[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

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

Search Results (4,104)

Search Parameters:
Keywords = optical band

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
11 pages, 1759 KiB  
Communication
All-Fiber Micro-Ring Resonator Based p-Si/n-ITO Heterojunction Electro-Optic Modulator
by Yihan Zhu, Ziqian Wang, Xing Chen, Honghai Zhu, Lizhuo Zhou, Yujie Zhou, Yi Liu, Yule Zhang, Xilin Tian, Shuo Sun, Jianqing Li, Ke Jiang, Han Zhang and Huide Wang
Materials 2025, 18(2), 307; https://doi.org/10.3390/ma18020307 (registering DOI) - 11 Jan 2025
Viewed by 230
Abstract
With the rapid advancement of information technology, the data demands in transmission rates, processing speed, and storage capacity have been increasing significantly. However, silicon electro-optic modulators, characterized by their weak electro-optic effect, struggle to balance modulation efficiency and bandwidth. To overcome this limitation, [...] Read more.
With the rapid advancement of information technology, the data demands in transmission rates, processing speed, and storage capacity have been increasing significantly. However, silicon electro-optic modulators, characterized by their weak electro-optic effect, struggle to balance modulation efficiency and bandwidth. To overcome this limitation, we propose an electro-optic modulator based on an all-fiber micro-ring resonator and a p-Si/n-ITO heterojunction, achieving high modulation efficiency and large bandwidth. ITO is introduced in this design, which exhibits an ε-near-zero (ENZ) effect in the communication band. The real and imaginary parts of the refractive index of ITO undergo significant changes in response to variations in carrier concentration induced by the reverse bias voltage, thereby enabling efficient electro-optic modulation. Additionally, the design of the all-fiber micro-ring eliminates coupling losses associated with spatial optical-waveguide coupling, thereby resolving the high insertion loss of silicon waveguide modulators and the challenges of integrating MZI modulation structures. The results demonstrate that this modulator can achieve significant phase shifts at low voltages, with a modulation efficiency of up to 3.08 nm/V and a bandwidth reaching 82.04 GHz, indicating its potential for high-speed optical chip applications. Full article
(This article belongs to the Special Issue Advances in Materials Science for Engineering Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Free electron concentration dependence of the real (ε<sub>real</sub>) and imaginary (ε<sub>imag</sub>) parts of the complex permittivity of ITO. (<b>b</b>) Complete simulation flow of the proposed p-Si /n-ITO junction electro-optic modulator based on all-fiber micro-ring resonator. (<b>c</b>) Voltage-dependent carrier concentration of ITO region in heterojunction devices with different thickness of Si layer.</p>
Full article ">Figure 2
<p>Schematic layout of the ring resonator-based modulator. The inset shows the cross-section of the ring.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>d</b>) The visualization data of the carrier concentration in the ITO region of the PN heterojunction at 0 V, 2 V, 3 V, and 5 V, respectively. (<b>e</b>) The relationship between the real and imaginary parts of the effective refractive index of the optical field in the micro-ring as a function of the externally applied voltage. (<b>f</b>) Schematic diagram of the energy band changes in the PN junction before and after the application of reverse bias.</p>
Full article ">Figure 4
<p>(<b>a</b>) Transmission spectrum of the microfiber knot resonator (MKR). (<b>b</b>) Transmission spectra of the microfiber knot resonator (MKR) (black solid line) and the MKR with the PN junction (red dashed line). (<b>c</b>) Transmission spectra of the all-fiber micro-ring modulator under different reverse bias voltages. (<b>d</b>) Scatter plot and fitting curve of the voltage-dependent resonance peak shift.</p>
Full article ">Figure 5
<p>(<b>a</b>) Voltage-dependent variation curve of the PN junction capacitance under a 10 GHz small-signal interference. (<b>b</b>) Voltage-dependent RC bandwidth limitation curve of the PN junction.</p>
Full article ">
19 pages, 7132 KiB  
Article
Green Synthesis of Sustainable and Cost-Effective TiO2-SiO2-Fe2O3 Heterojunction Nanocomposites for Rhodamine B Dye Degradation Under Sunlight
by Sara Oumenoune Tebbi, Abdeltif Amrane, Reguia Boudraa, Jean-Claude Bollinger, Stefano Salvestrini, Muhammad Imran Kanjal, Ammar Tiri, Lazhar Belkhiri, Maymounah N. Alharthi and Lotfi Mouni
Water 2025, 17(2), 168; https://doi.org/10.3390/w17020168 - 10 Jan 2025
Viewed by 303
Abstract
TiO2-SiO2-Fe2O3 heterojunction using the ceramic technique was used in this study to investigate its effectiveness as a photocatalyst for Rhodamine B (RhB) dye degradation. Structural, optical, and morphological characterizations of the synthesized materials were carried out [...] Read more.
TiO2-SiO2-Fe2O3 heterojunction using the ceramic technique was used in this study to investigate its effectiveness as a photocatalyst for Rhodamine B (RhB) dye degradation. Structural, optical, and morphological characterizations of the synthesized materials were carried out by X-ray diffraction (XRD), photoluminescence analysis (PL), scanning electron microscopy (SEM-EDS), and diffuse reflectance spectroscopy (DRS) to calculate the gap energy. In addition, a degradation rate of around 97% was obtained at a pH of 8, an initial RhB concentration of 10 mg·L−1, a TS-1F semiconductor dosage of 1 g·L−1, and a reaction time of 210 min. The ability of photocatalysis to degrade RhB at different ratios, pH, and with/without H2O2 in aqueous media was evaluated under UV light, visible light (250 W), and sunlight. When it comes to the degradation of RhB under visible light (250 W) and sunlight, respectively, the influence of the np junction showed promising results for the degradation of RhB. In contrast, there was no discernible photocatalytic activity under UV light, which proves that the absorbance switched from UV to visible, demonstrating the decrease in the band gap energy. Additionally, an analysis of the procedure’s cost-effectiveness and reusability through an economic study revealed that the synthesized material was interesting in terms of both cost and sustainability. Full article
(This article belongs to the Special Issue Water Treatment Using Nanomaterials and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>DRX spectra of Fe<sub>2</sub>O<sub>3</sub>, TS, and TS-1F composite.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) Fe<sub>2</sub>O<sub>3</sub>, (<b>b</b>) TiO<sub>2</sub>-SiO<sub>2</sub>, (<b>c</b>) TS-1F.</p>
Full article ">Figure 3
<p>SEM-EDS mapping of TS-10F, TS-5F, and TF-1F.</p>
Full article ">Figure 4
<p>FTIR spectra of Fe<sub>2</sub>O<sub>3</sub>; TiO<sub>2</sub>-SiO<sub>2</sub>; and TS-1F, TS-3F, TS-5F, and TS-10F.</p>
Full article ">Figure 5
<p>(<b>a</b>) UV–visible diffuse absorption spectra and (<b>b</b>) plots of (αhν)<sup>1/2</sup> versus hν for determination of gap energies.</p>
Full article ">Figure 6
<p>Photoluminescence spectra of different composites.</p>
Full article ">Figure 7
<p>The effect of RhB concentration on the photocatalytic activity of the TS-1F heterojunction under visible light (250 W) (<b>left</b>) and sunlight (<b>right</b>).</p>
Full article ">Figure 8
<p>Boxplots of C/C<sub>0</sub> and % inhibition for both sunlight and visible light.</p>
Full article ">Figure 8 Cont.
