at a heating rate of 5 °C/min in N<sub>2</sub>).</p>
Full article ">Figure 15
<p>(<b>a</b>) Images of the PAI films. (<b>b</b>) Transmittance UV-Vis spectra of the PAI films.</p>
Full article ">Scheme 1
<p>Synthetic scheme of the trifluoromethylated trimellitic anhydride and the diacid monomer.</p>
Full article ">Scheme 2
<p>Polymerization of the diacid monomer <b>6</b> with <b><span class="html-italic">s</span>DA</b>.</p>
Full article ">Scheme 3
<p>Model reaction of <b>6</b> with 3-aminobenzotrifluoride.</p>
Full article ">Scheme 4
<p>Scheme of the proposed model (<b>a</b>) and side (<b>b</b>) reactions (highlighted in red color).</p>
Full article ">Scheme 5
<p>Model reaction of <b><span class="html-italic">s</span>DAc</b> with 3-aminobenzotrifluoride.</p>
Full article ">Scheme 6
<p>Model reaction of <b>6</b> with <span class="html-italic">p</span>-anisidine.</p>
Full article ">Scheme 7
<p>Proposed mechanism of side reaction in the model reaction for <b>10</b> (highlighted in red color).</p>
Full article ">Scheme 8
<p>Synthesis of the diacid monomer <b>11</b>.</p>
Full article ">Scheme 9
<p>Polymerization of the diacid monomer <b>11</b>.</p>
Full article ">
13 pages, 2932 KiB  
Article
Encapsulation of Fatty Acids Using Linear Dextrin from Waxy Potato Starch: Effect of Debranching Time and Degree of Unsaturation
by Huifang Xie, Qingfei Duan, Guohua Hu, Xinyi Dong, Litao Ma, Jun Fu, Yiwen Yang, Huaran Zhang, Jiahui Song, Qunyu Gao and Long Yu
Gels 2025, 11(2), 91; https://doi.org/10.3390/gels11020091 - 24 Jan 2025
Viewed by 185
Abstract
This study investigates the effects of the debranching time of waxy potato starch using pullulanase and recrystallization on particle morphology, debranching degree, and crystal structure. The results demonstrated that after gelatinization and debranching, the surface of the starch crystals became rough and uneven [...] Read more.
This study investigates the effects of the debranching time of waxy potato starch using pullulanase and recrystallization on particle morphology, debranching degree, and crystal structure. The results demonstrated that after gelatinization and debranching, the surface of the starch crystals became rough and uneven due to hydrolysis, with most particles showing a fragmented surface. The crystalline state was not significantly changed with debranching time. X-ray diffraction analysis revealed no significant differences in the patterns of recrystallized linear dextrin (LD) after various debranching times. Notably, the short-range ordered structure of LD after debranching and recrystallization was more organized than that of the original or gelatinized starch. Additionally, polarized light microscopy showed that the birefringent pattern disappeared as a result of debranching and recrystallization, indicating the breakdown of particle structure, although the overall particle morphology did not change significantly with varying debranching times. Furthermore, linear dextrin derived from starch debranched for 6 h (with pullulanase at 15 μg/g) successfully embedded stearic acid, oleic acid, and linoleic acid, forming a VI-type starch–fatty acid complex. Full article
(This article belongs to the Special Issue Recent Advances in Food Gels (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>SEM micrographs of native WPS ((<b>A1</b>), 200×; (<b>A2</b>), 500×), gelatinized WPS ((<b>B1</b>), 200×; (<b>B2</b>), 500×), DBS-WPS-2 ((<b>C1</b>), 200×; (<b>C2</b>), 500×), DBS-WPS-4 ((<b>D1</b>), 200×; (<b>D2</b>), 500×), DBS-WPS-6 ((<b>E1</b>), 200×; (<b>E2</b>), 500×) and DBS-WPS-8 ((<b>F1</b>), 200×; (<b>F2</b>), 500×).</p>
Full article ">Figure 1 Cont.
<p>SEM micrographs of native WPS ((<b>A1</b>), 200×; (<b>A2</b>), 500×), gelatinized WPS ((<b>B1</b>), 200×; (<b>B2</b>), 500×), DBS-WPS-2 ((<b>C1</b>), 200×; (<b>C2</b>), 500×), DBS-WPS-4 ((<b>D1</b>), 200×; (<b>D2</b>), 500×), DBS-WPS-6 ((<b>E1</b>), 200×; (<b>E2</b>), 500×) and DBS-WPS-8 ((<b>F1</b>), 200×; (<b>F2</b>), 500×).</p>
Full article ">Figure 2
<p>XRD patterns of waxy potato starch debranched by pullulans with different times.</p>
Full article ">Figure 3
<p>(<b>a</b>) FT-IR patterns of waxy potato starch debranched by pullulans with different times, (<b>b</b>) enlarged by 1200~800 cm<sup>−1</sup>.</p>
Full article ">Figure 4
<p><sup>1</sup>H NMR patterns of waxy potato starch debranched by pullulans with different times.</p>
Full article ">Figure 5
<p>XRD pattern of the complexes of waxy potato starch debranched by 15 μ/g pullulanase for 2 h, 4 h, 6 h and 8 h with stearic acid at a ratio 10:1.</p>
Full article ">Figure 6
<p>XRD patterns of waxy potato starch debranched for 6 h and (<b>A</b>) stearic acid (SA), (<b>B</b>) oleic acid (OA) and (<b>C</b>) linoleic acid (LA) complexes.</p>
Full article ">
17 pages, 3279 KiB  
Article
Fabrication of Functional Polymers with Gradual Release of a Bioactive Precursor for Agricultural Applications
by Oscar G. Marambio, Rudy Martin-Trasancos, Julio Sánchez, Felipe A. Ramos and Guadalupe del C. Pizarro
Gels 2025, 11(2), 90; https://doi.org/10.3390/gels11020090 - 24 Jan 2025
Viewed by 182
Abstract
Biodegradable and biocompatible polymeric materials and stimulus-responsive hydrogels are widely used in the pharmaceutical, agricultural, biomedical, and consumer sectors. The effectiveness of these formulations depends significantly on the appropriate selection of polymer support. Through chemical or enzymatic hydrolysis, these materials can gradually release [...] Read more.
