[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
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 (3,602)

Search Parameters:
Keywords = reflectance spectroscopy

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
11 pages, 1951 KiB  
Article
Optimization of Non-Alloyed Backside Ohmic Contacts to N-Face GaN for Fully Vertical GaN-on-Silicon-Based Power Devices
by Youssef Hamdaoui, Sofie S. T. Vandenbroucke, Sondre Michler, Katir Ziouche, Matthias M. Minjauw, Christophe Detavernier and Farid Medjdoub
Micromachines 2024, 15(9), 1157; https://doi.org/10.3390/mi15091157 (registering DOI) - 15 Sep 2024
Abstract
In the framework of fully vertical GaN-on-Silicon device technology development, we report on the optimization of non-alloyed ohmic contacts on the N-polar n+-doped GaN face backside layer. This evaluation is made possible by using patterned TLMs (Transmission Line Model) through direct laser writing [...] Read more.
In the framework of fully vertical GaN-on-Silicon device technology development, we report on the optimization of non-alloyed ohmic contacts on the N-polar n+-doped GaN face backside layer. This evaluation is made possible by using patterned TLMs (Transmission Line Model) through direct laser writing lithography after locally removing the substrate and buffer layers in order to access the n+-doped backside layer. As deposited non-alloyed metal stack on top of N-polar orientation GaN layer after buffer layers removal results in poor ohmic contact quality. To significantly reduce the related specific contact resistance, an HCl treatment is applied prior to metallization under various time and temperature conditions. A 3 min HCl treatment at 70 °C is found to be the optimum condition to achieve thermally stable high ohmic contact quality. To further understand the impact of the wet treatment, SEM (Scanning Electron Microscopy) and XPS (X-ray Photoelectron Spectroscopy) analyses were performed. XPS revealed a decrease in Ga-O concentration after applying the treatment, reflecting the higher oxidation susceptibility of the N-polar face compared to the Ga-polar face, which was used as a reference. SEM images of the treated samples show the formation of pyramids on the N-face after HCl treatment, suggesting specific wet etching planes of the GaN crystal from the N-face. The size of the pyramids is time-dependent; thus, increasing the treatment duration results in larger pyramids, which explains the degradation of ohmic contact quality after prolonged high-temperature HCl treatment. Full article
(This article belongs to the Section D1: Semiconductor Devices)
19 pages, 1822 KiB  
Article
Renal Cell Carcinoma Discrimination through Attenuated Total Reflection Fourier Transform Infrared Spectroscopy of Dried Human Urine and Machine Learning Techniques
by Bogdan Adrian Buhas, Lucia Ana-Maria Muntean, Guillaume Ploussard, Bogdan Ovidiu Feciche, Iulia Andras, Valentin Toma, Teodor Andrei Maghiar, Nicolae Crișan, Rareș-Ionuț Știufiuc and Constantin Mihai Lucaciu
Int. J. Mol. Sci. 2024, 25(18), 9830; https://doi.org/10.3390/ijms25189830 - 11 Sep 2024
Viewed by 338
Abstract
Renal cell carcinoma (RCC) is the sixth most common cancer in men and is often asymptomatic, leading to incidental detection in advanced disease stages that are associated with aggressive histology and poorer outcomes. Various cancer biomarkers are found in urine samples from patients [...] Read more.
Renal cell carcinoma (RCC) is the sixth most common cancer in men and is often asymptomatic, leading to incidental detection in advanced disease stages that are associated with aggressive histology and poorer outcomes. Various cancer biomarkers are found in urine samples from patients with RCC. In this study, we propose to investigate the use of Attenuated Total Reflection-Fourier Transform Infrared Spectroscopy (ATR-FTIR) on dried urine samples for distinguishing RCC. We analyzed dried urine samples from 49 patients with RCC, confirmed by histopathology, and 39 healthy donors using ATR-FTIR spectroscopy. The vibrational bands of the dried urine were identified by comparing them with spectra from dried artificial urine, individual urine components, and dried artificial urine spiked with urine components. Urea dominated all spectra, but smaller intensity peaks, corresponding to creatinine, phosphate, and uric acid, were also identified. Statistically significant differences between the FTIR spectra of the two groups were obtained only for creatinine, with lower intensities for RCC cases. The discrimination of RCC was performed through Principal Component Analysis combined with Linear Discriminant Analysis (PCA–LDA) and Support Vector Machine (SVM). Using PCA–LDA, we achieved a higher discrimination accuracy (82%) (using only six Principal Components to avoid overfitting), as compared to SVM (76%). Our results demonstrate the potential of urine ATR-FTIR combined with machine learning techniques for RCC discrimination. However, further studies, especially of other urological diseases, must validate this approach. Full article
(This article belongs to the Special Issue Machine Learning in Disease Diagnosis and Treatment)
Show Figures

Figure 1

Figure 1
<p>ATR-FTIR spectrum of artificial urine (black) and the mean spectrum obtained from the urine of 39 control patients (red). The wavenumbers corresponding to the main peaks in the two sets of spectra are also indicated in cm<sup>−1</sup>.</p>
Full article ">Figure 2
<p>Comparison of the ATR-FTIR spectrum of artificial urine (black) with the spectra of the main organic urine components: urea (red), creatinine (blue), and uric acid (magenta). The vertical lines were traced to help identify the peaks of artificial urine with the peaks of the three components. The line and peak wavenumber colors indicate the compound for which we have the best match, and the black lines are traced for artificial urine peaks not matching the peaks of urea, creatinine, or uric acid.</p>
Full article ">Figure 3
<p>Matrix plot of the correlations between the ATR-FTIR absorption intensities measured for all the urine samples.</p>
Full article ">Figure 4
<p>The mean ATR-FTIR spectrum of urine from the RCC patients (red) and the healthy donors (CTRL) (blue) and the difference between the two mean spectra (black). Dashed areas represent the standard deviations. The difference spectrum was offset for better visualization.</p>
Full article ">Figure 5
<p>Loading plot for PC1 (black), PC2 (red), and PC4 (blue).</p>
Full article ">Figure 6
<p>Discrimination plot between the RCC and CTRL samples using a quadratic discrimination function and taking 15 PCs. For each sample, the software provides a score for the two groups CTRL and RCC. The sample is assigned to the group for which the score is highest. From a graphical point of view, the bi-dimensional space is split in two by the bisector of the first quadrant. The data points situated to the right from this bisector belong to the RCC group and the data points situated to the left from this bisector are assigned to the CTRL group. One can notice that three RCC cases (red circles) were assigned to the CTRL group (False Negative) and four CTRL samples (blue squares) were assigned to the RCC group (False Positive). The misassigned samples were marked with arrows. From 88 samples, 81 were assigned correctly, i.e., the accuracy was 92.05%.</p>
Full article ">Figure 7
<p>Accuracy of discrimination between the RCC and CTRL samples as a function of the number of PCs considered for the linear, quadratic, and Mahalanobis functions. The PCs were chosen in the order of their difference between the two groups (increasing the <span class="html-italic">p</span>, Pearson’s coefficient, from the Student’s <span class="html-italic">t</span>-Test, <a href="#ijms-25-09830-t003" class="html-table">Table 3</a>).</p>
Full article ">
20 pages, 3254 KiB  
Article
Acai Oil-Based Organogel Containing Hyaluronic Acid for Topical Cosmetic: In Vitro and Ex Vivo Assessment
by Suellen Christtine da Costa Sanches, Lindalva Maria de Meneses Costa Ferreira, Rayanne Rocha Pereira, Desireé Gyles Lynch, Ingryd Nayara de Farias Ramos, André Salim Khayat, José Otávio Carrera Silva-Júnior, Alessandra Rossi and Roseane Maria Ribeiro-Costa
Pharmaceutics 2024, 16(9), 1195; https://doi.org/10.3390/pharmaceutics16091195 - 11 Sep 2024
Viewed by 452
Abstract
Organogels are semi-solid pharmaceutical forms whose dispersing phase is an organic liquid, for example, an oil, such as acai oil, immobilized by a three-dimensional network formed by the gelling agent. Organogels are being highlighted as innovative release systems for cosmetic active ingredients such [...] Read more.
Organogels are semi-solid pharmaceutical forms whose dispersing phase is an organic liquid, for example, an oil, such as acai oil, immobilized by a three-dimensional network formed by the gelling agent. Organogels are being highlighted as innovative release systems for cosmetic active ingredients such as hyaluronic acid for topical applications. Acai oil was evaluated for its physicochemical parameters, fatty acid composition, lipid quality index, spectroscopic pattern (Attenuated total reflectance Fourier Transform Infrared Spectroscopy), thermal behavior, total phenolic, total flavonoids, and total carotenoids and β-carotene content. The effectiveness of the organogel incorporated with hyaluronic acid (OG + HA) was evaluated through ex vivo permeation and skin retention tests, in vitro tests by Attenuated total reflectance Fourier Transform Infrared Spectroscopy and Differential Scanning Calorimetry. The physicochemical analyses highlighted that the acai oil exhibited quality standards in agreement with the regulatory bodies. Acai oil also showed high antioxidant capacity, which was correlated with the identified bioactive compounds. The cytotoxicity tests demonstrated that the formulation OG + HA does not release toxic substances into the biological environment that could impede cell growth, adhesion, and efficacy. In vitro and ex vivo analyses demonstrated that after 6 h of application, OG + HA presented a high level of hydration, thermal protection and release of HA. Thus, it can be concluded that the OG + HA formulation has the potential for physical–chemical applications, antioxidant quality, and potentially promising efficacy for application in the cosmetic areas. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>ATR-FTIR spectrum of acai oil.</p>
Full article ">Figure 2
<p>Thermogravimetric and derived thermogravimetry curves of acai oil.</p>
Full article ">Figure 3
<p>Thermogram of the acai oil.</p>
Full article ">Figure 4
<p>Effect of different concentrations of OG + HA on the viability of fibroblast cells of the MRC5 lineage incubated for 48 h at 37 °C in DMEM. Data were expressed as mean ± standard deviation (n = 3). ***: significative values for <span class="html-italic">p</span> &lt; 0.0001 in relation to the control and other concentrations of OG + HA. NS: not significative values among themselves or compared to the control.</p>
Full article ">Figure 5
<p>Effect of OG + HA on pig skin fragments after different retention times in stratum corneum and epidermis/dermis.</p>
Full article ">Figure 6
<p>ATR-FTIR spectra of pig membranes after different retention times of OG + HA.</p>
Full article ">Figure 7
<p>Effect of OG + HA on pig skin fragments after different permeation/retention times. Data were expressed as mean ± standard deviation (n = 3). NS: not significant values among themselves or compared to the control.</p>
Full article ">Figure A1
<p>Ethics Committee for the use of Animals.</p>
Full article ">
21 pages, 4751 KiB  
Article
Green Synthesis of LaMnO3-Ag Nanocomposites Using Citrus limon (L.) Burm Peel Aqueous Extract: Photocatalytic Degradation of Rose Bengal Dye and Antibacterial Applications
by Nazim Hasan
Catalysts 2024, 14(9), 609; https://doi.org/10.3390/catal14090609 - 11 Sep 2024
Viewed by 265
Abstract
Perovskites can absorb solar energy and are extensively used in various catalytic and photocatalytic reactions. However, noble metal particles may enhance the catalytic, photocatalytic, and antibacterial activities. This study demonstrates the cost-effective green synthesis of the photocatalyst perovskite LaMnO3 and its modification [...] Read more.