<p>Boxplots of C/C<sub>0</sub> and % inhibition for both sunlight and visible light.</p>
Full article ">Figure 9
<p>Effect of pH values on the degradation rate under sunlight, C<sub>0</sub> = 10 ppm, TS-1F = 1 g/L, and 210 min.</p>
Full article ">Figure 10
<p>Effect of H<sub>2</sub>O<sub>2</sub> on RhB removal under sunlight, pH 8, C<sub>0</sub> = 10 ppm, TS-1F = 1 g/L, and 210 min.</p>
Full article ">Figure 11
<p>Effect of TS-1F ratio on the degradation efficiency under sunlight, pH 8, C<sub>0</sub> = 10 ppm, and 210 min.</p>
Full article ">Figure 12
<p>Recovery performance and stability of TS-1F degradation of RhB under sunlight, pH 8, C<sub>0</sub> = 10 ppm, TS-1F = 1 g/L, and 210 min.</p>
Full article ">Figure 13
<p>Effect of scavenger on photodegradation of RB, where TS-1F under sunlight C<sub>0</sub> = 10 mg/L, TS-1F = 0.1 g/100 mL, pH = 8, and C<sub>scavenger</sub> = 5 mM after 150 min.</p>
Full article ">Figure 14
<p>The suggested mechanism of RhB degradation using TS-1F under sunlight.</p>
Full article ">
25 pages, 7245 KiB  
Article
Long-Term Evaluation of GCOM-C/SGLI Reflectance and Water Quality Products: Variability Among JAXA G-Portal and JASMES
by Salem Ibrahim Salem, Mitsuhiro Toratani, Hiroto Higa, SeungHyun Son, Eko Siswanto and Joji Ishizaka
Remote Sens. 2025, 17(2), 221; https://doi.org/10.3390/rs17020221 - 9 Jan 2025
Viewed by 286
Abstract
The Global Change Observation Mission-Climate (GCOM-C) satellite, launched in December 2017, is equipped with the Second-generation Global Imager (SGLI) sensor, featuring a moderate spatial resolution of 250 m and 19 spectral bands, including the unique 380 nm band. After six years in orbit, [...] Read more.
The Global Change Observation Mission-Climate (GCOM-C) satellite, launched in December 2017, is equipped with the Second-generation Global Imager (SGLI) sensor, featuring a moderate spatial resolution of 250 m and 19 spectral bands, including the unique 380 nm band. After six years in orbit, a comprehensive evaluation of SGLI products and their temporal consistency is needed. Remote sensing reflectance (Rrs) is the primary product for monitoring water quality, forming the basis for deriving key oceanic constituents such as chlorophyll-a (Chla) and total suspended matter (TSM). The Japan Aerospace Exploration Agency (JAXA) provides Rrs products through two platforms, G-Portal and JASMES, each employing different atmospheric correction methodologies and assumptions. This study aims to evaluate the SGLI full-resolution Rrs products from G-Portal and JASMES at regional scales (Japan and East Asia) and assess G-Portal Rrs products globally between January 2018 and December 2023. The evaluation employs in situ matchups from NASA’s Aerosol Robotic Network-Ocean Color (AERONET-OC) and cruise measurements. We also assess the retrieval accuracy of two water quality indices, Chla and TSM. The AERONET-OC data analysis reveals that JASMES systematically underestimates Rrs values at shorter wavelengths, particularly at 412 nm. While the Rrs accuracy at 412 nm is relatively low, G-Portal’s Rrs products perform better than JASMES at shorter wavelengths, showing lower errors and stronger correlations with AERONET-OC data. Both G-Portal and JASMES show lower agreement with AERONET-OC and cruise datasets at shorter wavelengths but demonstrate improved agreement at longer wavelengths (530 nm, 565 nm, and 670 nm). JASMES generates approximately 12% more matchup data points than G-Portal, likely due to G-Portal’s stricter atmospheric correction thresholds that exclude pixels with high reflectance. In situ measurements indicate that G-Portal provides better overall agreement, particularly at lower Rrs magnitudes and Chla concentrations below 5 mg/m3. This evaluation underscores the complexities and challenges of atmospheric correction, particularly in optically complex coastal waters (Case 2 waters), which may require tailored atmospheric correction methods different from the standard approach. The assessment of temporal consistency and seasonal variations in Rrs data shows that both platforms effectively capture interannual trends and maintain temporal stability, particularly from the 490 nm band onward, underscoring the potential of SGLI data for long-term monitoring of coastal and oceanic environments. Full article
Show Figures

Figure 1

Figure 1
<p>Global distribution of matchups between validation datasets and GCOM-C/SGLI. (<b>a</b>) Locations of 331 in situ <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> matchups (red cross), 753 in situ Chla and TSM matchups (green circle) of cruise measurements, and 22 sites for G-Portal and JASMES comparison (blue cross). (<b>b</b>) Distribution of 3704 AERONET-OC matchups, with red circles indicating locations and sizes proportional to the number of measurements at each site. The dashed box represents the footprint of the JASMES full-resolution <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> product over Japan.</p>
Full article ">Figure 2
<p>Schematic diagram of atmospheric correction processes for (<b>a</b>) SGLI G-Portal and (<b>b</b>) SGLI JASMES.</p>
Full article ">Figure 3
<p>Comparison of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> spectra for various matchup scenarios. The first row represents <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> spectra for matchups between (<b>a</b>) AERONET-OC and both (<b>b</b>) SGLI G-Portal and (<b>c</b>) SGLI JASMES. The second row shows <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> spectra for matchups between (<b>d</b>) cruise measurements and both (<b>e</b>) SGLI G-Portal and (<b>f</b>) SGLI JASMES. Grey lines represent individual matchups, with black lines indicating the mean <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> values across different wavelengths. The legend at the top identifies the mean <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> values at various AERONET-OC sites, as shown in panel (<b>a</b>).</p>
Full article ">Figure 4