Biodegradable and biocompatible polymeric materials and stimulus-responsive hydrogels are widely used in the pharmaceutical, agricultural, biomedical, and consumer sectors. The effectiveness of these formulations depends significantly on the appropriate selection of polymer support. Through chemical or enzymatic hydrolysis, these materials can gradually release bioactive agents, enabling controlled drug release. The objective of this work is to synthesize, characterize, and apply two controlled-release polymeric systems, focusing on the release of a phyto-pharmaceutical agent (herbicide) at varying pH levels. The copolymers were synthesized via free radical polymerization in solution, utilizing tetrahydrofuran (THF) as the organic solvent and benzoyl peroxide (BPO) as the initiator, without the use of a cross-linking agent. Initially, the herbicide was grafted onto the polymeric chains, and its release was subsequently tested across different pH environments in a heterogeneous phase using an ultrafiltration (UF) system. The development of these two controlled-release polymer systems aimed to measure the herbicide’s release across different pH levels. The goal is to adapt these materials for agricultural use, enhancing soil quality and promoting efficient water usage in farming practices. The results indicate that the release of the herbicide from the conjugate systems exceeded 90% of the bioactive compound after 8 days at pH 10 for both systems. Furthermore, the two polymeric systems demonstrated first-order kinetics for herbicide release in aqueous solutions at different pH levels. The kinetic constant was found to be higher at pH 7 and 10 compared to pH 3. These synthetic hydrogels are recognized as functional polymers suitable for the sustained release of herbicides in agricultural applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Polymerization reaction of the P(HEMA-<span class="html-italic">alt</span>-MAn) and P(HPMA-<span class="html-italic">alt</span>-MAn) and copolymers under hydrolysis conditions.</p>
Full article ">Figure 2
<p>(<b>a</b>) FTIR spectrum of P(HEMA-<span class="html-italic">alt</span>-MAn) and (<b>b</b>) its copolymer conjugate 2,4-D.</p>
Full article ">Figure 3
<p>(<b>a</b>) FTIR spectrum of P(HPMA-<span class="html-italic">alt</span>-MAn) and (<b>b</b>) its copolymer conjugate 2,4-D.</p>
Full article ">Figure 4
<p>(<b>a</b>) Grafting reaction of 2,4-D onto P(HEMA-<span class="html-italic">alt</span>-MAn) and P(HPMA-<span class="html-italic">alt</span>-MAn) and (<b>b</b>) the corresponding hydrolysis of the anhydride upon swelling.</p>
Full article ">Figure 5
<p><sup>1</sup>H-NMR spectrum of P(HPMA-<span class="html-italic">alt</span>-MAn) and its copolymer conjugate 2,4-D.</p>
Full article ">Figure 6
<p>Thermograms of P(HEMA-<span class="html-italic">alt</span>-MAn) and its conjugate P(HEMA-<span class="html-italic">alt</span>-MAn)-2,4-D (<b>left</b>), and P(HPMA-<span class="html-italic">alt</span>-MAn) and its conjugate P(HPMA-<span class="html-italic">alt</span>-MAn)-2,4-D (<b>right</b>). Heating rate, 10 °C min<sup>−1</sup>, in inert environment (gaseous N<sub>2</sub>).</p>
Full article ">Figure 7
<p>Swelling isotherm of P(HEMA-<span class="html-italic">alt</span>-MAn) and P(HPMA-<span class="html-italic">alt</span>-MAn) as a function of time in buffered solutions at specific pH levels: 3 (●), 7 (■), and 10 (▲) at 25 °C.</p>
Full article ">Figure 8
<p>Release of 2,4-D (%) from copolymer–herbicide conjugates in an aqueous solution at pH 3, 7, and 10.</p>
Full article ">Figure 9
<p>Schematic representation of inter- or intra-hydrogen bonds in fabricating stimulus-responsive hydrogels.</p>
Full article ">Figure 10
<p>Determination of kinetic release constants at different pH. Plot of ln <span class="html-italic">c<sup>bound</sup></span> vs. time (days).</p>
Full article ">Figure 11
<p>Experimental set-up of ultrafiltration system: (a) filtration cell, (b) filtrate, (c) magnetic stirrer, (d) selector, (e) reservoir, and (f) pressure source.</p>
Full article ">
17 pages, 2080 KiB  
Article
Multi-Responsive Amphiphilic Hyperbranched Poly[(2-dimethyl aminoethyl methacrylate)-co-(benzyl methacrylate)]copolymers: Self-Assembly and Curcumin Encapsulation in Aqueous Media
by Foteini Ginosati, Dimitrios Vagenas, Angelica Maria Gerardos and Stergios Pispas
Materials 2025, 18(3), 513; https://doi.org/10.3390/ma18030513 - 23 Jan 2025
Viewed by 203
Abstract
In this study, we report the synthesis of amphiphilic hyperbranched poly[(2-dimethylaminoethyl methacrylate)-co-(benzyl methacrylate)] statistical copolymers with two different stoichiometric compositions using the reversible addition–fragmentation chain transfer polymerization (RAFT) technique. The selection of monomers was made to incorporate a pH and thermoresponsive polyelectrolyte (DMAEMA) [...] Read more.