Perovskites can absorb solar energy and are extensively used in various catalytic and photocatalytic reactions. However, noble metal particles may enhance the catalytic, photocatalytic, and antibacterial activities. This study demonstrates the cost-effective green synthesis of the photocatalyst perovskite LaMnO3 and its modification with noble metal Ag nanoparticles. The green synthesis of nanocomposite was achieved through a hydrothermal method employing aqueous extract derived from Citrus limon (L.) Burm peels. The properties of fabricated perovskites LaMnO3 and LaMnO3-Ag nanocomposites were evaluated and characterized by Ultraviolet-Visible spectroscopy (UV-Vis), Diffuse Reflectance Spectroscopy (DRS), X-ray diffraction (XRD), Fourier-Transform Infrared Spectroscopy (FT-IR), High-Resolution Transmission Electron Microscopy (HRTEM), Scanning Electron Microscopy (SEM), Energy-Dispersive X-ray spectroscopy (EDX) and Brunauer–Emmett–Teller (BET) surface area techniques. The particle size distribution % of LaMnO3 and LaMnO3-Ag was observed to be 20 to 60 nm after using TEM images. The maximum percentage size distribution was 37 nm for LaMnO3 and 43 nm for LaMnO3-Ag. In addition, LaMnO3-Ag nanocomposite was utilized as a photocatalyst for the degradation of Rose Bengal (RB) dye and its antibacterial activities against Staphylococcus aureus (S. aureus) and Escherichia coli (E. coli). The surface area and band gap for perovskite LaMnO3 nanoparticles were calculated as 12.642 m2/g and 3.44 eV, respectively. The presence of noble metal and hydrothermal-bio reduction significantly impacted the crystallinity. The BET surface area was found to be 16.209 m2/g, and band gap energy was calculated at 2.94 eV. The LaMnO3 nanocomposite with noble metal shows enhanced photocatalytic effectiveness against RB dye (20 PPM) degradation (92%, R2 = 0.995) with pseudo-first-order chemical kinetics (rate constant, k = 0.05057 min−1) within 50 min due to the ultimate combination of the hydrothermal and bio-reduction technique. The photocatalytic activity of the LaMnO3-Ag nanocomposite was optimized at different reaction times, photocatalyst doses (0.2, 0.4, 0.6, and 0.8 g/L), and various RB dye concentrations (20, 30, 40, and 50 ppm). The antibacterial activities of green synthesized LaMnO3 and LaMnO3-Ag nanoparticles were explored based on colony-forming unit (cfu) reduction and TEM images of bacterial and nanoparticle interactions for S. aureus and E. coli. An amount of 50 µg/mL LaMnO3-Ag nanocomposite was sufficient to work as the highest antibacterial activity for both bacteria. The perovskite LaMnO3-Ag nanocomposite synthesis process is economically and environmentally friendly. Additionally, it has a wide range of effective and exclusive applications for remediating pollutants. Full article
Show Figures

Figure 1

Figure 1
<p>XRD pattern of the green-synthesized (<b>a</b>) LaMnO<sub>3</sub>-Ag and (<b>b</b>) LaMnO<sub>3</sub> nanocomposites (XRD peaks are highlighted and matched), (<b>c</b>) ICPDS reference peaks of LaMnO<sub>3</sub> with card number 00-050-0297, (<b>d</b>) ICPDS reference peaks of Ag with card number 00-004-0783, and W–H plot for (<b>e</b>) LaMnO<sub>3</sub>-Ag and (<b>f</b>) LaMnO<sub>3</sub> nanocomposite.</p>
Full article ">Figure 2
<p>UV-Vis DRS spectra of LaMnO<sub>3</sub> and LaMnO<sub>3</sub>-Ag nanocomposites (<b>a</b>), band gap energy (eV) of LaMnO<sub>3</sub> (<b>b</b>), band gap of LaMnO<sub>3</sub>-Ag nanocomposite (<b>c</b>).</p>
Full article ">Figure 3
<p>TEM and HRTEM images of LaMnO<sub>3</sub> (<b>a</b>,<b>b</b>), LaMnO<sub>3</sub>-Ag (<b>c</b>,<b>d</b>), and their respective particle size distribution %.</p>
Full article ">Figure 4
<p>HRSEM images; (<b>a</b>) LaMnO<sub>3</sub> and (<b>b</b>) LaMnO<sub>3</sub>-EDX, (<b>c</b>) LaMnO<sub>3</sub>-Ag and (<b>d</b>) LaMnO<sub>3</sub>-Ag EDX.</p>
Full article ">Figure 5
<p>FTIR spectra of the synthesized LaMnO<sub>3</sub> and LaMnO<sub>3</sub>-Ag nanocomposites.</p>
Full article ">Figure 6
<p>N<sub>2</sub> adsorption–desorption isotherms ((<b>a</b>) LaMnO<sub>3</sub> and (<b>b</b>) LaMnO<sub>3</sub>-Ag nanocomposites) and pore size distribution patterns ((<b>c</b>) LaMnO<sub>3</sub> and (<b>d</b>) LaMnO<sub>3</sub>-Ag nanocomposites).</p>
Full article ">Figure 7
<p>(<b>a</b>) Photodegradation percentage efficiency, (<b>b</b>) apparent rate constant for the degradation of different ppm of dye, (<b>c</b>) apparent rate constant for the degradation of 20 ppm dye after using various amounts of LaMnO<sub>3</sub> nanocomposite, (<b>d</b>) recyclability of LaMnO<sub>3</sub>-Ag nanocomposite as catalyst, (<b>e</b>) UV-Vis spectra of RB dye (20 PPM, 50 mL) degradation using 40 mg LaMnO<sub>3</sub>-Ag nanocomposite catalyst, (<b>f</b>) scavenging agents effect on RB dye degradation.</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>) Photodegradation percentage efficiency, (<b>b</b>) apparent rate constant for the degradation of different ppm of dye, (<b>c</b>) apparent rate constant for the degradation of 20 ppm dye after using various amounts of LaMnO<sub>3</sub> nanocomposite, (<b>d</b>) recyclability of LaMnO<sub>3</sub>-Ag nanocomposite as catalyst, (<b>e</b>) UV-Vis spectra of RB dye (20 PPM, 50 mL) degradation using 40 mg LaMnO<sub>3</sub>-Ag nanocomposite catalyst, (<b>f</b>) scavenging agents effect on RB dye degradation.</p>
Full article ">Figure 8
<p>Schematic diagram of the reaction mechanism involved in the photocatalytic activity of LaMnO<sub>3</sub>-Ag nanocomposites.</p>
Full article ">Figure 9
<p>TEM images of <span class="html-italic">S. aureus</span> and <span class="html-italic">E. coli</span> interaction with LaMnO<sub>3</sub>-Ag nanoparticles. Images (<b>a</b>) and (<b>d</b>) show control bacterial cells, (<b>b</b>,<b>e</b>) represent the interaction of LaMnO<sub>3</sub>-Ag NPs towards <span class="html-italic">S. aureus</span> and <span class="html-italic">E. coli</span>, respectively. In contrast (<b>c</b>,<b>f</b>) represent <span class="html-italic">S. aureus</span> and <span class="html-italic">E. coli</span>. cell debris after the antibacterial effects of LaMnO<sub>3</sub>-Ag NPs.</p>
Full article ">Figure 10
<p>Schematic preparation of LaMnO<sub>3</sub> and LaMnO<sub>3</sub>-Ag nanocomposites.</p>
Full article ">
17 pages, 10970 KiB  
Article
Disclosing Colors and Pigments on Archaeological Objects from the Aga Khan Necropolis (West Aswan Egypt) through On-Site Analytical Methods: Preliminary Results
by Paola Fermo, Chiara Andrea Lombardi, Alfonsina D’Amato, Vittoria Guglielmi, Benedetta Giudici, Alice Tomaino, Massimiliana Pozzi, Valeria Comite, Andrea Bergomi, Lorenzo Guardiano and Patrizia Piacentini
Heritage 2024, 7(9), 4980-4996; https://doi.org/10.3390/heritage7090235 - 9 Sep 2024
Viewed by 290
Abstract
The present study is aimed at the characterization of artifacts excavated in the necropolis surrounding the mausoleum of the Aga Khan in Aswan (Egypt), as part of the Mummies Investigations Anthropological & Scientific West Aswan Necropolis (MIASWAN) project. Four cartonnages and some pottery [...] Read more.