<p>Scatterplots (<b>a</b>–<b>f</b>) comparing AERONET-OC <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> with SGLI <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> of G-Portal (x symbols) and JASMES (• symbols) over Japan at four AERONET-OC sites: Ieodo_Station, Socheongcho, Ariake_Tower, and Kemigawa_Offshore. N refers to the number of matchups, R<sup>2</sup> to the coefficient of determination, β to bias, δ to mean absolute difference, ∆ to root mean square difference, and σ to mean absolute percentage difference in percent (%). For each wavelength, the evaluation metrics for G-Portal are listed first, followed by those for JASMES in parentheses. The dashed grey line represents the 1:1 line.</p>
Full article ">Figure 5
<p>Scatterplots (<b>a</b>–<b>f</b>) comparing AERONET-OC <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> values with SGLI G-Portal <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> values at 29 AERONET-OC sites, covering z global scale. N refers to the number of matchups, R<sup>2</sup> to the coefficient of determination, β to bias, δ to mean absolute difference, ∆ to root mean square difference, and σ to mean absolute percentage difference in percent (%). The dashed grey line represents the 1:1 line.</p>
Full article ">Figure 6
<p>Time series plots comparing <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> values from AERONET-OC (green circles), SGLI G-Portal (blue crosses), and SGLI JASMES (orange fork) across three observation sites: Kemigawa_Offshore, Ariake_Tower, and Socheongcho. Each row corresponds to specific SGLI bands (412 nm, 443 nm, 490 nm, 530 nm, 565 nm, and 670 nm). Parenthetical numbers adjacent to each SGLI band on the Y-axes indicate the closest corresponding AERONET-OC band. For instance, 670 (667) on the <span class="html-italic">Y</span>-axis of the last row represents the SGLI and AERONET-OC bands 670 nm and 667 nm, respectively.</p>
Full article ">Figure 7
<p>Scatterplots (<b>a</b>–<b>g</b>) compare cruise <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> with SGLI <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> of G-Portal (x symbols) and JASMES (• symbols). The number of matchups (N), coefficient of determination (R<sup>2</sup>), bias (β), mean absolute difference (δ), root mean square difference (∆), mean absolute percentage difference (σ) in percent (%), and chlorophyll-a (Chla) are shown. For each wavelength, the metrics for G-Portal are listed first, followed by those for JASMES in parentheses. The dashed grey line represents the 1:1 line.</p>
Full article ">Figure 8
<p>Difference in <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> values (Δ<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>) between cruise measurement and SGLI of (<b>a</b>) G-Portal globally, (<b>b</b>) G-Portal over Japan, and (<b>c</b>) JASMES. N refers to the number of matchups. Each grey line represents individual observations, while the black line indicates the mean Δ<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>, and the vertical bars represent the standard deviation of Δ<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> at each wavelength.</p>
Full article ">Figure 9
<p>Scatterplots comparing in situ measurements and SGLI-derived values from G-Portal and JASMES for two products: (<b>a</b>) Chla and (<b>b</b>) TSM. The number of matchups (N), coefficient of determination (R<sup>2</sup>), bias (β), mean absolute difference (δ), root mean square difference (∆), and mean absolute percentage difference (σ) in percent (%) are shown. The metrics for G-Portal are listed first, followed by those for JASMES in parentheses. The dashed grey line represents the 1:1 line.</p>
Full article ">Figure 10
<p>Density scatterplots (<b>a</b>–<b>g</b>) illustrate SGLI G-Portal <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> versus SGLI JASMES <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> over 22 sites (blue cross, <a href="#remotesensing-17-00221-f001" class="html-fig">Figure 1</a>a). N refers to the number of matchups, while R<sup>2</sup> and S denote the coefficient of determination and the slope, respectively. The dashed grey line represents the 1:1 line.</p>
Full article ">Figure 11
<p>Density scatterplots comparing SGLI G-Portal and SGLI JASMES for (<b>a</b>) Chla and (<b>b</b>) TSM over 22 sites covering Case 1 and Case 2 waters (blue cross, <a href="#remotesensing-17-00221-f001" class="html-fig">Figure 1</a>a). N refers to the number of matchups, while R<sup>2</sup> and S denote the coefficient of determination and the slope, respectively. The dashed grey line represents the 1:1 line.</p>
Full article ">Figure A1
<p>Trend and seasonality components of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> for five key bands (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>_412, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>_443, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>_490, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>_530, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>_670) at the Socheongcho AERONET-OC site. The trend plots (left column) and seasonality plots (right column) illustrate the temporal behavior of remote sensing reflectance across G-Portal, JASMES, and AERONET-OC data.</p>
Full article ">
16 pages, 7205 KiB  
Article
Comprehensive Structural, Chemical, and Optical Characterization of Cu2ZnSnS4 Films on Kapton Using the Automated Successive Ionic Layer Adsorption and Reaction Method
by Perla J. Vázquez-González, Martha L. Paniagua-Chávez, Lizette A. Zebadua-Chavarria, Rafael Mota-Grajales, C. A. Meza-Avendaño, Enrique Campos-González, A. Escobosa-Echavarría, Yaoqiao Hu, Aldo E. Pérez-Ramos, Manuel-Matuz and Carlos A. Hernández-Gutiérrez
Nanomaterials 2025, 15(2), 85; https://doi.org/10.3390/nano15020085 - 8 Jan 2025
Viewed by 320
Abstract
This study provides a comprehensive structural, chemical, and optical characterization of CZTS thin films deposited on flexible Kapton substrates via the Successive Ionic Layer Adsorption and Reaction (SILAR) method. The investigation explored the effects of varying deposition cycles (40, 60, 70, and 80) [...] Read more.