In this study, we report the synthesis of amphiphilic hyperbranched poly[(2-dimethylaminoethyl methacrylate)-co-(benzyl methacrylate)] statistical copolymers with two different stoichiometric compositions using the reversible addition–fragmentation chain transfer polymerization (RAFT) technique. The selection of monomers was made to incorporate a pH and thermoresponsive polyelectrolyte (DMAEMA) component and a hydrophobic component (BzMA) to achieve amphiphilicity and study the effects of architecture and environmental factors on the behavior of the novel branched copolymers. Molecular characterization was performed through size exclusion chromatography (SEC) and spectroscopic characterization techniques (1H-NMR and FT-IR). The self-assembly behavior of the hyperbranched copolymers in aqueous media, in response to variations in pH, temperature, and ionic strength, was studied using dynamic light scattering (DLS), electrophoretic light scattering (ELS), and fluorescence spectroscopy (FS). Finally, the efficacy of the two novel copolymers to encapsulate curcumin (CUR), a hydrophobic, polyphenolic drug with proven anti-inflammatory and fluorescence properties, was established. Its encapsulation was evaluated through DLS, UV–Vis, and fluorescence measurements, investigating the change of hydrodynamic radius of the produced mixed copolymer–CUR nanoparticles in each case and their fluorescence emission properties. Full article
(This article belongs to the Special Issue Applied Stimuli-Responsive Polymer Based Materials)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H-NMR spectra of HB1 (green) and HB2 (blue) hyperbranched copolymers (the solvent is deuterated acetone (CD<sub>3</sub>)<sub>2</sub>CO, and the letters indicate the corresponding hydrogens in the chemical structure).</p>
Full article ">Figure 2
<p>SEC chromatogram for HB1 and HB2 hyperbranched copolymers.</p>
Full article ">Figure 3
<p>Critical aggregation concentration (CAC) determination for HB1 (<b>a</b>) and HB2 (<b>b</b>) at pH 7.</p>
Full article ">Figure 4
<p>Size distributions from DLS for the HB1 and HB2 hyperbranched copolymer solutions at (<b>a</b>) pH 3, (<b>b</b>) pH 7, and (<b>c</b>) pH 10.</p>
Full article ">Figure 5
<p>Size distributions from DLS for the HB1 (<b>a</b>) and HB2 (<b>b</b>) hyperbranched copolymer solutions as a function of temperature.</p>
Full article ">Figure 6
<p>Ionic strength dependence plots for (<b>a</b>) HB1 and (<b>b</b>) HB2 copolymers (polymer concentration 10<sup>−3</sup> g/mL, at pH = 7 and at a temperature of 25 °C).</p>
Full article ">Figure 7
<p>Comparative size distributions of HB1 and HB2 copolymers with curcumin encapsulationat 10%<span class="html-italic">w</span>/<span class="html-italic">w</span> (<b>a</b>,<b>c</b>) and 20% <span class="html-italic">w</span>/<span class="html-italic">w</span> (<b>b</b>,<b>d</b>) in different media.</p>
Full article ">Figure 8
<p>Fluorescence of curcumin in acetone (<b>a</b>) and of HB1 aggregates at 10%<span class="html-italic">w</span>/<span class="html-italic">w</span> and 20%<span class="html-italic">w</span>/<span class="html-italic">w</span> curcumin encapsulation (<b>b</b>) and HB2 ones at 10%<span class="html-italic">w</span>/<span class="html-italic">w</span> and 20%<span class="html-italic">w/w</span> curcumin encapsulation (<b>c</b>).</p>
Full article ">Scheme 1
<p>Reaction scheme of the synthesis of P(DMAEMA-co-BzMA) hyperbranched copolymers.</p>
Full article ">
14 pages, 4050 KiB  
Article
Stability and Controlled Polymerization of Trithiocarbonate Chain Transfer Agents Under Harsh Conditions
by Thi Ngan Vu, Tomoya Nishimura, Yu Osaki, Toyohiro Otani and Shin-ichi Yusa
Polymers 2025, 17(3), 297; https://doi.org/10.3390/polym17030297 - 23 Jan 2025
Viewed by 216
Abstract
This study investigates the stability and application of trithiocarbonate-based chain transfer agents (CTAs) in reversible addition–fragmentation chain transfer (RAFT) radical polymerization under harsh conditions. We evaluated the stability of 4-cyano-4-(2-carboxyethylthiothioxomethylthio) pentanoic acid (Rtt-17) and 4-cyano-4-(dodecylsulfanylthiocarbonyl) sulfanylpentanoic acid (Rtt-05) at 60 °C under basic [...] Read more.