The present study is aimed at the characterization of artifacts excavated in the necropolis surrounding the mausoleum of the Aga Khan in Aswan (Egypt), as part of the Mummies Investigations Anthropological & Scientific West Aswan Necropolis (MIASWAN) project. Four cartonnages and some pottery shards were investigated on-site by means of non-destructive and micro-destructive techniques, such as attenuated total reflection/Fourier transform infrared spectroscopy (ATR/FTIR) and visible reflectance spectroscopy Vis-RS). Thanks to the use of these techniques, several pigments employed in the creation of the artifacts were identified. Due to the impossibility of transporting the investigated objects out of Egypt, a first-ever on-site characterization of the artifacts from this important excavation was carried out through scientific methodologies. These extreme conditions made the use of analytical instrumentation very challenging. Nevertheless, several characteristic pigments and hues were successfully identified. Full article
(This article belongs to the Section Archaeological Heritage)
Show Figures

Figure 1

Figure 1
<p>Some of the researchers at work on the excavation (on the <b>left</b>) and the entrance of tomb AGH026 (on the <b>right</b>).</p>
Full article ">Figure 2
<p>FT-IR spectroscopy measurements at the warehouse where the objects were moved.</p>
Full article ">Figure 3
<p>ATR/FTIR spectra of sample 1-B14 and 2-B14. Marker bands of calcite (+), hydroxyl group (*) organic matter (°), water (▼) silicates (^)and ochre (§) are highlighted.</p>
Full article ">Figure 4
<p>Observation of the surface of <span class="html-italic">cartonnage</span> AGH026-B14 carried out by means of a USB microscope (on the <b>left</b>) and an image acquired on a black area where blue grains mixed with the black pigment are evident (on the <b>right</b>).</p>
Full article ">Figure 5
<p>ATR/FTIR spectra ofAGH026-B13 sample. Marker bands of calcite (+), Egyptian blue (•), hydroxyl group (*), water (▼), silicates (^), red ochre (§), alizarine (x) (with the signal at 1100 cm<sup>−1</sup> overlaid to gypsum), and yellow ochre (▲) are highlighted.</p>
Full article ">Figure 6
<p>ATR/FTIR spectra ofAGH026-B31sample. Marker bands of calcite (+), Egyptian blue (•), hydroxyl group (*), water (▼), silicates (^), and red ochre (§) are highlighted.</p>
Full article ">Figure 7
<p><span class="html-italic">Cartonnage</span> AGH026-B41 and the list of the points analyzed by portable visible reflectance spectroscopy.</p>
Full article ">Figure 8
<p>Visible reflectance spectra acquired on the points (<b>1</b>–<b>12c</b>) indicated in <a href="#heritage-07-00235-f007" class="html-fig">Figure 7</a> and respective first derivative of all samples of <span class="html-italic">cartonnage</span> AGH026-B41.</p>
Full article ">Figure 9
<p>Scatter plot obtained after PCA analysis carried out on visible reflectance spectra acquired on different points on <span class="html-italic">cartonnage</span> AGH026-B41.</p>
Full article ">Figure 10
<p>Pottery AGH032-14 and the list of the points analyzed by Vis-RS.</p>
Full article ">Figure 11
<p>Visible reflectance spectra of all samples of pottery acquired on the points (<b>POT1</b>–<b>6</b>) indicated in <a href="#heritage-07-00235-f010" class="html-fig">Figure 10</a>.</p>
Full article ">
9 pages, 2010 KiB  
Communication
High-Sensitivity and In Situ Multi-Component Detection of Gases Based on Multiple-Reflection-Cavity-Enhanced Raman Spectroscopy
by Dewang Yang, Wenhua Li, Haoyue Tian, Zhigao Chen, Yuhang Ji, Hui Dong and Yongmei Wang
Sensors 2024, 24(17), 5825; https://doi.org/10.3390/s24175825 - 7 Sep 2024
Viewed by 457
Abstract
Raman spectroscopy with the advantages of the in situ and simultaneous detection of multi-components has been widely used in the identification and quantitative detection of gas. As a type of scattering spectroscopy, the detection sensitivity of Raman spectroscopy is relatively lower, mainly due [...] Read more.
Raman spectroscopy with the advantages of the in situ and simultaneous detection of multi-components has been widely used in the identification and quantitative detection of gas. As a type of scattering spectroscopy, the detection sensitivity of Raman spectroscopy is relatively lower, mainly due to the low signal collection efficiency. This paper presents the design and assembly of a multi-channel cavity-enhanced Raman spectroscopy system, optimizing the structure of the sample pool to reduce the loss of the laser and increase the excitation intensity of the Raman signals. Moreover, three channels are used to collect Raman signals to increase the signal collection efficiency for improving the detection sensitivity. The results showed that the limits of detection for the CH4, H2, CO2, O2, and N2 gases were calculated to be 3.1, 34.9, 17.9, 27, and 35.2 ppm, respectively. The established calibration curves showed that the correlation coefficients were all greater than 0.999, indicating an excellent linear correlation and high level of reliability. Meanwhile, under long-time integration detection, the Raman signals of CH4, H2, and CO2 could be clearly distinguished at the concentrations of 10, 10, and 50 ppm, respectively. The results indicated that the designed Raman system possesses broad application prospects in complex field environments. Full article
(This article belongs to the Section Optical Sensors)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of the Raman spectroscopy system and the optical path.</p>
Full article ">Figure 2
<p>Original Raman spectra of the mixed gases (Nos. 3, 6, 9, and 12) with 5× magnification for the ranges of 0–1100 cm<sup>−1</sup> and 3400–4400 cm<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Raman spectra of different gas samples CH<sub>4</sub> (<b>a</b>), H<sub>2</sub> (<b>b</b>) and CO<sub>2</sub> (<b>c</b>) under different concentrations and the calibration curves of gas concentration versus Raman peak intensity (<b>d</b>).</p>
Full article ">Figure 4
<p>Raman spectra of low concentrations of CH<sub>4</sub> (<b>a</b>), H<sub>2</sub> (<b>b</b>), and CO<sub>2</sub> (<b>c</b>) under long-time integration.</p>
Full article ">
11 pages, 13353 KiB  
Article
In Situ Studies on the Influence of Surface Symmetry on the Growth of MoSe2 Monolayer on Sapphire Using Reflectance Anisotropy Spectroscopy and Differential Reflectance Spectroscopy
by Yufeng Huang, Mengjiao Li, Zhixin Hu, Chunguang Hu, Wanfu Shen, Yanning Li and Lidong Sun
Nanomaterials 2024, 14(17), 1457; https://doi.org/10.3390/nano14171457 - 7 Sep 2024
Viewed by 361
Abstract
The surface symmetry of the substrate plays an important role in the epitaxial high-quality growth of 2D materials; however, in-depth and in situ studies on these materials during growth are still limited due to the lack of effective in situ monitoring approaches. In [...] Read more.
The surface symmetry of the substrate plays an important role in the epitaxial high-quality growth of 2D materials; however, in-depth and in situ studies on these materials during growth are still limited due to the lack of effective in situ monitoring approaches. In this work, taking the growth of MoSe2 as an example, the distinct growth processes on Al2O3 (112¯0) and Al2O3 (0001) are revealed by parallel monitoring using in situ reflectance anisotropy spectroscopy (RAS) and differential reflectance spectroscopy (DRS), respectively, highlighting the dominant role of the surface symmetry. In our previous study, we found that the RAS signal of MoSe2 grown on Al2O3 (112¯0) initially increased and decreased ultimately to the magnitude of bare Al2O3 (112¯0) when the first layer of MoSe2 was fully merged, which is herein verified by the complementary DRS measurement that is directly related to the film coverage. Consequently, the changing rate of reflectance anisotropy (RA) intensity at 2.5 eV is well matched with the dynamic changes in differential reflectance (DR) intensity. Moreover, the surface-dominated uniform orientation of MoSe2 islands at various stages determined by RAS was further investigated by low-energy electron diffraction (LEED) and atomic force microscopy (AFM). By contrast, the RAS signal of MoSe2 grown on Al2O3 (0001) remains at zero during the whole growth, implying that the discontinuous MoSe2 islands have no preferential orientations. This work demonstrates that the combination of in situ RAS and DRS can provide valuable insights into the growth of unidirectional aligned islands and help optimize the fabrication process for single-crystal transition metal dichalcogenide (TMDC) monolayers. Full article
Show Figures

Figure 1

Figure 1
<p>MoSe<sub>2</sub> thin film growth on Al<sub>2</sub>O<sub>3</sub> (<math display="inline"><semantics> <mrow> <mn>11</mn> <mover accent="true"> <mrow> <mn>2</mn> </mrow> <mo>¯</mo> </mover> <mn>0</mn> </mrow> </semantics></math>) at 530 °C monitored using optical reflection measurement is schematically presented in (<b>a</b>). The setup allows the simultaneous determination of the RA and DR spectrum in real time during the growth. (<b>b</b>) Two-dimensional contour map of the RA signal over photon energy (the horizontal axis) and deposition time (the vertical axis). The RA spectra recorded at selected deposition times of <span class="html-italic">t<sub>a</sub></span>~31 min, <span class="html-italic">t<sub>b</sub></span>~47 min, <span class="html-italic">t<sub>c</sub></span>~63 min, <span class="html-italic">t<sub>d</sub></span>~87 min, and <span class="html-italic">t<sub>e</sub></span>~140 min (indicated by the horizontal dashed lines in (<b>b</b>)), are presented in (<b>c</b>), whereas the black dotted line represents the initial RA spectrum of bare Al<sub>2</sub>O<sub>3</sub> (<math display="inline"><semantics> <mrow> <mn>11</mn> <mover accent="true"> <mrow> <mn>2</mn> </mrow> <mo>¯</mo> </mover> <mn>0</mn> </mrow> </semantics></math>) substrate. The evolution of the RA intensity at 2.5 eV (along the vertical dashed line in (<b>b</b>)) is plotted in (<b>d</b>) as a function of deposition time (solid blue line). In a similar fashion, the corresponding DR spectra are exhibited, namely, 2D contour map for an overview in (<b>e</b>), spectra recorded at <span class="html-italic">t<sub>a</sub></span>, <span class="html-italic">t<sub>b</sub></span>, <span class="html-italic">t<sub>c</sub></span>, <span class="html-italic">t<sub>d</sub></span>, and <span class="html-italic">t<sub>e</sub></span> in (<b>f</b>), and the variation of the DR signals as a function of the deposition time at 2.3 eV and 3.1 eV in (<b>d</b>). The first derivative curve of the change in RA intensity at 2.5 eV is also shown in (<b>d</b>).</p>