This study provides a comprehensive structural, chemical, and optical characterization of CZTS thin films deposited on flexible Kapton substrates via the Successive Ionic Layer Adsorption and Reaction (SILAR) method. The investigation explored the effects of varying deposition cycles (40, 60, 70, and 80) and annealing treatments on the films. An X-ray diffraction (XRD) analysis demonstrated enhanced crystallinity and phase purity, particularly in films deposited with 70 cycles. These films exhibited a notable reduction in secondary phases in the as-deposited state, with further improvements observed after annealing at 400 °C and 450 °C in a sulfur atmosphere. A pole figure analysis indicates a decrease in texture disorder with annealing, suggesting improved crystalline orientation at higher temperatures. Field emission scanning electron microscopy (FE-SEM) showed enhancements in surface morphology, with increased grain size and uniformity post-annealing. Chemical uniformity was confirmed through Secondary Ion Mass Spectrometry (SIMS), Energy-Dispersive Spectroscopy (EDS), and X-ray Photoelectron Spectroscopy (XPS). XPS revealed the presence of CZTS phases alongside oxidized phases. Annealing effectively reduced secondary phases, such as ZnO, SnO2, CuO, and SO2, enhancing the CZTS phase. An optical analysis demonstrated that annealing at 200 °C in an air atmosphere reduced the band gap from 1.53 eV to 1.38 eV. In contrast, annealing at 400 °C and 450 °C in a sulfur atmosphere increased the band gap to 1.59 eV and 1.63 eV, respectively. The films exhibited p-type conductivity, as inferred from a valence band structure analysis. Density Functional Theory (DFT) calculations provided insights into the observed band gap variations, further substantiating the findings. Full article
Show Figures

Figure 1

Figure 1
<p>The chemical process of CZTS deposition by SILAR method, and the deposited CZTS samples over the Kapton as a function of the number of cycles.</p>
Full article ">Figure 2
<p>XRD patterns of (<b>a</b>) the as-deposited samples as a function of deposition cycles, and (<b>b</b>) the samples deposited with 70 cycles and those annealed at 200 °C, 400 °C, and 450 °C.</p>
Full article ">Figure 3
<p>Texture analysis of the studied samples: (<b>a</b>) 60 cycles as-deposited, (<b>b</b>) 60 cycles after annealing, (<b>c</b>) 70 cycles as-deposited, (<b>d</b>) 70 cycles after annealing at 200 °C, (<b>e</b>) 70 cycles after annealing at 450 °C, (<b>f</b>) 70 cycles after annealing at 400 °C.</p>
Full article ">Figure 4
<p>F-SEM analysis of the studied samples: (<b>a</b>) 60 cycles as-deposited, (<b>b</b>) 60 cycles after annealing, (<b>c</b>) 70 cycles as-deposited, and (<b>d</b>) 70 cycles after annealing at 200 °C.</p>
Full article ">Figure 5
<p>F-SEM cross-section analysis of the studied samples, (<b>a</b>) 60 cycles as-deposited, (<b>b</b>) 60 cycles after annealing, (<b>c</b>) 70 cycles as-deposited, and (<b>d</b>) 70 cycles after annealing at 200 °C.</p>
Full article ">Figure 6
<p>(<b>a</b>) SIMS depth profile of the CZTS deposited on Kapton, (<b>b</b>) EDS and XPS quantification results for the analyzed samples.</p>
Full article ">Figure 7
<p>XPS analysis of CZTS elements for as-deposited films and films annealed at 200 °C and 450 °C.</p>
Full article ">Figure 7 Cont.
<p>XPS analysis of CZTS elements for as-deposited films and films annealed at 200 °C and 450 °C.</p>
Full article ">Figure 8
<p>(<b>a</b>) Band gap energy measurements of the studied samples, and (<b>b</b>) band gap extraction graph using the Kubelka–Munk method for the studied samples.</p>
Full article ">Figure 9
<p>(<b>a</b>) Valence band maximum analysis for 70-cycle as-deposited and annealed samples, (<b>b</b>) band gap diagram of the as-deposited sample, (<b>c</b>) band gap diagram of the sample annealed at 200 °C, and (<b>d</b>) band gap diagram of the sample annealed at 450 °C.</p>
Full article ">Figure 10
<p>(<b>a</b>) Atomic structure of CZTS used for DFT modeling. Band structures of (<b>b</b>) pristine CZTS, (<b>c</b>) CZTS with a sulfur (S) vacancy, and (<b>d</b>) CZTS with a zinc (Zn) vacancy.</p>
Full article ">
12 pages, 2120 KiB  
Article
First Principles Study of Electronic and Optical Properties of Cadmium-Tin-Oxide
by Adel Bandar Alruqi
Inorganics 2025, 13(1), 14; https://doi.org/10.3390/inorganics13010014 - 7 Jan 2025
Viewed by 255
Abstract
Cadmium-tin-oxide (CTO), also referred to as cadmium stannate (Cd2SnO4), is known for its interesting electrical, electronic, and optical properties, making it useful in various applications such as in transparent conducting oxides for optoelectronic devices and also in photovoltaic applications. [...] Read more.
Cadmium-tin-oxide (CTO), also referred to as cadmium stannate (Cd2SnO4), is known for its interesting electrical, electronic, and optical properties, making it useful in various applications such as in transparent conducting oxides for optoelectronic devices and also in photovoltaic applications. While its properties have been investigated experimentally, there is not much record in the literature on the computational study of the electronic and optical properties of CTO. This study employed density functional theory to explore the two properties of CTO. The hybrid functionals were used to widen the band gap from 0.381 eV (for PBE) to 3.13 eV, which replicates the experimental values very well. The other properties obtained were a refractive index of 2.53, absorption coefficient of 1.43 × 104 cm−1, and dielectric constant of 6.401 eV. The optical energy loss of 0.00691 that was investigated for the first time in this work adds to the literature on the properties of CTO. However, the electrical properties of CTO, which also play a key role in the working of optoelectronic devices, need to be investigated. Full article
Show Figures

Figure 1

Figure 1
<p>The electronic properties of cadmium-tin-oxide showing (<b>a</b>) the band structure and (<b>b</b>) the density of states.</p>
Full article ">Figure 2
<p>Density of states of cadmium-tin-oxide: (<b>a</b>) Projected DOS, and (<b>b</b>) Effective DOS.</p>
Full article ">Figure 3
<p>The real part, imaginary part, and effective dielectric function.</p>
Full article ">Figure 4
<p>The calculated refractive index as a function of energy.</p>
Full article ">Figure 5
<p>The computed optical absorption coefficient.</p>
Full article ">Figure 6
<p>The calculated optical energy loss.</p>
Full article ">Figure 7
<p>A 3D structure of cadmium-tin-oxide.</p>
Full article ">
10 pages, 2772 KiB  
Article
Carbon Nanosphere-Based TiO2 Double Inverse Opals
by Dániel Attila Karajz, Kincső Virág Rottenbacher, Klára Hernádi and Imre Miklós Szilágyi
Molecules 2025, 30(2), 205; https://doi.org/10.3390/molecules30020205 - 7 Jan 2025
Viewed by 379
Abstract
Inverse opals (IOs) are intensively researched in the field of photocatalysis, since their optical properties can be fine-tuned by the initial nanosphere size and material. Another possible route for photonic crystal programming is to stack IOs with different pore sizes. Accordingly, single and [...] Read more.