This study investigates the stability and application of trithiocarbonate-based chain transfer agents (CTAs) in reversible addition–fragmentation chain transfer (RAFT) radical polymerization under harsh conditions. We evaluated the stability of 4-cyano-4-(2-carboxyethylthiothioxomethylthio) pentanoic acid (Rtt-17) and 4-cyano-4-(dodecylsulfanylthiocarbonyl) sulfanylpentanoic acid (Rtt-05) at 60 °C under basic conditions using 1H NMR and UV–vis absorption spectra, showing that Rtt-05 is more stable than Rtt-17. The greater stability of Rtt-05 is attributed to the hydrophobic dodecyl group, which allows it to form micelles in water, thereby protecting the trithiocarbonate group from the surrounding aqueous phase. In contrast, hydrophilic Rtt-17, without long alkyl chains, cannot form micelles in water. Following the stability assessment, Rtt-17 and Rtt-05 were employed for RAFT polymerization of hydrophilic monomers, such as N,N-dimethylacrylamide (DMA) and 2-(methacryloyloxy)ethyl phosphorylcholine (MPC). DMA can dissolve in both water and organic solvents, and MPC can dissolve in water and polar solvents. Both CTAs successfully controlled the polymerization of DMA, producing polymers with narrow molecular weight distributions (Mw/Mn) less than 1.2. Also, Rtt-17 demonstrated effective control of MPC polymerization, yielding Mw/Mn values of around 1.2. However, during the polymerization of MPC, Rtt-05 failed to maintain control, resulting in a broad Mw/Mn (≥1.9). The inability of Rtt-05 to control MPC polymerization is due to the formation of micelles, which disrupts the interaction between the hydrophilic MPC propagating radicals and the trithiocarbonate group in the hydrophobic core of Rtt-05 micelles. The findings provide critical insights into designing CTAs for specific applications, particularly for biomedical and industrial uses of hydrophilic polymers, highlighting the potential for precise molecular weight control and tailored polymer properties. Full article
(This article belongs to the Collection Polymerization and Kinetic Studies)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Chemical structures of MPC and DMA and (<b>b</b>) conceptual illustrations of polymerization of the monomers under harsh conditions (pH 10 and 70 °C) using Rtt-17 and Rtt-05 as chain transfer agents.</p>
Full article ">Figure 2
<p>UV–vis absorption spectra of (<b>a</b>) Rtt-17 and (<b>b</b>) Rtt-05 with 0.025 g/L at 25 °C under various pH conditions after 24 h at 60 °C: The pH values were indicated in the spectra.</p>
Full article ">Figure 3
<p>The decomposition rates estimated from (<b>a</b>) <sup>1</sup>H NMR integral intensity ratios of signals at 1.74 ppm and (<b>b</b>) UV–vis absorption spectra for Rtt-17 (<span style="color:red">●</span>) and Rtt-05 (<span style="color:blue">■</span>) under various pH conditions after 24 h at 60 °C.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV–vis absorption spectra of Nile red in water in the presence of Rtt-05 at various concentrations at pH 11: The concentration of Rtt-05 ([Rtt-05]) was indicated in the spectra and (<b>b</b>) absorbance at 572 nm for Nile red as a function of CTA concentrations at pH 11 for Rtt-17 (<b><span style="color:blue">□</span></b>) and Rtt-05 (<b><span style="color:red">○</span></b>) at 25 °C.</p>
Full article ">Figure 5
<p>Hydrodynamic radius (<span class="html-italic">R</span><sub>h</sub>) distributions for Rtt-05 at 10 g/L in water under pH 10 at (<b>a</b>) 25 °C and (<b>b</b>) 70 °C.</p>
Full article ">Figure 6
<p>Time–conversion (<span class="html-italic">p</span>) (<b><span style="color:red">□</span></b>) and first-order kinetic plots (<b><span style="color:blue">○</span></b>) of PDMA using (<b>a</b>) Rtt-17 and (<b>b</b>) Rtt-05; [M]<sub>0</sub> and [M] are the monomer concentrations at the polymerization time = 0 min and the corresponding time, respectively: The arrows indicated the axis.</p>
Full article ">Figure 7
<p>The early-stage, pre-equilibrium reaction between V-501 and CTAs.</p>
Full article ">Figure 8
<p>GPC elution curves of PDMA prepared using (<b>a</b>) Rtt-17 and (<b>b</b>) Rtt-05 and conversion (<span class="html-italic">p</span>)–<span class="html-italic">M</span><sub>n</sub> (<b><span style="color:red">□</span></b>) with theoretical line (<b><span style="color:#538135">—</span></b>) and <span class="html-italic">p</span>-<span class="html-italic">M</span><sub>w</sub>/<span class="html-italic">M</span><sub>n</sub> plots (<b><span style="color:blue">○</span></b>) using (<b>c</b>) Rtt-17 and (<b>d</b>) Rtt-05: The arrows indicated the axis.</p>
Full article ">Figure 9
<p>Time–conversion (<span class="html-italic">p</span>) (<b><span style="color:red">□</span></b>) and first-order kinetic plots (<b><span style="color:blue">○</span></b>) of PMPC using (<b>a</b>) Rtt-17 and (<b>b</b>) Rtt-05; [M]<sub>0</sub> and [M] are the monomer concentrations at the polymerization time = 0 min and the corresponding time, respectively: The arrows indicated the axis.</p>
Full article ">Figure 10
<p>GPC elution curves of PMPC prepared using (<b>a</b>) Rtt-17 and (<b>b</b>) Rtt-05 and conversion (<span class="html-italic">p</span>)–<span class="html-italic">M</span><sub>n</sub> (<b><span style="color:red">□</span></b>) with theoretical line (<b><span style="color:#538135">—</span></b>) and <span class="html-italic">p</span>-<span class="html-italic">M</span><sub>w</sub>/<span class="html-italic">M</span><sub>n</sub> plots (<b><span style="color:blue">○</span></b>) using (<b>c</b>) Rtt-17 and (<b>d</b>) Rtt-05: The arrows indicated the axis.</p>
Full article ">
19 pages, 2411 KiB  
Article
Modification of Structure, Pasting, and In Vitro Digestion Properties of Glutinous Rice Starch by Different Lactic Acid Bacteria Fermentation
by Dongliang Shao, Jigang Zhang, Tiantian Shao, Yuhui Li, Hongkui He, Yanli Wang, Jintong Ma, Runjie Cao, Anjun Li and Xianfeng Du
Foods 2025, 14(3), 367; https://doi.org/10.3390/foods14030367 - 23 Jan 2025
Viewed by 320
Abstract
This research evaluated the effect of fermentation with Lactobacillus plantarum 11122, Lactobacillus casei 23184, and Lactobacillus lactis 1011 on structure, pasting, and in vitro digestion properties of glutinous rice starch varying in TN and HY genotype, respectively. The results showed that fermentation decreased [...] Read more.