Full article ">Figure 2
<p>(<b>a</b>) The band structure of monolayer MoSe<sub>2</sub> calculated by DFT. The arrows indicate the transition in C and D from the valance band to the conduction band. (<b>b</b>) The imaginary part of the calculated dielectric function for monolayer MoSe<sub>2</sub>. The main features are labeled as A to D. (<b>c</b>) Representative DR spectra recorded from the growth of MoSe<sub>2</sub> on Al<sub>2</sub>O<sub>3</sub> (<math display="inline"><semantics> <mrow> <mn>11</mn> <mover accent="true"> <mrow> <mn>2</mn> </mrow> <mo>¯</mo> </mover> <mn>0</mn> </mrow> </semantics></math>) surface. The dark circles and blue lines represent the Lorentz fit of DR spectra recorded at different times. (<b>d</b>) The evolutions of center energy and integral area of peaks C and D as a function of deposition time.</p>
Full article ">Figure 3
<p>LEED results of as-grown MoSe<sub>2</sub> sample. (<b>a</b>) LEED pattern of MoSe<sub>2</sub> thin film (after the deposition time of <span class="html-italic">t<sub>e</sub></span>) using an electron beam energy of 135 eV. (<b>b</b>) The line profiles (dotted lines) taken across the six first-order diffraction spots on LEED patterns (indicated by the arrowed gray circle in (<b>a</b>)) obtained after various deposition times of <span class="html-italic">t<sub>b</sub></span>, <span class="html-italic">t<sub>c</sub></span>, <span class="html-italic">t<sub>d</sub></span>, and <span class="html-italic">t<sub>e</sub></span>, respectively. The lines are normalized and offset for clarity. (<b>c</b>) The variation of the FWHM obtained by fitting the experimental profiles (solid lines) in (<b>b</b>) as a function of the deposition time.</p>
Full article ">Figure 4
<p>AFM images of MoSe<sub>2</sub> films obtained after deposition times from <span class="html-italic">t<sub>a</sub></span> to <span class="html-italic">t<sub>e</sub></span> are displayed in (<b>a</b>–<b>e</b>), respectively. The white circle in (<b>b</b>) represents the coalescence of islands. (<b>f</b>) Height profiles for the green line in (<b>c</b>), the red line in (<b>d</b>), and the blue line in (<b>e</b>).</p>
Full article ">Figure 5
<p>Real-time monitoring for the MBE growth of MoSe<sub>2</sub> layer on Al<sub>2</sub>O<sub>3</sub> (0001) substrate at 530 °C. (<b>a</b>) Two-dimensional contour map of the RA signal over photon energy (the horizontal axis) and deposition time (the vertical axis). (<b>b</b>) RA spectra recorded during the growth of MoSe<sub>2</sub> layer on the bare Al<sub>2</sub>O<sub>3</sub> (0001) surface (marked by black dot line) from the beginning to the end (marked by black solid line) of the growth process. The time interval is about 5 min. The evolution of the RA intensity at 2.5 eV (along the vertical dashed line in (<b>a</b>)) as a function of the deposition time is plotted in (<b>c</b>) (solid green line). The inset shows the LEED pattern of MoSe<sub>2</sub> measured at an electron energy of 135 eV. The corresponding 2D contour map of DRS, the DR spectra over photon energy with about 5 min interval, and the evolution of DR intensity at 2.4 eV and 3.0 eV are shown in (<b>c</b>–<b>e</b>), respectively. (<b>f</b>) The corresponding Raman spectrum and the AFM image. The green curve represents the height profile across the green line.</p>
Full article ">
21 pages, 16002 KiB  
Article
Comparative Studies on Nanocellulose as a Bio-Based Consolidating Agent for Ancient Wood
by Anastasia Fornari, Daniele Rocco, Leonardo Mattiello, Martina Bortolami, Marco Rossi, Laura Bergamonti, Claudia Graiff, Stefania Bani, Fabio Morresi and Fabiana Pandolfi
Appl. Sci. 2024, 14(17), 7964; https://doi.org/10.3390/app14177964 - 6 Sep 2024
Viewed by 327
Abstract
In this work, nanocellulose aqueous dispersions were studied as a bio-inspired consolidating agent for the recovery and conservation of ancient wood and compared with two of the most used traditional consolidants: the synthetic resins Paraloid B-72 and Regalrez 1126. The morphology of crystalline [...] Read more.
In this work, nanocellulose aqueous dispersions were studied as a bio-inspired consolidating agent for the recovery and conservation of ancient wood and compared with two of the most used traditional consolidants: the synthetic resins Paraloid B-72 and Regalrez 1126. The morphology of crystalline nanocellulose (CNC), determined using Scanning Electron Microscopy (SEM), presents with a rod-like shape, with a size ranging between 15 and 30 nm in width. Chemical characterization performed using the Fourier-Transform Infrared Spectroscopy (FT-IR) technique provides information on surface modifications, in this case, demonstrating the presence of only the characteristic peaks of nanocellulose. Moreover, conductometric, pH, and dry matter measurements were carried out, showing also in this case values perfectly conforming to what is found in the literature. The treated wood samples were observed under an optical microscope in reflected light and under a scanning electron microscope to determine, respectively, the damage caused by xylophages and the morphology of the treated surfaces. The images acquired show the greater similarity of the surfaces treated with nanocellulose to untreated wood, compared with other consolidating agents. Finally, a colorimetric analysis of these samples was also carried out before and after a first consolidation treatment, and after a second treatment carried out on the same samples three years later. The samples treated with CNC appeared very homogeneous and uniform, without alterations in their final color appearance, compared to other traditional synthetic products. Full article
(This article belongs to the Special Issue Advanced Technologies in Cultural Heritage)
Show Figures

Figure 1

Figure 1
<p>SEM micrograph of CNC.</p>
Full article ">Figure 2
<p>FTIR spectrum of CNC.</p>
Full article ">Figure 3
<p>Wood samples treated with CNC (A), Paraloid B-72 (B), and Regalrez 1126 (C), and samples untreated (NT) during the processes of impregnation: (<b>1</b>) before treatment; (<b>2</b>) immediately after; (<b>3</b>) after 19 h; and (<b>4</b>) after 50 days.</p>
Full article ">Figure 4
<p>Detailed images of samples treated with (<b>a</b>) CNC, (<b>b</b>) Paraloid B-72, and (<b>c</b>) Regalrez 1126, immediately after application.</p>
Full article ">Figure 5
<p>Reflected light microscope images of samples that were not-treated (NT) and those treated with CNC (A), Paraloid B-72 (B), and Regalrez 1126 (C). Images with UV light and 5× magnification (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>); images with visible light and 2.5× magnification (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>). White arrows indicate in all samples the signs of degradation due to the action of xylophagous insects.</p>
Full article ">Figure 6
<p>Reflectance spectra of untreated (NT) and treated samples (CNC, Paraloid B-72, Regalrez 1126), acquired in SCI and SCE mode 24 h after the consolidation treatment.</p>
Full article ">Figure 7
<p>Reflectance spectra of untreated (NT) and treated samples (CNC, Paraloid B-72, Regalrez 1126), acquired in SCI and SCE mode, one month after the consolidation treatment.</p>
Full article ">Figure 8
<p>Reflectance spectra of untreated (NT) and treated samples (CNC, Paraloid B-72, Regalrez 1126), acquired in SCI and SCE mode, three years after the first consolidating treatment.</p>
Full article ">Figure 9
<p>Reflectance spectra of untreated (NT) and treated samples (CNC, Paraloid B-72, Regalrez 1126), acquired in SCI and SCE mode, one week after the second consolidating treatment, carried out three years after the first treatment.</p>
Full article ">Figure 10
<p>SEM images of untreated wood sample; cross section in correspondence of a woodworm hole (<b>a</b>); longitudinal sections (<b>b</b>,<b>c</b>); magnification of a fiber channel (<b>c</b>).</p>
Full article ">Figure 11
<p>SEM images of untreated sample (<b>1</b>) and of the consolidant coating films of CNC (<b>2</b>), Paraloid B-72 (<b>3</b>), and Regalrez 1126 (<b>4</b>).</p>
Full article ">Figure 12
<p>SEM images of longitudinal section of wood samples, where it is possible to see the fibers channels: untreated sample (<b>1</b>); sample treated with CNC (<b>2</b>), Paraloid B-72 (<b>3</b>), Regalrez 1126 (<b>4</b>).</p>
Full article ">
17 pages, 4400 KiB  
Article
Preparation of Composite Hydrogels Based on Cysteine–Silver Sol and Methylene Blue as Promising Systems for Anticancer Photodynamic Therapy
by Dmitry V. Vishnevetskii, Fedor A. Metlin, Yana V. Andrianova, Elizaveta E. Polyakova, Alexandra I. Ivanova, Dmitry V. Averkin and Arif R. Mekhtiev
Gels 2024, 10(9), 577; https://doi.org/10.3390/gels10090577 - 5 Sep 2024
Viewed by 429
Abstract
In this study, a novel supramolecular composite, “photogels”, was synthesized by mixing of cysteine–silver sol (CSS) and methylene blue (MB). A complex of modern physico-chemical methods of analysis such as viscosimetry, UV spectroscopy, dynamic and electrophoretic light scattering, scanning electron microscopy and energy-dispersive [...] Read more.