Inverse opals (IOs) are intensively researched in the field of photocatalysis, since their optical properties can be fine-tuned by the initial nanosphere size and material. Another possible route for photonic crystal programming is to stack IOs with different pore sizes. Accordingly, single and double IOs were synthesized using vertical deposition and atomic layer deposition. In the case of the double IOs, the alternating use of the two preparation methods was successfully performed. Hydrothermally synthesized 326 and 458 nm carbon nanospheres were utilized to manufacture two different IOs; hence the name 326 nm and 458 nm IOs. Heat treatment removed the sacrificial template carbon nanospheres, and the as-deposited TiO2 crystallized upon annealing into nanocrystalline anatase form. Reflectance mode UV–visible spectroscopy showed that most IOs had photonic properties, i.e., a photonic band gap, and by the “slow” photon effect enhanced absorbance, except the 326 nm IO, even though it also had an increase in absorbance. The IOs were tested by photocatalytic degradation of Rhodamine 6-G under visible light. Photocatalytic experiments showed that the 458 nm IO was more active and the double IOs showed higher efficiency compared to monolayers, even if the less effective 326 nm IO was the top layer. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of the IO samples.</p>
Full article ">Figure 2
<p>XRD diffractograms of the inverse opal samples.</p>
Full article ">Figure 3
<p>UV–vis spectra of the inverse opals (IOs), absorbance (<b>left</b>) and reflectance (<b>right</b>).</p>
Full article ">Figure 4
<p>Photocatalysis experiments of the inverse opal (IO) samples, under visible light irradiation.</p>
Full article ">
24 pages, 2793 KiB  
Article
Dispersive Sweatt Model for Broadband Lens Design with Metasurfaces
by Weiyu Chen, Ko-Han Shih and C. Kyle Renshaw
Photonics 2025, 12(1), 43; https://doi.org/10.3390/photonics12010043 - 6 Jan 2025
Viewed by 288
Abstract
The Sweatt model has been extensively used to design optical systems containing diffractive optical elements (DOEs) because it captures the dispersive characteristics of DOEs. We introduce a new dispersive Sweatt model (DSM) that can describe meta-atom (MA) dispersion, which has material and geometric [...] Read more.
The Sweatt model has been extensively used to design optical systems containing diffractive optical elements (DOEs) because it captures the dispersive characteristics of DOEs. We introduce a new dispersive Sweatt model (DSM) that can describe meta-atom (MA) dispersion, which has material and geometric contributions in addition to diffraction. It uses a wavelength-dependent scalar coefficient to modify the diffractive dispersion and describe the dispersion of a given MA basis. This provides a robust framework to design systems containing metasurface (MS) elements while including their unique dispersive properties in the design optimization. Importantly, the DSM is based on ray optics and enables the design of MS-containing systems using conventional optical design software such as Zemax and Code V. We use the DSM to demonstrate the design of a hybrid refractive/MS achromatic doublet for the midwave infrared (MWIR) band. The design example includes multiple wavelengths and field angles during optimization and demonstrates excellent agreement between the DSM and real hybrid lens performance modeled using wave optics. We discuss the limits of the DSM and present a simple model to predict performance limits due to phase mismatch at Fresnel zone boundaries. Full article
(This article belongs to the Special Issue Advancements in Optical Metamaterials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic illustrating the ideal phase delay across a flat, singlet metalens to give a perfect spherical wavefront focusing at a distance <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>f</mi> </mrow> <mrow> <mi>d</mi> </mrow> </msub> </mrow> </semantics></math> from the lens, as illustrated by the blue dashed line. The red line illustrates a ray from the exit pupil to the focal point; it is perpendicular to the spherical wavefront at the crossing point. <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mfenced separators="|"> <mrow> <mi>r</mi> </mrow> </mfenced> <mo>=</mo> <mi>l</mi> </mrow> </semantics></math> is the additional phase required at location <math display="inline"><semantics> <mrow> <mi>r</mi> </mrow> </semantics></math>; this is in excess of the phase delay at <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. To enhance the detail at both ends, the middle section was removed as indicated by the broken line. (<b>b</b>) Schematic illustrating chromatic focal shift of a metalens. A shorter wavelength ray (shown by the green line) focuses at a longer distance from the exit pupil. <math display="inline"><semantics> <mrow> <mi>ξ</mi> </mrow> </semantics></math> is the transverse ray-pierce error at the focal plane, and <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> </mrow> </semantics></math> is the axial chromatic focal shift.</p>
Full article ">Figure 2
<p>Meta-atom geometry for nanopillars (<b>a</b>) and nanoholes (<b>d</b>). Meta-atom data for nanopillars are shown on the top (<b>a</b>–<b>c</b>), and nanoholes are shown on the bottom (<b>d</b>–<b>f</b>). Phase delay is shown in the middle (<b>b</b>,<b>e</b>), and transmission on the right (<b>c</b>,<b>f</b>). Results for five wavelengths (λ) are shown: 3 μm (green), 3.5 μm (red), 4 μm (blue), 4.5 μm (yellow), and 5 μm (purple). The dashed lines in (<b>b</b>) show the DSM phase used to approximate the meta-atom dispersion via a dispersion parameter <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>(</mo> <mi>λ</mi> <mo>)</mo> </mrow> </semantics></math>, as described in the text.</p>
Full article ">Figure 3
<p>Singlet metalens prescription for Si nanopillars (top) and Air-nanoholes (bottom). (<b>a</b>,<b>c</b>) Phase profile of the metalenses across the MWIR band. (<b>b</b>,<b>d</b>) Corresponding phase profile for a DSM model of the singlet.</p>
Full article ">Figure 4
<p>Phase profiles of MS singlets using Si nanopillars. The first column (<b>a</b>,<b>c</b>) shows the real (i.e., wrapped) phase of the meta-atoms, and the second column (<b>b</b>,<b>d</b>) shows the unwrapped phase. Solid curves show the phase for the design wavelength 4 μm, alongside off-design wavelengths 3.5 μm (<b>a</b>,<b>b</b>) and 4.5 μm (<b>c</b>,<b>d</b>). For comparison, ideal phase profiles are shown for the off-design wavelength at the same 10 mm focal length (dashed).</p>
Full article ">Figure 5
<p>Focal spot of pillar (<b>left</b>) and hole (<b>right</b>) metalens for off-design wavelengths at best focus.</p>
Full article ">Figure 6
<p>(<b>a</b>) The phase error model. (<b>b</b>,<b>c</b>) Focusing efficiency as a function of phase error for a singlet metalens with F-number (<b>b</b>) F/5 and (<b>c</b>) F/2. The blue line shows the focusing efficiency of the linear phase error model described in the text. The open (NH) and filled (NP) symbols overlay results for the metalens singlets described in <a href="#sec4-photonics-12-00043" class="html-sec">Section 4</a>. Different symbol colors represent different wavelengths: green (3 μm), red (3.5 μm), blue (4 μm), yellow (4.5 μm), and purple (5 μm).</p>
Full article ">Figure 7
<p>An achromatic design example for a hybrid (refractive/metasurface) doublet. Design layouts are shown in the first column alongside MTFs for each layout. Each row corresponds to a different optical design: (<b>a</b>–<b>c</b>) Initial planoconvex Si singlet. (<b>d</b>–<b>f</b>) Si/MS doublet designed using the DSM. (<b>g</b>–<b>i</b>) Meta-atom arrangement of the “real” MS layout and doublet performance using the LUA to describe the MS. For each design, the modulation transfer function (MTF) is computed using wave optics and then averaged over the tangential and sagittal axes. For reference, the diffraction-limited MTF is shown for a wavelength of 4 μm.</p>
Full article ">
11 pages, 7884 KiB  
Article
Tuning Electronic Structure and Optical Properties of Monolayered h-BN by Doping C, Cu and Al
by Qun Li, Tengchao Gao, Kuo Zhang, Xiangming Che and Guolong Ni
Molecules 2025, 30(1), 192; https://doi.org/10.3390/molecules30010192 - 6 Jan 2025
Viewed by 300
Abstract
As a graphene-like material, h-BN has stimulated great research interest recently due to its potential application for next-generation electronic devices. Herein, a systematic theoretical investigation of electronic structures and optical properties of C-doped and Cu-Al co-doped h-BN is carried out by the first-principles [...] Read more.