This research evaluated the effect of fermentation with Lactobacillus plantarum 11122, Lactobacillus casei 23184, and Lactobacillus lactis 1011 on structure, pasting, and in vitro digestion properties of glutinous rice starch varying in TN and HY genotype, respectively. The results showed that fermentation decreased the weight-average molecular weight and increased the radius of gyration. The short chain was increased by degrading the medium chain (B2, DP 24−35) of amorphous in starch, which directly led to the increase of branching degree and rearrangement of the starch chain. LAB fermentation increases the short-range ordered structure, helix structure, and crystallinity by polymerization or interactions of short chains between intermolecular and intramolecular. Furthermore, the pasting characteristic of the fermented starch sample obtained obvious improvement in terms of hydration capacity, including breakdown and setback value. Fermentation facilitated the forming of both slowly digestible starch (17.1–30.79%) and resistant starch (32.3–46.8%) in TN but caused a decline in the content of rapidly digestible starch (25.47–43.6% in TN, 9.36–17.8% in HY). The result of Pearson’s correlation tests and PCA showed the variety of structural and physicochemical of fermentation-treated starch depend highly on the starter culture and starch resources. These results provided new data support for the potential application of modified starch by fermentation with LABs. Full article
Show Figures

Figure 1

Figure 1
<p>Changes of (<b>A</b>) pH value, (<b>B</b>) TTA, and (<b>C</b>) viable cell number in fermented glutinous rice mixtures.</p>
Full article ">Figure 2
<p>Profiles of <sup>13</sup>C CP/MAS NMR spectra (<b>A</b>,<b>B</b>), XRD patterns (<b>C</b>,<b>D</b>), and <sup>1</sup>H NMR spectra (<b>E</b>,<b>F</b>) of glutinous rice starch with different LAB fermentation.</p>
Full article ">Figure 3
<p>FTIR spectra curve of native and LAB fermented glutinous rice starch.</p>
Full article ">Figure 4
<p>RVA graphs of single LAB strain fermentation on the starch pasting properties of (<b>A</b>) TN and (<b>B</b>) HY glutinous rice.</p>
Full article ">Figure 5
<p>Digestive properties of native and modified glutinous rice starch.</p>
Full article ">Figure 6
<p>Pearson correlation analysis of structure–property relationships and Principal components analysis of physicochemical properties. Mw: weight average molecular weight, Rz: radius of gyration, PDI: Mw/Mn, DP: degree of polymerization, CR: relative crystallinity, DH: double helices, SH: single helix, AM: amorphous components, DB: Branching Degree, DO: short-range order, DD: degree of double helix, PV: peak viscosity, TV: trough viscosity, BD: breakdown, FV: final viscosity, SB: setback, GT: Gelatinization temperature, RDS: rapidly digestible starch, SDS: Slowly digestible starch, and RS: resistant starch.</p>
Full article ">
32 pages, 19962 KiB  
Review
Noncovalent Interactions in Coordination Chemistry of Cyclic Trinuclear Copper(I) and Silver(I) Pyrazolates
by Arina Olbrykh, Gleb Yakovlev, Aleksei Titov and Elena Shubina
Crystals 2025, 15(2), 115; https://doi.org/10.3390/cryst15020115 - 23 Jan 2025
Viewed by 273
Abstract
Group 11 metals form with pyrazolate ligand complexes with a general formula of [MPz]n. The value of “n” varies depending on the type of substituent in the ligand and the metal atom. Copper(I) and silver(I) ions mainly form cyclic di-, tri-, [...] Read more.