In this study, a novel supramolecular composite, “photogels”, was synthesized by mixing of cysteine–silver sol (CSS) and methylene blue (MB). A complex of modern physico-chemical methods of analysis such as viscosimetry, UV spectroscopy, dynamic and electrophoretic light scattering, scanning electron microscopy and energy-dispersive X-ray spectroscopy showed that MB molecules are uniformly localized mainly in the space between fibers of the gel-network formed by CSS particles. Molecules of the dye also bind with the surface of CSS particles by non-covalent interactions. This fact is reflected in the appearance of a synergistic anticancer effect of gels against human squamous cell carcinoma even in the absence of light irradiation. A mild toxic influence of hydrogels was observed in normal keratinocyte cells. Photodynamic exposure significantly increased gel activity, and there remained a synergistic effect. The study of free-radical oxidation in cells has shown that gels are not only capable of generating reactive oxygen species, but also have other targets of action. Flow cytometric analysis allowed us to find out that obtained hydrogels caused cell cycle arrest both without irradiation and with light exposure. The obtained gels are of considerable interest both from the point of view of academics and applied science, for example, in the photodynamic therapy of superficial neoplasms. Full article
(This article belongs to the Special Issue Synthesis and Applications of Hydrogels (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) The scheme of preparation of gels with MB. (<b>B</b>) The MB content in the gels (table) and photo of the gels. CSMBG—cysteine silver methylene blue gel.</p>
Full article ">Figure 2
<p>(<b>a</b>) Gel viscosity over time. (<b>b</b>) Gel viscosity dependence on MB concentration. See the sample numbers (1, 5, 11 and 17) in the table (<a href="#gels-10-00577-f001" class="html-fig">Figure 1</a>B).</p>
Full article ">Figure 3
<p>SEM images and EDS of gels: (<b>a</b>) no MB; (<b>b</b>–<b>e</b>) samples 1, 5, 11 and 17. See the sample numbers in the table in <a href="#gels-10-00577-f001" class="html-fig">Figure 1</a>B.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV spectra in various regions of wavelengths, (<b>b</b>) particle size distribution and cross-correlation functions and (<b>c</b>) zeta potential measurements for gels under study. See the sample numbers in the table in <a href="#gels-10-00577-f001" class="html-fig">Figure 1</a>B.</p>
Full article ">Figure 5
<p>The cytotoxicity (MTT) of hydrogels for (<b>a</b>,<b>b</b>) SiHa and (<b>c</b>,<b>d</b>) HaCat cells without irradiation (<b>a</b>,<b>c</b>) and at a laser exposure of 638 nm (<b>b</b>,<b>d</b>). SiHa and HaCat cells’ incubation time with the systems is 48 h (n = 3). The information about the samples is given in the table in <a href="#gels-10-00577-f001" class="html-fig">Figure 1</a>B.</p>
Full article ">Figure 6
<p>The ROS generation in (<b>a,b</b>) SiHa and (<b>c,d</b>) HaCat cells treated by hydrogels without irradiation (<b>a</b>,<b>c</b>) and at a laser exposure of 638 nm (<b>b</b>,<b>d</b>). SiHa and HaCat cells’ incubation time with the systems is 48 h (n = 4). The information about the samples is given in the table in <a href="#gels-10-00577-f001" class="html-fig">Figure 1</a>B.</p>
Full article ">Figure 7
<p>The flow cytometric analysis of DNA content in (<b>a</b>,<b>b</b>) SiHa and (<b>c</b>,<b>d</b>) HaCat cells treated by hydrogel (0—no MB) without irradiation (<b>a</b>,<b>c</b>) and at a laser exposure of 638 nm (<b>b</b>,<b>d</b>). SiHa and HaCat cells’ incubation time with the systems is 48 h. The concentration of the samples is 30 µM. The subG0–G1 peak on the histogram indicates apoptotic cells; G0–G1, S and G2-M—phases of the cell cycle.</p>
Full article ">
17 pages, 4184 KiB  
Article
Enhanced Photocatalytic Degradation of Herbicide 2,4-Dichlorophenoxyacetic Acid Using Sulfated CeO2
by Carlos Rodríguez, Claudia Castañeda, Edwin Sosa, José J. Martínez, Sonia Mancipe, Hugo Rojas, Francisco Tzompantzi and Ricardo Gómez
Catalysts 2024, 14(9), 594; https://doi.org/10.3390/catal14090594 - 4 Sep 2024
Viewed by 442
Abstract
The present study presents the results obtained from evaluating the photocatalytic behavior of a series of sulfated CeO2 materials in the photocatalytic degradation of the herbicide 2,4-dichlorophenoxyacetic acid. The CeO2 photocatalytic support was prepared using the precipitation synthesis method. Subsequently, the [...] Read more.
The present study presents the results obtained from evaluating the photocatalytic behavior of a series of sulfated CeO2 materials in the photocatalytic degradation of the herbicide 2,4-dichlorophenoxyacetic acid. The CeO2 photocatalytic support was prepared using the precipitation synthesis method. Subsequently, the support was wetly impregnated with different contents of sulfate ions (0.5, 1.0, and 2.0 wt.%). The materials were characterized using X-ray diffraction, nitrogen physisorption, infrared spectroscopy, diffuse reflectance UV–Vis spectrophotometry, and thermal analysis. The characterization results showed that the sulfation of the material promoted an increase in the surface area and a decrease in the average size of the crystallites. Likewise, it was possible to demonstrate the surface sulfation of the support through bidentate coordination of the sulfate groups to the semiconductor metal. Concerning photoactivity, the convenience of the surface modification of CeO2 was confirmed because the sulfate groups acted as capturers of the electrons generated during the photocatalytic process, reducing the frequency of recombination of the charge carriers and allowing the availability of the gaps to favor the degradation reaction of the contaminant. Finally, it was evident that a percentage of 1.0 wt.% of the sulfate anion is the optimal content to improve the photocatalytic properties of CeO2. Full article
(This article belongs to the Special Issue Catalytic Energy Conversion and Catalytic Environmental Purification)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction results of the materials studied.</p>
Full article ">Figure 2
<p>N<sub>2</sub> adsorption–desorption isotherms of the materials studied.</p>
Full article ">Figure 3
<p>ATR-FTIR spectra of the materials studied.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV–Vis spectra and (<b>b</b>) Tauc plot of the materials studied.</p>
Full article ">Figure 5
<p>TGA-MS results of (<b>a</b>) uncalcined CeO<sub>2</sub> support and (<b>b</b>) 1.0SO<sub>4</sub><sup>2−</sup>/CeO<sub>2</sub> photocatalyst.</p>
Full article ">Figure 6
<p>Photoluminescence spectra of the materials studied.</p>
Full article ">Figure 7
<p>(<b>a</b>) UV spectra of the 2,4-D photodegradation using 1.0SO<sub>4</sub><sup>2−</sup>/CeO<sub>2</sub> photocatalyst, and (<b>b</b>) variation of the relative concentration of 2,4-D as a function of reaction time using materials studied.</p>
Full article ">Figure 8
<p>(<b>a</b>) UV spectra of the 2,4-D adsorption using 1.0SO<sub>4</sub><sup>2−</sup>/CeO<sub>2</sub> photocatalyst. (<b>b</b>) Comparison between the adsorption reaction and the photocatalytic degradation of the 2,4-D herbicide.</p>
Full article ">Figure 9
<p>Representation of the photocatalytic process in the degradation of 2,4-D using the 1.0SO<sub>4</sub><sup>2−</sup>/CeO<sub>2</sub> material.</p>
Full article ">Figure 10
<p>(<b>a</b>) Surface plots and (<b>b</b>) contour plots representing the effect of 2,4-D concentration and photocatalyst mass on the degradation percentage.</p>
Full article ">Figure 11
<p>Reuse of the 1.0SO<sub>4</sub><sup>2−</sup>/CeO<sub>2</sub> photocatalyst in the photocatalytic degradation of 2,4-dichlorophenoxyacetic acid (2,4-D).</p>
Full article ">Figure 12
<p>ATR-FTIR spectra of the 1.0SO<sub>4</sub><sup>2−</sup>/CeO<sub>2</sub> material reused in the cycle experiments and the material used in the adsorption test.</p>
Full article ">
17 pages, 2698 KiB  
Article
Assessment of Optical and Phonon Characteristics in MOCVD-Grown (AlxGa1−x)0.5In0.5P/n+-GaAs Epifilms
by Devki N. Talwar and Zhe Chuan Feng
Molecules 2024, 29(17), 4188; https://doi.org/10.3390/molecules29174188 - 4 Sep 2024
Viewed by 435
Abstract
Quaternary (AlxGa1−x)yIn1−yP alloys grown on GaAs substrates have recently gained considerable interest in photonics for improving visible light-emitting diodes, laser diodes, and photodetectors. With two degrees of freedom (x, y) and keeping growth on a [...] Read more.