As a graphene-like material, h-BN has stimulated great research interest recently due to its potential application for next-generation electronic devices. Herein, a systematic theoretical investigation of electronic structures and optical properties of C-doped and Cu-Al co-doped h-BN is carried out by the first-principles calculations. Firstly, two different C-doped h-BN structures for the para-position and ortho-position are constructed. The results show that the C ortho-doped h-BN (BCN) structure with a band gap of 3.05 eV is relatively stable, which is selected as a substate to achieve the Cu-Al co-doped h-BN. Based on this, the effect of the concentration of C atom doping on the electronic and optical properties of Cu-Al co-doped BCxN (x = 0, 11.1% and 22.2%) is investigated. The results demonstrate that the band gap of Cu-Al co-doped BCxN decreases and the optical properties improve with the increase in C atom concentration. The band gap and static dielectric constant of Cu-Al co-doped BC0N, BC1N and BC2N are 0.98 eV, 0.87 eV and 0.23 eV and 2.34, 3.03 and 3.77, respectively. As for all Cu-Al co-doped BCxN systems, the adsorption peak is red-shifted, and the peak intensity obviously decreases compared to the undoped h-BN. Additionally, the Cu-Al co-doped BC2N exhibits the best response to visible light. This work will provide valuable guidance for designing and developing h-BN-based doping systems with good performance in the field of optical and photocatalysis. Full article
(This article belongs to the Special Issue Chemical Research on Photosensitive Materials)
Show Figures

Figure 1

Figure 1
<p>Structure of (<b>a</b>) C para-doped and (<b>b</b>) ortho-doped h-BN and (<b>c</b>) the corresponding formation energy.</p>
Full article ">Figure 2
<p>Band structure of (<b>a</b>) undoped h-BN and (<b>b</b>) C para-doped and (<b>c</b>) C ortho-doped h-BN. PDOS for (<b>d</b>) undoped h-BN and (<b>e</b>) C para-doped h-BN. (<b>f</b>) PDOS of p-orbital for B, N and C atoms.</p>
Full article ">Figure 3
<p>Structure of (<b>a</b>) Cu-Al co-doped BC<sub>0</sub>N, (<b>b</b>) Cu-Al co-doped BC<sub>1</sub>N and (<b>c</b>) Cu-Al co-doped BC<sub>2</sub>N.</p>
Full article ">Figure 4
<p>Band structure of (<b>a</b>) Cu-Al co-doped BC<sub>0</sub>N, (<b>b</b>) Cu-Al co-doped BC<sub>1</sub>N and (<b>c</b>) Cu-Al co-doped BC<sub>2</sub>N. PDOS for (<b>d</b>) C ortho-doped h-BN and (<b>e</b>) Cu-Al co-doped BC<sub>0</sub>N.</p>
Full article ">Figure 5
<p>(<b>a</b>) Real parts and (<b>b</b>) imaginary parts of complex dielectric function for h-BN and Cu-Al co-doped BC<sub>x</sub>N.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) Optical absorption of undoped h-BN and Cu-Al co-doped BC<sub>x</sub>N and (<b>c</b>,<b>d</b>) reflection of undoped h-BN and Cu-Al co-doped BC<sub>x</sub>N.</p>
Full article ">Figure 7
<p>Complex refractive index of undoped h-BN and Cu-Al co-doped BC<sub>x</sub>N (<b>a</b>,<b>b</b>) n, (<b>c</b>,<b>d</b>) k.</p>
Full article ">Figure 8
<p>(<b>a</b>) Structure of h-BN supercell, (<b>b</b>) cutoff energy and (<b>c</b>) K-point convergence test of h-BN.</p>
Full article ">
12 pages, 2636 KiB  
Article
MoTe2 Photodetector for Integrated Lithium Niobate Photonics
by Qiaonan Dong, Xinxing Sun, Lang Gao, Yong Zheng, Rongbo Wu and Ya Cheng
Nanomaterials 2025, 15(1), 72; https://doi.org/10.3390/nano15010072 - 5 Jan 2025
Viewed by 383
Abstract
The integration of a photodetector that converts optical signals into electrical signals is essential for scalable integrated lithium niobate photonics. Two-dimensional materials provide a potential high-efficiency on-chip detection capability. Here, we demonstrate an efficient on-chip photodetector based on a few layers of MoTe [...] Read more.