Group 11 metals form with pyrazolate ligand complexes with a general formula of [MPz]n. The value of “n” varies depending on the type of substituent in the ligand and the metal atom. Copper(I) and silver(I) ions mainly form cyclic di-, tri-, and tetra-nuclear complexes or polymeric structures. Cyclic trinuclear d10 metal pyrazolates [MPzm]3 (M = Cu(I) and Ag(I); Pz = substituted pyrazolate ligand) are of particular interest because their planar structure allows them to form supramolecular aggregates via noncovalent metal–metal, metal–π, and metal–electron donor interactions. Designing complexes based on these interactions has been a focus of research for the last two decades. The ability of cyclic trinuclear copper(I) and silver(I) pyrazolates to form coordination and supramolecular structures determines their properties and potential applications in catalysis, gas sensing, molecular recognition, and photoluminescence. In this review, we discuss noncovalent interactions between cyclic trinuclear silver(I) and copper(I) complexes with various types of ligands. Full article
(This article belongs to the Special Issue Reviews of Crystal Engineering)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The representation of dimer formation for [CuPz<sup>1</sup>]<sub>3</sub> (<b>A</b>) and [CuPz<sup>2</sup>]<sub>3</sub> (<b>B</b>). Data from Ref. [<a href="#B55-crystals-15-00115" class="html-bibr">55</a>].</p>
Full article ">Figure 2
<p>(<b>A</b>) Fragment of the infinite stack of [AgPz]<sub>3</sub>. (<b>B</b>) Top view of the infinite [AgPz] stack demonstrating the regularity of the fragment’s position. Data from Refs. [<a href="#B63-crystals-15-00115" class="html-bibr">63</a>,<a href="#B64-crystals-15-00115" class="html-bibr">64</a>].</p>
Full article ">Figure 3
<p>Fragments of the infinite stacks of [CuPz<sup>3</sup>]<sub>3</sub> (<b>A</b>) and [AgPz<sup>3</sup>]<sub>3</sub> (<b>B</b>).</p>
Full article ">Figure 4
<p>Frontier orbitals for the optimized T<sub>1</sub> (<b>A</b>) and S<sub>0</sub> (<b>B</b>) states of dimers of [CuPz<sup>3</sup>]<sub>3</sub>. Reprinted with permission from Ref. [<a href="#B67-crystals-15-00115" class="html-bibr">67</a>]. Copyright 2006 American Chemical Society.</p>
Full article ">Figure 5
<p>Fragments of supramolecular chains of [AgPz<sup>4</sup>]<sub>3</sub> (<b>A</b>) and [AgPz<sup>5</sup>]<sub>3</sub> (<b>B</b>) formed by Ag<sub>3</sub>…Hal contacts.</p>
Full article ">Figure 6
<p>Fragments of supramolecular chains of [CuPz<sup>6</sup>]<sub>3</sub> (<b>A</b>) and [AgPz<sup>6</sup>]<sub>3</sub> (<b>B</b>) formed by M…Hal contacts.</p>
Full article ">Figure 7
<p>Emission and excitation spectra of [CuPz<sup>6</sup>]<sub>3</sub> (<b>A</b>) and [AgPz<sup>6</sup>]<sub>3</sub> (<b>B</b>) at 298 and 77 K. Reprinted with permission from Ref. [<a href="#B70-crystals-15-00115" class="html-bibr">70</a>]. Copyright 2013 American Chemical Society.</p>
Full article ">Figure 8
<p>Schematic representation of the synthesis of [CuPz<sup>9</sup>]<sub>3</sub>(CuX)<sub>m</sub> complexes(<b>a</b>) and structural illustration (<b>b</b>). Reprinted with permission from Ref. [<a href="#B74-crystals-15-00115" class="html-bibr">74</a>] Copyright 2019 American Chemical Society.</p>
Full article ">Figure 9
<p>Dimeric fragments of [CuPz<sup>8</sup>]<sub>3</sub> supported by intermolecular Cu….N<sup>Py</sup> interactions.</p>
Full article ">Figure 10
<p>Fragment of the [CuPz<sup>11</sup>]<sub>3</sub> in a crystal, demonstrating the formation of dimers due to Cu…Cu contacts and the supramolecular organization of these dimers through Cu…O intermolecular interactions.</p>
Full article ">Figure 11
<p>A general representation of the different supramolecular packing of [AgPz<sup>3</sup>]<sub>3</sub> with aromatic compounds. Curly brackets in (<b>a</b>–<b>e</b>) highlight the BAB, BAAB, BAB, BA and BAA fragment in {BAB}<sub>∞</sub>, {BAAB}<sub>∞</sub>, {BAB}<sub>∞</sub>, {BA}<sub>∞</sub> and {BAA}<sub>∞</sub>, respectively.</p>
Full article ">Figure 12
<p>Fragments of supramolecular chains of [CuPz<sup>3</sup>]<sub>3</sub>∙(naphtalene) (<b>A</b>) and [AgPz<sup>3</sup>]<sub>3</sub>∙(naphtalene) (<b>B</b>).</p>
Full article ">Figure 13
<p>General procedures for synthesis of {[MPz<sup>n</sup>]<sub>3</sub>∙(coronene)} adducts. Adapted from Ref. [<a href="#B88-crystals-15-00115" class="html-bibr">88</a>]. Copyright 2023 The Royal Society of Chemistry.</p>
Full article ">Figure 14
<p>Fragment of supramolecular chain of {[AgPz<sup>3</sup>]<sub>3</sub> (<span class="html-italic">o</span>-terphenyl)}.</p>
Full article ">Figure 15
<p>Normalized emission spectra of [AgPz<sup>3</sup>]<sub>3</sub>∙(<span class="html-italic">o</span>-terphenyl) upon excitation from 250 to 370 nm in the solid state at 77 K. Reprinted with permission from Ref. [<a href="#B89-crystals-15-00115" class="html-bibr">89</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 16