Quaternary (AlxGa1−x)yIn1−yP alloys grown on GaAs substrates have recently gained considerable interest in photonics for improving visible light-emitting diodes, laser diodes, and photodetectors. With two degrees of freedom (x, y) and keeping growth on a lattice-matched GaAs substrate, the (AlxGa1−x)0.5In0.5P alloys are used for tuning structural, phonon, and optical characteristics in different energy regions from far-infrared (FIR) → near-infrared (NIR) → ultraviolet (UV). Despite the successful growth of (AlxGa1−x)0.5In0.5P/n+-GaAs epilayers, limited optical, phonon, and structural characteristics exist. Here, we report our results of carefully examined optical and vibrational properties on highly disordered alloys using temperature-dependent photoluminescence (TD-PL), Raman scattering spectroscopy (RSS), and Fourier-transform infrared reflectivity (FTIR). Macroscopic models were meticulously employed to analyze the TD-PL, RSS, and FTIR data of the (Al0.24Ga0.76)0.5In0.5P/n+-GaAs epilayers to comprehend the energy-dependent characteristics. The Raman scattering and FTIR results of phonons helped analyze the reflectivity spectra in the FIR region. Optical constants were carefully integrated in the transfer matrix method for evaluating the reflectivity R(E) and transmission T(E) spectra in the NIR → UV regions, validating the TD-PL measurements of bandgap energies (EgPL). Full article
(This article belongs to the Special Issue Chemical Research on Photosensitive Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Temperature-dependent photoluminescence measurements exhibiting a shift in intensity peaks of electronic energy bandgaps (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="normal">E</mi> </mrow> <mrow> <mi mathvariant="normal">g</mi> </mrow> <mrow> <mi mathvariant="normal">P</mi> <mi mathvariant="normal">L</mi> </mrow> </msubsup> </mrow> </semantics></math>) from 2.1 eV to 2.03 eV, in good agreement with the experimental/theoretical results. (<b>b</b>) Comparison of the experimental (open red color square) energy bandgap of the (Al<sub>0.24</sub>Ga<sub>0.76</sub>)<sub>0.50</sub>In<sub>0.50</sub>P alloy with Varshni’s (Ref. [<a href="#B70-molecules-29-04188" class="html-bibr">70</a>]) formula (blue open circles) using an appropriate set of values for the adjustable parameters α and β (see text).</p>
Full article ">Figure 2
<p>Composition-dependent energy bandgap of (Al<sub>x</sub>Ga<sub>1−x</sub>)<sub>0.5</sub>In<sub>0.5</sub>P alloys as a function of composition by using Equation (2) following Ref. [<a href="#B71-molecules-29-04188" class="html-bibr">71</a>]. The PL result of E<sub>g</sub> at x = 0.24 is shown by red color open square.</p>
Full article ">Figure 3
<p>Raman scattering spectra of four MOCVD-grown (Al<sub>x</sub>Ga<sub>1−x</sub>)<sub>0.5</sub>In<sub>0.5</sub>P/n<sup>+</sup>-GaAs epilayers with x = 0.24. These measurements were performed by using a Renishaw Raman Microscope model-100 in the backscattering geometry with a 633 nm laser beam from a He–Ne source. Vertical magenta-colored arrows are used to indicate the TO mode of the GaAs (substrate) as well as the AlP-, GaP-, and InP-like LO and TO modes of the quaternary alloy (see text).</p>
Full article ">Figure 4
<p>(<b>a</b>) Experimental reflectivity spectra using a high-resolution Brüker IFS 120 v/S FTIR spectrometer on four MOCVD-grown (Ga<sub>1−x</sub>Al<sub>x</sub>)<sub>0.5</sub>In<sub>0.5</sub>P/GaAs T<sub>i</sub> (i = 1, 4) samples with x = 0.24 (see <a href="#molecules-29-04188-t001" class="html-table">Table 1</a>). Magenta vertical arrows identify the major optical modes. (<b>b</b>) Experimental (calculated) reflectivity spectra of n<sup>+</sup> GaAs substrate using blue-colored open circles (red-colored line) (see text).</p>
Full article ">Figure 5
<p>Comparison of the simulated (violet-colored solid line) IR reflectivity spectra <math display="inline"><semantics> <mrow> <mi>R</mi> <mfenced separators="|"> <mrow> <mi>ω</mi> </mrow> </mfenced> </mrow> </semantics></math> with the experimental (black-colored inverted triangles) data for the (Al<sub>x</sub>Ga<sub>1−x</sub>)<sub>0.5</sub>In<sub>0.5</sub>P/n<sup>+</sup>-GaAs epilayer (sample T<sub>1</sub>). The simulation was performed using a three-phase ‘air–film–substrate’ model (Ref. [<a href="#B64-molecules-29-04188" class="html-bibr">64</a>]), while experimental data were obtained by a high-resolution Brüker IFS 120 v/S FTIR spectrometer (see text).</p>
Full article ">Figure 6
<p>(<b>a</b>) Simulation of E-dependent <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">ε</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> <mo>(</mo> <mi>E</mi> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">ε</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> <mo>(</mo> <mi>E</mi> <mo>)</mo> </mrow> </semantics></math> for the (Al<sub>x</sub>Ga<sub>1−x</sub>)<sub>0.5</sub>In<sub>0.5</sub>P alloys with an increment of composition x. (<b>b</b>) Simulation of E-dependent <math display="inline"><semantics> <mrow> <mi mathvariant="normal">n</mi> <mo>(</mo> <mi>E</mi> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">κ</mi> <mo>(</mo> <mi>E</mi> <mo>)</mo> </mrow> </semantics></math> for the (Al<sub>x</sub>Ga<sub>1−x</sub>)<sub>0.5</sub>In<sub>0.5</sub>P alloys with an increment of composition x.</p>
Full article ">Figure 7
<p>Transfer-matrix-based simulated spectra of reflectance <math display="inline"><semantics> <mrow> <mi mathvariant="normal">R</mi> <mo>(</mo> <mi mathvariant="normal">E</mi> <mo>)</mo> </mrow> </semantics></math> (red-colored line) and transmission <math display="inline"><semantics> <mrow> <mi mathvariant="normal">T</mi> <mo>(</mo> <mi mathvariant="normal">E</mi> <mo>)</mo> </mrow> </semantics></math> (green-colored line) for the 0.9 μm thick (Al<sub>0.23</sub>Ga0.74)<sub>0.5</sub>In<sub>0.5</sub>P epifilm grown on the GaAs substrate. The magenta-colored vertical arrows drawn near 605 nm predicted the bandgap of ~2.05 eV, in excellent agreement with our PL and existing data from the literature [<a href="#B71-molecules-29-04188" class="html-bibr">71</a>].</p>
Full article ">
18 pages, 4453 KiB  
Article
Electrospun PVP Fibers as Carriers of Ca2+ Ions to Improve the Osteoinductivity of Titanium-Based Dental Implants
by Janina Roknić, Ines Despotović, Jozefina Katić and Željka Petrović
Molecules 2024, 29(17), 4181; https://doi.org/10.3390/molecules29174181 - 3 Sep 2024
Viewed by 654
Abstract
Although titanium and its alloys are widely used as dental implants, they cannot induce the formation of new bone around the implant, which is a basis for the functional integrity and long-term stability of implants. This study focused on the functionalization of the [...] Read more.
Although titanium and its alloys are widely used as dental implants, they cannot induce the formation of new bone around the implant, which is a basis for the functional integrity and long-term stability of implants. This study focused on the functionalization of the titanium/titanium oxide surface as the gold standard for dental implants, with electrospun composite fibers consisting of polyvinylpyrrolidone and Ca2+ ions. Polymer fibers as carriers of Ca2+ ions should gradually dissolve, releasing Ca2+ ions into the environment of the implant when it is immersed in a model electrolyte of artificial saliva. Scanning electron microscopy, energy dispersive X-ray spectroscopy and attenuated total reflectance Fourier transform infrared spectroscopy confirmed the successful formation of a porous network of composite fibers on the titanium/titanium oxide surface. The mechanism of the formation of the composite fibers was investigated in detail by quantum chemical calculations at the density functional theory level based on the simulation of possible molecular interactions between Ca2+ ions, polymer fibers and titanium substrate. During the 7-day immersion of the functionalized titanium in artificial saliva, the processes on the titanium/titanium oxide/composite fibers/artificial saliva interface were monitored by electrochemical impedance spectroscopy. It can be concluded from all the results that the composite fibers formed on titanium have application potential for the development of osteoinductive and thus more biocompatible dental implants. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The ATR-FTIR spectra of (<b>a</b>) CaCl<sub>2</sub> salt, PVP and (PVP+Ca<sup>2+</sup>) fibers; (<b>b</b>) freshly prepared Ti, Ti modified with thermally prepared oxide film (Ti/Ti oxide) and Ti modified with the composite fibers [Ti/Ti oxide/(PVP+Ca<sup>2+</sup>) fibers].</p>
Full article ">Figure 2
<p>The SEM images and corresponding EDS spectra of (<b>a</b>,<b>b</b>) (PVP+Ca<sup>2+</sup>) fibers; (<b>c</b>,<b>d</b>) Ti modified with thermally prepared oxide film (Ti/Ti oxide) and (<b>e</b>,<b>f</b>) Ti modified with the composite fibers [Ti/Ti oxide/(PVP+Ca<sup>2+</sup>) fibers].</p>
Full article ">Figure 3
<p>The optimized structures of (<b>a</b>) PVP-Ca-I, in which the PVP-Ca bond is established via a nitrogen atom; (<b>b</b>) PVP-Ca-II, in which the PVP-Ca bond is established via oxygen atoms of the carbonyl groups. The bond distances are given in Å. The bond energies are given in kcal mol<sup>−1</sup>. O—red, C—gray, N—blue, H—white, Ca—yellow-green.</p>
Full article ">Figure 4
<p>The optimized structures of (<b>a</b>) PVP-Ca-TiO<sub>2</sub>-I, in which the Ca<sup>2+</sup> is complexed in between PVP layer and TiO<sub>2</sub> surface, (<b>b</b>) PVP-Ca-TiO<sub>2</sub>-II, in which the Ca<sup>2+</sup> is bound on the top of PVP layer. The bond distances are given in Å. The bond energies are given in kcal mol<sup>−1</sup>. O—red, C—gray, N—blue, H—white, Ca—yellow-green.</p>
Full article ">Figure 5
<p>The EIS plots in the form of (<b>a</b>) magnitude vs. log <span class="html-italic">f,</span> (<b>b</b>) phase angle vs. log <span class="html-italic">f</span> recorded on the [Ti/Ti oxide/(PVP+Ca<sup>2+</sup>) fibers] sample at open circuit potential after 1 h, 2 days and 7 days of immersion in artificial saliva, pH = 6.8.</p>
Full article ">Figure 6
<p>The SEM images and corresponding EDS spectra of (<b>a</b>,<b>b</b>) freshly electrospun (PVP+Ca<sup>2+</sup>) fibers on Al foil; (<b>c</b>,<b>d</b>) the Al foil surface with residual fibers after 7 days of immersion in artificial saliva, pH = 6.8. Fe, visible in the spectrum (<b>d</b>), originates from Al foil.</p>
Full article ">Figure 7
<p>The photographs of (<b>a</b>) the freshly abraded and degreased surface of the Ti samples; (<b>b</b>) the thermally generated oxide film on the Ti (Ti/Ti oxide); and (<b>c</b>) the (PVP+Ca<sup>2+</sup>) fibers on the Ti/Ti oxide sample.</p>
Full article ">
25 pages, 3593 KiB  
Article
Simulations of Infrared Reflectivity and Transmission Phonon Spectra for Undoped and Doped GeC/Si (001)
by Devki N. Talwar and Jason T. Haraldsen
Nanomaterials 2024, 14(17), 1439; https://doi.org/10.3390/nano14171439 - 3 Sep 2024
Viewed by 440
Abstract
Exploring the phonon characteristics of novel group-IV binary XC (X = Si, Ge, Sn) carbides and their polymorphs has recently gained considerable scientific/technological interest as promising alternatives to Si for high-temperature, high-power, optoelectronic, gas-sensing, and photovoltaic applications. Historically, the effects of phonons on [...] Read more.