The integration of a photodetector that converts optical signals into electrical signals is essential for scalable integrated lithium niobate photonics. Two-dimensional materials provide a potential high-efficiency on-chip detection capability. Here, we demonstrate an efficient on-chip photodetector based on a few layers of MoTe2 on a thin film lithium niobate waveguide and integrate it with a microresonator operating in an optical telecommunication band. The lithium-niobate-on-insulator waveguides and micro-ring resonator are fabricated using the femtosecond laser photolithography-assisted chemical–mechanical etching method. The lithium niobate waveguide-integrated MoTe2 presents an absorption coefficient of 72% and a transmission loss of 0.27 dB µm−1 at 1550 nm. The on-chip photodetector exhibits a responsivity of 1 mA W−1 at a bias voltage of 20 V, a low dark current of 1.6 nA, and a photo–dark current ratio of 108 W−1. Due to effective waveguide coupling and interaction with MoTe2, the generated photocurrent is approximately 160 times higher than that of free-space light irradiation. Furthermore, we demonstrate a wavelength-selective photonic device by integrating the photodetector and micro-ring resonator with a quality factor of 104 on the same chip, suggesting potential applications in the field of on-chip spectrometers and biosensors. Full article
(This article belongs to the Section Nanofabrication and Nanomanufacturing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of the MoTe<sub>2</sub>-based on-chip photodetector. (<b>b</b>) Raman spectrum of 2H-MoTe<sub>2</sub> (20 layers) on the LNOI platform under 532 nm laser excitation. Insert schematic of the MoTe<sub>2</sub> structure: Mo (purple) and Te (yellow), the arrows indicate the direction of atom vibration. (<b>c</b>) AFM image of 2H-MoTe<sub>2</sub> covering the LNOI waveguide (WG). (<b>d</b>) The thickness of the 2H-MoTe<sub>2</sub> layer corresponds to the region of the solid line (red color) marked in (<b>c</b>).</p>
Full article ">Figure 2
<p>Simulation of the light field distribution in the LNOI waveguide without 2H-MoTe<sub>2</sub> (<b>a</b>) and with 2H-MoTe<sub>2</sub> (<b>b</b>). (<b>c</b>) The simulation of the electric field intensity |E<sup>2</sup>| in the coupling section for TE polarization. (<b>d</b>) Measured transmission loss of the waveguide without and with 2H-MoTe<sub>2</sub>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic band diagrams of the Au-2H-MoTe<sub>2</sub>-Au structure: (<b>top</b>) in thermal equilibrium; (<b>bottom</b>) under illumination and applied bias voltage. (<b>b</b>) Comparison of the I-V curves under dark and in-coupled light intensity with 190 μW. (<b>c</b>) I–V curves at varying light intensities. (<b>d</b>) Photocurrent as a function of light intensity within the waveguide at different bias voltages. (<b>e</b>) Responsivity and EQE as functions of bias voltage. (<b>f</b>) Impulse response curve of the photodetector.</p>
Full article ">Figure 4
<p>(<b>a</b>) I–V curves at a constant in-coupled light intensity of 300 μW with different wavelengths. (<b>b</b>) Comparison of the responsivity of the photodetector under in-coupled light via waveguide and spatial illumination. (<b>c</b>) Simulated absorption rate of 2H-MoTe<sub>2</sub> at varying thicknesses. (<b>d</b>) 2H-MoTe<sub>2</sub> thickness-dependent responsivity of on-chip photodetectors.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic diagram of the PD on a waveguide-coupled ring resonant cavity, the arrows indicate the direction of light propagation. (<b>b</b>) Optical micrograph of the micro-ring resonator (MRR) and the corresponding waveguide-integrated photodetector. (<b>c</b>) Output energy of the waveguide with and without 2H-MoTe<sub>2</sub>. (<b>d</b>) Comparison of transmission spectrum measured using commercial detectors and photocurrent measured using on-chip integrated PD, and Lorentz fitting curve.</p>
Full article ">
24 pages, 10077 KiB  
Article
Deep-Learning-Based Method for the Identification of Typical Crops Using Dual-Polarimetric Synthetic Aperture Radar and High-Resolution Optical Images
by Xiaoshuang Ma, Le Li and Yinglei Wu
Remote Sens. 2025, 17(1), 148; https://doi.org/10.3390/rs17010148 - 3 Jan 2025
Viewed by 391
Abstract
Timely monitoring of distribution and growth state of crops is crucial for agricultural management. Remote sensing (RS) techniques provide an effective tool to monitor crops. This study proposes a novel approach for the identification of typical crops, including rapeseed and wheat, using multisource [...] Read more.
Timely monitoring of distribution and growth state of crops is crucial for agricultural management. Remote sensing (RS) techniques provide an effective tool to monitor crops. This study proposes a novel approach for the identification of typical crops, including rapeseed and wheat, using multisource remote sensing data and deep learning technology. By adopting an improved DeepLabV3+ network architecture that integrates a feature-enhanced module and an attention module, multiple features from both optical data and synthetic aperture radar (SAR) data are fully mined to take into account the spectral reflectance traits and polarimetric scattering straits of crops. The proposal can effectively address the limitations of using a single data source, alleviating the misclassification problem brought by the spectral similarity of crops in certain bands. Experimental results demonstrate that the proposed crop identification DeepLabV3+ (CI-DeepLabV3+) method outperforms traditional classification methods and the original DeepLabV3+ network, with an overall accuracy and F1 score of 94.54% and 94.55%, respectively. Experimental results also support the conclusion that using multiple features from multi-source data can indeed improve the performance of the network. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the study area.</p>
Full article ">Figure 2
<p>Phenological calendar of the two crops.</p>
Full article ">Figure 3
<p>Sub-images of the study area (orange rectangle: forest, yellow rectangle: wheat, red rectangle: rapeseed). (<b>a</b>) GF-2 optical image. (<b>b</b>) Sentinel-1 PolSAR image.</p>
Full article ">Figure 4
<p>Pictures taken in the field. (<b>a</b>) Wheat; (<b>b</b>) rapeseed; and (<b>c</b>) vegetation.</p>
Full article ">Figure 5
<p>Spectral reflectance of different vegetation type.</p>
Full article ">Figure 6
<p>The scattering characteristics of different land cover types.</p>
Full article ">Figure 7
<p>Diagram of the CI-DeepLabV3+ network.</p>
Full article ">Figure 8
<p>Schematic diagram of the enhanced module structure.</p>
Full article ">Figure 9
<p>Schematic diagram of the structure of the attention module.</p>
Full article ">Figure 10
<p>(<b>a</b>) Classification result with Feature Set 1: utilizing solely PolSAR feature information; (<b>b</b>) classification result with Feature Set 2: utilizing solely optical feature information; and (<b>c</b>) classification result with Feature Set 3: concurrently leveraging both optical and PolSAR feature information.</p>
Full article ">Figure 11
<p>Classification results of the CI-DeepLabV3+ model with different feature sets (wheat is marked in red and rapeseed is marked in green): (<b>a</b>) GF-2 optical image (false color); (<b>b</b>) ground truth label image; (<b>c</b>) classification result with Feature Set 1; (<b>d</b>) classification result with Feature Set 2; and (<b>e</b>) classification result with Feature Set 3.</p>
Full article ">Figure 12
<p>Classification results of different methods. (<b>a</b>) GF-2 optical image (false color display); (<b>b</b>) ground truth label image; (<b>c</b>) classification result with SVM; (<b>d</b>) classification result with U-net; (<b>e</b>) classification result with DeepLabV3+; and (<b>f</b>) classification result with CI-DeepLabV3+.</p>
Full article ">Figure 13
<p>The results for different combinations of hyperparameters.</p>
Full article ">Figure 14
<p>Performance metrics for varying sample sizes. (<b>a</b>) F1 score for wheat and rapeseed; (<b>b</b>) IOU for wheat and rapeseed; and (<b>c</b>) OA for wheat and rapeseed.</p>
Full article ">Figure 14 Cont.