<p>Fragment of the supramolecular chain of complex {[AgPz<sup>3</sup>]<sub>3</sub>∙(4,4′–difluoro–1,1′–biphenyl)}.</p>
Full article ">Figure 17
<p>Fragment of the {[AgPz<sup>3</sup>]<sub>3</sub><b>∙</b>(4,4′–diiodo–1,1′–biphenyl)} 1D infinite chain.</p>
Full article ">Figure 18
<p>(<b>a</b>) The photoluminescence spectra of complex {[AgPz<sup>3</sup>]<sub>3</sub>∙(4,4–dibromo–1,1′–biphenyl)} under different excitations. (<b>b</b>) Photographs of complex {[AgPz<sup>3</sup>]<sub>3</sub>∙(4,4′–dibromo–1,1′–biphenyl)} crystals under 365 and 254 nm UV irradiation. (<b>c</b>) Photographs of the abbreviation “INEOS” on a polyurethane-based matrix under UV irradiation at 365 and 254 nm. (<b>d</b>) The CIE 1931 coordinates for {[AgPz<sup>3</sup>]<sub>3</sub>∙(4–chloro–4′–fluoro–1,1′–biphenyl)}.</p>
Full article ">Figure 19
<p>Fragments of the supramolecular chain of the [Cu<sub>2</sub>Pz<sup>13</sup>]<sub>3</sub> cage with benzene (<b>A</b>), pyridine (<b>B</b>), and nitrobenzene (<b>C</b>), showing the intermolecular Cu…Cu and Cu…C intermolecular interactions.</p>
Full article ">Figure 20
<p>The photographs of crystals of cage complexes under 365 nm irradiation and their emission spectra. Adapted from Ref. [<a href="#B92-crystals-15-00115" class="html-bibr">92</a>] with permission from The Royal Society of Chemistry.</p>
Full article ">Figure 21
<p>Fragments of supramolecular chains of {[AgPz<sup>6</sup>]<sub>3</sub>∙(A-PyC)} (<b>A</b>) and {[AgPz<sup>6</sup>]<sub>3</sub>∙(P-PyC)} (<b>B</b>) formed by M…π contacts.</p>
Full article ">Figure 22
<p>Fragments of supramolecular chains of {[AgPz<sup>3</sup>]<sub>3</sub>∙(dmt)} (<b>A</b>) and {[AgPz<sup>6</sup>]<sub>3</sub>∙(dmdbt)} (<b>B</b>) formed by M…π contacts. The Ag…S distances are presented.</p>
Full article ">Figure 23
<p>The crystal structure of {[CuPz<sup>15</sup>]<sub>3</sub>∙(DMT)}.</p>
Full article ">Figure 24
<p>Fragment of the {([CuPz<sup>3</sup>]<sub>3</sub>)<sub>4</sub>∙(C60)}<sub>∞</sub> structure, demonstrating the central core and the tetrahedral arrangement of [CuPz<sup>3</sup>]<sub>3</sub> around fullerene molecule.</p>
Full article ">Figure 25
<p>Crystal structure of {[AgPz<sup>12</sup>]<sub>3</sub>∙2(CH<sub>3</sub>CN)}.</p>
Full article ">Figure 26
<p>Crystal packing fragment of a co-crystal of [AgPz<sup>4</sup>] with CH<sub>3</sub>CN.</p>
Full article ">Figure 27
<p>Fragments of a supramolecular chain of [AgPz<sup>4</sup>]<sub>3</sub>∙crystal, demonstrating coordination with CH<sub>2</sub>Cl<sub>2</sub> (<b>A</b>) and PhCN (<b>B</b>).</p>
Full article ">Figure 28
<p>Fragment of a supramolecular chain of [CuPz<sup>4</sup>]<sub>3</sub>∙crystal, demonstrating coordination with CH<sub>3</sub>CN.</p>
Full article ">Figure 29
<p>Fragments of the {[AgPz<sup>3</sup>]<sub>3</sub>∙(NO<sub>3</sub>)<sub>2</sub>} (<b>A</b>) and {[AgPz<sup>3</sup>]<sub>3</sub> (ba)} packing (<b>B</b>).</p>
Full article ">Figure 30
<p>Crystal structure of {([CuPz<sup>3</sup>]<sub>3</sub>)<sub>2</sub>∙(Ph<sub>2</sub>CO)}.</p>
Full article ">Figure 31
<p>Crystal structures of {[AgPz<sup>3</sup>]<sub>3</sub>∙(Ph<sub>2</sub>CO)} (<b>A</b>) and the dimer of {[AgPz<sup>3</sup>]<sub>3</sub>∙(Ph<sub>2</sub>CO)} (<b>B</b>).</p>
Full article ">Figure 32
<p>Crystal structure of {[AgPz<sup>3</sup>]<sub>3</sub>∙(Ph<sub>2</sub>CO)} (<b>A</b>). Fragment of supramolecular packing of {[AgPz<sup>3</sup>]<sub>3</sub>∙(Ph<sub>2</sub>CO)}<sub>∞</sub> (<b>B</b>).</p>
Full article ">Figure 33
<p>Crystal structures of {[CuPz<sup>3</sup>]<sub>3</sub>∙(AcFc)<sub>2</sub>} (<b>A</b>) and {[AgPz<sup>3</sup>]<sub>3</sub>∙(AcFc)<sub>2</sub>} (<b>B</b>).</p>
Full article ">Figure 34
<p>Fragment of the supramolecular chain of {[AgPz<sup>3</sup>]<sub>3</sub>∙(PhAcFc)}.</p>
Full article ">Figure 35
<p>Workflow for the co-crystallization of <span class="html-italic">C. operculatus</span> crude extract with [AgPz<sup>3</sup>]<sub>3</sub>. Reprinted with permission from Ref. [<a href="#B107-crystals-15-00115" class="html-bibr">107</a>]. Copyright 2024 Elsevier.</p>
Full article ">Figure 36
<p>Crystal structure of dimer {[AgPz<sup>3</sup>]<sub>3</sub>·(BH<sub>3</sub>NEt<sub>3</sub>)}<sub>2</sub>. The proton and non-coordinated hydride atoms are omitted for clarity.</p>
Full article ">Scheme 1
<p>Representation of the main types of d<sup>10</sup> metal pyrazolate complexes.</p>
Full article ">Scheme 2
<p>The chemical structures of pyrazoles, which are used as starting materials for the synthesis of cyclic metal pyrazolate complexes.</p>
Full article ">Scheme 3
<p>The representation of trinuclear complexes from Ref. [<a href="#B70-crystals-15-00115" class="html-bibr">70</a>].</p>