Exploring the phonon characteristics of novel group-IV binary XC (X = Si, Ge, Sn) carbides and their polymorphs has recently gained considerable scientific/technological interest as promising alternatives to Si for high-temperature, high-power, optoelectronic, gas-sensing, and photovoltaic applications. Historically, the effects of phonons on materials were considered to be a hindrance. However, modern research has confirmed that the coupling of phonons in solids initiates excitations, causing several impacts on their thermal, dielectric, and electronic properties. These studies have motivated many scientists to design low-dimensional heterostructures and investigate their lattice dynamical properties. Proper simulation/characterization of phonons in XC materials and ultrathin epilayers has been challenging. Achieving the high crystalline quality of heteroepitaxial multilayer films on different substrates with flat surfaces, intra-wafer, and wafer-to-wafer uniformity is not only inspiring but crucial for their use as functional components to boost the performance of different nano-optoelectronic devices. Despite many efforts in growing strained zinc-blende (zb) GeC/Si (001) epifilms, no IR measurements exist to monitor the effects of surface roughness on spectral interference fringes. Here, we emphasize the importance of infrared reflectivity Rω  and transmission Tω spectroscopy at near normal θi = 0 and oblique θi ≠ 0 incidence (Berreman effect) for comprehending the phonon characteristics of both undoped and doped GeC/Si (001) epilayers. Methodical simulations of Rω and Tω revealing atypical fringe contrasts in ultrathin GeC/Si are linked to the conducting transition layer and/or surface roughness. This research provided strong perspectives that the Berreman effect can complement Raman scattering spectroscopy for allowing the identification of longitudinal optical ωLO phonons, transverse optical ωTO phonons, and LO-phonon–plasmon coupled ωLPP+  modes, respectively. Full article
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Sketch of a three-phase ideal model (‘air/epifilm/substrate’) with dielectric functions 1 air <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (air), 2 <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>2</mn> </msub> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mrow> <mi>tf</mi> </mrow> </msub> </mrow> </semantics></math> (zb GeC thin film), and 3 <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>3</mn> </msub> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mi mathvariant="normal">s</mi> </msub> </mrow> </semantics></math> (Si substrate) for studying the reflectance/transmission spectra of thin zb GeC/Si (001) films grown on a substrate. The modified model with the dielectric functions 1 air <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>1</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (air), 2<math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="true">∼</mo> </mover> </mrow> <mn>2</mn> </msub> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mrow> <mi>tf</mi> </mrow> </msub> </mrow> </semantics></math> (thin film) transition layer 2′ <math display="inline"><semantics> <mrow> <mrow> <msup> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> <mo>′</mo> </msup> </mrow> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mrow> <mi>tl</mi> </mrow> </msub> </mrow> </semantics></math>, and 3 <math display="inline"><semantics> <mrow> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mn>3</mn> </msub> <mo>=</mo> <msub> <mrow> <mover> <mi mathvariant="sans-serif">ε</mi> <mo stretchy="false">~</mo> </mover> </mrow> <mi mathvariant="normal">s</mi> </msub> </mrow> </semantics></math> (substrate). Scattering factors χ and χ<sub>2</sub> due to roughness between GeC/air and GeC//TL surface (see Equations (8b,c)) are also included for studying the reflectivity and transmission spectra of thin films grown on a substrate [see text].</p>
Full article ">Figure 2
<p>Calculated reflectance spectra at near-normal incidence for semi-infinite n-type zb GeC. The blue and red lines reflect the spectra for undoped η = 0 and n-doped with η = 0.5 E+19 cm<sup>−3</sup>, respectively. The positions of <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>TO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>LO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>modes of zb GeC are also marked (see text).</p>
Full article ">Figure 3
<p>(<b>a</b>) Calculated infrared reflectance spectra at near-normal incidence θ<sub>i</sub> ≈ 0 for the GeC/Si (001) epilayers of different film thicknesses. The results include bulk zb GeC as well as 8 μm, 6 μm, 4 μm, 2 μm, 1 μm, 0.5 μm, and 0.05 μm thick films. (<b>b</b>) Reflectivity spectra of 4 μm thick GeC/Si (001) epifilm, with blue- and red-colored lines indicating undoped η = 0 and n-doped η = 0.5 E+19 cm<sup>−3</sup>, respectively. (<b>c</b>) Polarization-dependent reflectivity of 0.5 μm thick GeC/Si (001) epifilm at oblique incident angle θ<sub>i</sub> = 45°, where blue- and red-colored lines indicate s- and p-polarization spectra. The positions of <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>TO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>LO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>modes of GeC are also marked (see text).</p>
Full article ">Figure 4
<p>(<b>a</b>) Calculated infrared transmission spectra at near-normal incidence θ<sub>i</sub> ≈ 0 for the GeC/Si (001) epilayers of different film thicknesses. The results include 8 μm, 6 μm, 4 μm, 2 μm, 1 μm, 0.5 μm, and 0.05 μm thick films. (<b>b</b>) Transmission spectra of 4 μm thick GeC/Si (001) epifilm, with blue- and red-colored lines indicating undoped η = 0 and n-doped η = 0.5 E+19 cm<sup>−3</sup>, respectively. (<b>c</b>) Polarization-dependent transmission spectra of 0.5 μm thick GeC/Si (001) epifilm at oblique incident angle θ<sub>i</sub> = 45°, where blue- and red-colored lines indicate s- and p-polarization spectra. The positions of <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>TO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">ω</mi> <mrow> <mi>LO</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>modes of GeC are also marked (see text).</p>
Full article ">Figure 5
<p>Calculated LO-phonon–plasmon coupled <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">ω</mi> <mrow> <mrow> <mi>LPP</mi> <mo> </mo> </mrow> </mrow> <mo>±</mo> </msubsup> </mrow> </semantics></math>mode frequencies in n-type GeC as a function of free carrier concentration η. The values of ω<sub>LO</sub>, ω<sub>TO</sub> modes (dotted lines) of GeC are indicated by sky blue arrows. Variation in ω<sub>P</sub> (sky blue line) with η is also displayed (see text).</p>
Full article ">Figure 6
<p>(<b>a</b>) Calculated plasma frequency <span class="html-italic">ω</span><sub>P</sub> in cm<sup>−1</sup> versus charge carrier concentration <span class="html-italic">η</span> (cm<sup>−3</sup>) in n-type GeC. (<b>b</b>) Calculated low field mobility μ in (cm<sup>2</sup>/Vs) (left) and plasmon coupling coefficient γ in cm<sup>−1</sup> versus charge carrier concentration <span class="html-italic">η</span> (cm<sup>−3</sup>) in n-type GeC (see text).</p>
Full article ">Figure 7
<p>Calculated infrared spectrum of 5.0 μm thick n-type GeC/Si(100) epifilm: (<b>a</b>) Reflectivity spectra as a function of frequency (cm<sup>−1</sup>) for a fixed value of γ = 150 cm<sup>−1</sup> while changing ω<sub>P</sub> from 300, 500, 700, 900, and 1200 cm<sup>−1</sup>. (<b>b</b>) Same key as for (<b>a</b>) but for the transmission spectra of 5.0 μm thick n-type GeC/Si(100) epifilm. (<b>c</b>) Reflectivity spectra as a function of frequency (cm<sup>−1</sup>) for a fixed value of ω<sub>P</sub> = 1000 cm<sup>−1</sup> while changing γ from 100, 200, 300, 400, and 500 cm<sup>−1</sup>. (<b>d</b>) Same key as for (<b>c</b>) but for the transmission spectra of 5.0 μm thick n-type GeC/Si(100) epifilm (see text).</p>
Full article ">Figure 8
<p>(<b>a</b>) Calculated infrared reflectivity spectra at oblique incidence (θ<sub>i</sub> = 45°) for n-type GeC/Si (001) 1.0 μm thick film in the s- and p-polarization (different colors). The charge carrier concentration η increased from 6.2 E+18 cm<sup>−3</sup> → 1.1 E+19 cm<sup>−3</sup> → 1.7 E+19 cm<sup>−3</sup> → 2.5 E+19 cm<sup>−3</sup>, respectively. The calculated shifts of <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">ω</mi> <mrow> <mrow> <mi>LPP</mi> <mo> </mo> </mrow> </mrow> <mo>+</mo> </msubsup> </mrow> </semantics></math> modes in the p-polarization spectra of GeC/Si are shown by the magenta-colored vertical arrows (see text). (<b>b</b>) Same key as for (<b>a</b>) but for the simulated transmission spectra of 1.0 μm thick epifilm with different charge carrier concentrations η.</p>