Full article ">Scheme 4
<p>Schematic representation of the <span class="html-italic">syn-</span> and <span class="html-italic">anti</span>-conformers of trimeric copper pyrazolate.</p>
Full article ">Scheme 5
<p>Chemical structures of pyridine-containing chalcones.</p>
Full article ">
18 pages, 2465 KiB  
Article
An In-Vitro Evaluation of Strength, Hardness, and Color Stability of Heat-Polymerized and 3D-Printed Denture Base Polymers After Aging
by Abdulrahman Al-Ameri, Othman Y. Alothman, Omar Alsadon and Durgesh Bangalore
Polymers 2025, 17(3), 288; https://doi.org/10.3390/polym17030288 - 23 Jan 2025
Viewed by 283
Abstract
This study evaluated the strength, hardness, and color stability of 3D-printed denture base resins and compared the outcome with conventional heat-cured denture base resins after aging by thermocycling. A total of 72 specimens from conventional and 3D-printed materials were fabricated in different shapes [...] Read more.
This study evaluated the strength, hardness, and color stability of 3D-printed denture base resins and compared the outcome with conventional heat-cured denture base resins after aging by thermocycling. A total of 72 specimens from conventional and 3D-printed materials were fabricated in different shapes and dimensions based on the mechanical and color tests performed. The specimens were divided into five groups: flexural, tensile, and compressive strengths (n = 20), hardness, and color stability (n = 6). In all these groups, half of the specimens were stored in a distilled water bath at 37 °C for 24 h, and the remaining half of the specimens were subjected to aging by thermocycling. The 3D-printed specimens demonstrated the highest means of tensile strength (32.20 ± 3.8 MPa), compressive strength (106.31 ± 4.07 MPa), and Vickers hardness number (24.51 ± 0.36), and the lowest means of flexural strength (54.29 ± 13.17 MPa) and color difference (ΔE = 2.18 ± 1.09). Conventional heat-cured specimens demonstrated the highest means of flexural strength (59.96 ± 8.39 MPa) and color difference (ΔE = 4.74 ± 2.37) and the lowest means of tensile strength (32.17 ± 9.06 MPa), compressive strength (46.05 ± 4.98 MPa), and Vickers hardness number (10.42 ± 1.05). Aging significantly reduced the flexural strength (−27%), tensile strength (−44%), and hardness (−7%) of 3D-printed resins in contrast to the conventional resin’s compressive strength (−15%) and color stability (p < 0.05). The 3D-printed resin had comparable flexural and tensile strength and significantly superior compressive strength, hardness, and color stability compared with conventional resins. Aging significantly and negatively affected the flexural strength, tensile strength, and hardness of 3D-printed resin. Full article
(This article belongs to the Special Issue 3D Printing and Molding Study in Polymeric Materials)
Show Figures

Figure 1

Figure 1
<p>Specimen shapes and dimensions used in this study. (<b>a</b>) flexural strength, (<b>b</b>) Tensile strength, (<b>c</b>) compressive strength and (<b>d</b>) hardness and color stability.</p>
Full article ">Figure 2
<p>Schematic illustration of strength test set-up. (<b>a</b>) Flexural, (<b>b</b>) Tensile, (<b>c</b>) Compressive.</p>
Full article ">Figure 3
<p>Mean flexural strength of conventional and 3D-printed resins at baseline and after thermocycling (Baseline, BL; Thermocycled, AT). Materials with the same lowercase letters are not significantly different.</p>
Full article ">Figure 4
<p>Mean elastic modulus of conventional and 3D-printed resins at baseline and after thermocycling (Baseline, BL; Thermocycled, AT). Materials with the same lowercase letters are not significantly different.</p>
Full article ">Figure 5
<p>Mean tensile strength of conventional and 3D-printed resins at baseline and after thermocycling (Baseline, BL; Thermocycled, AT). Materials with the same lowercase letters are not significantly different.</p>
Full article ">Figure 6
<p>Mean compressive strength of conventional and 3D-printed resins at baseline and after thermocycling (Baseline, BL; Thermocycled, AT). Materials with the same lowercase letters are not significantly different.</p>
Full article ">Figure 7
<p>Mean hardness of conventional and 3D-printed resins at baseline and after thermocycling (Baseline, BL; Thermocycled, AT). Materials with the same lowercase letters are not significantly different.</p>
Full article ">Figure 8
<p>Mean ΔE of conventional and 3D-printed resins at baseline and after thermocycling (Baseline, BL; Thermocycled, AT). Materials with the same lowercase letters are not significantly different.</p>
Full article ">Figure 9
<p>FTIR waveforms of the conventional and 3D-printed resins.</p>
Full article ">
Back to TopTop