Full article ">Figure 9
<p>(<b>a</b>) Calculated reflectance at near-normal incidence for a 4 μm thick GeC/Si (100) epifilm (<math display="inline"><semantics> <mi mathvariant="sans-serif">η</mi> </semantics></math>~1.01 E+17 cm<sup>−3</sup>) with different air/film surface roughnesses δ (≡0.05 μm, 0.10 μm, and 0.15 μm). (<b>b</b>) Same key as for (<b>a</b>) but for different film/substrate interface roughnesses δ<sub>2</sub> (≡0.10 μm, 0.15 μm, 0.20 μm, and 0.25 μm) (see text).</p>
Full article ">Figure 10
<p>(<b>a</b>) Calculated reflectance spectra at near normal incidence (θ<sub>i</sub> ≈ 0) for a 4 μm thick GeC/Si (100) epifilm (<math display="inline"><semantics> <mi mathvariant="sans-serif">η</mi> </semantics></math>~1.01 E+17 cm<sup>−3</sup>) for a fixed value of transition layer thickness<math display="inline"><semantics> <mrow> <msub> <mrow> <mrow> <mo> </mo> <mi mathvariant="normal">d</mi> </mrow> </mrow> <mn>2</mn> </msub> <mo stretchy="false">(</mo> <mo>≡</mo> <mo> </mo> </mrow> </semantics></math>0.05 μm) and varying air/film surface roughness δ (≡ 0.10 μm, 0.15 μm, 0.20 μm, and 0.25 μm). (<b>b</b>) Same key as for (<b>a</b>) with a fixed value of transition layer thickness<math display="inline"><semantics> <mrow> <msub> <mrow> <mrow> <mo> </mo> <mi mathvariant="normal">d</mi> </mrow> </mrow> <mn>2</mn> </msub> <mo stretchy="false">(</mo> <mo>≡</mo> <mo> </mo> </mrow> </semantics></math>0.05 μm) and varying film/substrate interface roughness δ<sub>2</sub> (≡ 0.10 μm, 0.15 μm, 0.20 μm, and 0.25 μm) (see text).</p>
Full article ">
16 pages, 2554 KiB  
Article
Ball-Milling Enhanced UV Protection Performance of Ca2Fe-Sulisobenzone Layered Double Hydroxide Organic Clay
by Márton Szabados, Rebeka Mészáros, Dorina Gabriella Dobó, Zoltán Kónya, Ákos Kukovecz and Pál Sipos
Nanomaterials 2024, 14(17), 1436; https://doi.org/10.3390/nano14171436 - 2 Sep 2024
Viewed by 535
Abstract
Using a co-precipitation technique, the anionic form of sulisobenzone (benzophenone-4) sunscreen ingredient was incorporated into the interlayer space of CaFe-hydrocalumite for the first time. Using detailed post-synthetic millings of the photoprotective nanocomposite obtained, we aimed to study the mechanochemical effects on complex, hybridized [...] Read more.
Using a co-precipitation technique, the anionic form of sulisobenzone (benzophenone-4) sunscreen ingredient was incorporated into the interlayer space of CaFe-hydrocalumite for the first time. Using detailed post-synthetic millings of the photoprotective nanocomposite obtained, we aimed to study the mechanochemical effects on complex, hybridized layered double hydroxides (LDHs). Various physicochemical properties of the ground and the intact LDHs were compared by powder X-ray diffractometry, N2 adsorption-desorption, UV–Vis diffuse reflectance, infrared and Raman spectroscopy, scanning electron microscopy and thermogravimetric measurements. The data showed significant structural and morphological deformations, surface and textural changes and multifarious thermal behavior. The most interesting development was the change in the optical properties of organic LDHs; the milling significantly improved the UV light blocking ability, especially around 320 nm. Spectroscopic results verified that this could be explained by a modification in interaction between the LDH layers and the sulisobenzone anions, through modulated π–π conjugation and light absorption of benzene rings. In addition to the vibrating mill often used in the laboratory, the photoprotection reinforcement can also be induced by the drum mill grinding system commonly applied in industry. Full article
(This article belongs to the Section Nanocomposite Materials)
Show Figures

Figure 1

Figure 1
<p>X-ray powder diffraction patterns (<b>A</b>) of disodium salt of SB, starting nitrate-containing Ca<sub>2</sub>Fe-LDH and solids prepared with various iron(III):sulisobenzone molar ratios and (<b>B</b>) unmilled and milled Ca<sub>2</sub>Fe-sulisobenzone LDH organic clays with increasing ball-to-powder mass ratio (crystal thicknesses and basal distances signed in nm and Å, respectively).</p>
Full article ">Figure 2
<p>SEM photos of intact (<b>top</b>) and milled (BPR: 600) (<b>bottom</b>) Ca<sub>2</sub>Fe-SB LDHs.</p>
Full article ">Figure 3
<p>UV–Vis diffuse reflectance spectroscopy analysis of starting nitrate-containing Ca<sub>2</sub>Fe-LDH, SB, disodium salt of SB and unmilled, milled SB-LDHs with increasing ball-to-powder mass ratios ((<b>A</b>,<b>C</b>)—absorption, (<b>B</b>)—reflection spectra). Particle size distribution of unmilled and milled SB-LDH nanocomposites (<b>D</b>).</p>
Full article ">Figure 4
<p>Infrared (<b>A</b>,<b>A1</b>) and Raman (<b>B</b>,<b>B1</b>) spectra of starting nitrate-containing Ca<sub>2</sub>Fe-LDH, disodium salt of SB and unmilled, milled Ca<sub>2</sub>Fe-sulisobenzone LDHs with growing BPR.</p>
Full article ">Figure 5
<p>Thermogravimetric, derivative thermogravimetric (DTG) and evolved gas analysis of the starting (<b>A</b>,<b>A1</b>) and milled (<b>B</b>,<b>B1</b>) organic clays.</p>
Full article ">Figure 6
<p>X-ray powder diffraction curves (<b>A</b>) and UV–Vis diffuse reflectance spectra (<b>B</b>) of the unmilled and milled Ca<sub>2</sub>Fe-sulisobenzone LDHs with increasing drum mill operation time.</p>
Full article ">
11 pages, 3700 KiB  
Article
Preparation of Bi@Ho3+:TiO2/Composite Fiber Photocatalytic Materials and Hydrogen Production via Visible Light Decomposition of Water
by Tieping Cao, Yue Gao, Wei Xia and Xuan Qi
Catalysts 2024, 14(9), 588; https://doi.org/10.3390/catal14090588 - 2 Sep 2024
Viewed by 438
Abstract
Using electrospun nanofibers doped with TiO2 and rare-earth ion Ho3+ as the matrix, and sodium gluconate as the reducing agent, Bi(NO3)3 was reduced using hydrothermal technology to produce Bi@Ho3+:TiO2 composite fiber materials. The materials’ [...] Read more.
Using electrospun nanofibers doped with TiO2 and rare-earth ion Ho3+ as the matrix, and sodium gluconate as the reducing agent, Bi(NO3)3 was reduced using hydrothermal technology to produce Bi@Ho3+:TiO2 composite fiber materials. The materials’ phase, morphology, and photoelectric properties were characterized using various analytical testing methods, including X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray photoelectron spectroscopy (XPS), ultraviolet-visible diffuse reflectance spectroscopy (UV-Vis DRS), and transient photocurrent (IP). During the hydrothermal process, it was confirmed that Bi3+ was reduced by sodium gluconate to form pure Bi nanoparticles, which combined with Ho3+:TiO2 nanofibers to form heterojunctions. By leveraging the surface plasmon resonance (SPR) effect of metallic Bi and the abundant energy level structure and 4f electron transition properties of rare-earth Ho3+, the TiO2 nanofibers underwent dual modification, effectively enhancing the photocatalytic activity and stability of TiO2. Under visible light irradiation, the rate of hydrogen production through water decomposition reached 43.6 μmol·g−1·h−1. Full article
(This article belongs to the Section Photocatalysis)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns of Ho<sup>3+</sup>:TiO<sub>2</sub> nanofibers and TiO<sub>2</sub> nanofibers. (<b>b</b>) XRD patterns of the composite samples BHT2, BHT4, BHT6, BHT8, and HT.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) TiO<sub>2</sub> nanofibers and (<b>b</b>) Ho<sup>3+</sup>:TiO<sub>2</sub> nanofibers; (<b>c</b>) composite fibers BHT2, (<b>d</b>) BHT4, (<b>e</b>) BHT6, and (<b>f</b>) BHT8.</p>
Full article ">Figure 3
<p>(<b>a</b>) TEM, (<b>b</b>) HRTEM, and (<b>c</b>) EDS diagram of composite fiber sample BHT6.</p>
Full article ">Figure 4
<p>(<b>a</b>) Full spectrum of composite fiber sample BHT6; (<b>b</b>) high-resolution XPS spectra of Bi 4f, (<b>c</b>) Ti 2p, and (<b>d</b>) O 1s.</p>
Full article ">Figure 5
<p>UV-vis DRS of TiO<sub>2</sub> nanofibers and their composite fibers.</p>
Full article ">Figure 6
<p>PL diagram of TiO<sub>2</sub> nanofibers and their composite fibers.</p>
Full article ">Figure 7
<p>Transient photocurrent response of TiO<sub>2</sub> nanofibers and their composite fibers.</p>
Full article ">Figure 8
<p>Hydrogen production rate of TiO<sub>2</sub> nanofibers and their composite fibers.</p>
Full article ">Figure 9
<p>Schematic diagram showing the energy band structure and electron–hole pair separation of composite fiber sample Bi@Ho<sup>3+</sup>:TiO<sub>2</sub>.</p>
Full article ">
Back to TopTop