<p>Performance metrics for varying sample sizes. (<b>a</b>) F1 score for wheat and rapeseed; (<b>b</b>) IOU for wheat and rapeseed; and (<b>c</b>) OA for wheat and rapeseed.</p>
Full article ">
31 pages, 9251 KiB  
Article
Seasonal Land Use and Land Cover Mapping in South American Agricultural Watersheds Using Multisource Remote Sensing: The Case of Cuenca Laguna Merín, Uruguay
by Giancarlo Alciaturi, Shimon Wdowinski, María del Pilar García-Rodríguez and Virginia Fernández
Sensors 2025, 25(1), 228; https://doi.org/10.3390/s25010228 - 3 Jan 2025
Viewed by 385
Abstract
Recent advancements in Earth Observation sensors, improved accessibility to imagery and the development of corresponding processing tools have significantly empowered researchers to extract insights from Multisource Remote Sensing. This study aims to use these technologies for mapping summer and winter Land Use/Land Cover [...] Read more.
Recent advancements in Earth Observation sensors, improved accessibility to imagery and the development of corresponding processing tools have significantly empowered researchers to extract insights from Multisource Remote Sensing. This study aims to use these technologies for mapping summer and winter Land Use/Land Cover features in Cuenca de la Laguna Merín, Uruguay, while comparing the performance of Random Forests, Support Vector Machines, and Gradient-Boosting Tree classifiers. The materials include Sentinel-2, Sentinel-1 and Shuttle Radar Topography Mission imagery, Google Earth Engine, training and validation datasets and quoted classifiers. The methods involve creating a multisource database, conducting feature importance analysis, developing models, supervised classification and performing accuracy assessments. Results indicate a low significance of microwave inputs relative to optical features. Short-wave infrared bands and transformations such as the Normalised Vegetation Index, Land Surface Water Index and Enhanced Vegetation Index demonstrate the highest importance. Accuracy assessments indicate that performance in mapping various classes is optimal, particularly for rice paddies, which play a vital role in the country’s economy and highlight significant environmental concerns. However, challenges persist in reducing confusion between classes, particularly regarding natural vegetation features versus seasonally flooded vegetation, as well as post-agricultural fields/bare land and herbaceous areas. Random Forests and Gradient-Boosting Trees exhibited superior performance compared to Support Vector Machines. Future research should explore approaches such as Deep Learning and pixel-based and object-based classification integration to address the identified challenges. These initiatives should consider various data combinations, including additional indices and texture metrics derived from the Grey-Level Co-Occurrence Matrix. Full article
(This article belongs to the Special Issue Feature Papers in Remote Sensors 2024)
Show Figures

Figure 1

Figure 1
<p>Number of relevant Scopus-indexed studies using Sentinel-1 and Sentinel-2 for remote sensing (2016–2023).</p>
Full article ">Figure 2
<p>The study area.</p>
Full article ">Figure 3
<p>Harmonized Sentinel-2MultiSpectral Instrument—Level-2A composites.</p>
Full article ">Figure 4
<p>Harmonized Sentinel-2MultiSpectral Instrument-Level-2A indices.</p>
Full article ">Figure 5
<p>Sentinel-1 Ground Range Detected medium composites per polarisation and season.</p>
Full article ">Figure 6
<p>Elevation and slope derived from the shuttle radar topography mission.</p>
Full article ">Figure 7
<p>Representative features of each class.</p>
Full article ">Figure 8
<p>Simplified flow chart of the methodology.</p>
Full article ">Figure 9
<p>Feature importance according to the feature and the classifier.</p>
Full article ">Figure 10
<p>Maps according to the models.</p>
Full article ">Figure 11
<p>Potential applications of LULC cartography in air, water and soil quality assessments.</p>
Full article ">
22 pages, 2943 KiB  
Article
Characterization of 77 GHz Radar Backscattering from Sea Surfaces at Low Incidence Angles: Preliminary Results
by Qinghui Xu, Chen Zhao, Zezong Chen, Sitao Wu, Xiao Wang and Lingang Fan
Remote Sens. 2025, 17(1), 116; https://doi.org/10.3390/rs17010116 - 1 Jan 2025
Viewed by 425
Abstract
Millimeter-wave (MMW) radar is capable of providing high temporal–spatial measurements of the ocean surface. Some topics, such as the characterization of the radar echo, have attracted widespread attention from researchers. However, most existing research studies focus on the backscatter of the ocean surface [...] Read more.
Millimeter-wave (MMW) radar is capable of providing high temporal–spatial measurements of the ocean surface. Some topics, such as the characterization of the radar echo, have attracted widespread attention from researchers. However, most existing research studies focus on the backscatter of the ocean surface at low microwave bands, while the sea surface backscattering mechanism in the 77 GHz frequency band remains not well interpreted. To address this issue, in this paper, the investigation of the scattering mechanism is carried out for the 77 GHz frequency band ocean surface at small incidence angles. The backscattering coefficient is first simulated by applying the quasi-specular scattering model and the corrected scattering model of geometric optics (GO4), using two different ocean wave spectrum models (the Hwang spectrum and the Kudryavtsev spectrum). Then, the dependence of the sea surface normalized radar cross section (NRCS) on incidence angles, azimuth angles, and sea states are investigated. Finally, by comparison between model simulations and the radar-measured data, the 77 GHz frequency band scattering characterization of sea surfaces at the near-nadir incidence is verified. In addition, experimental results from the wave tank are shown, and the difference in the scattering mechanism is further discussed between water surfaces and oceans. The obtained results seem promising for a better understanding of the ocean surface backscattering mechanism in the MMW frequency band. It provides a new method for fostering the usage of radar technologies for real-time ocean observations. Full article
(This article belongs to the Topic Radar Signal and Data Processing with Applications)
Show Figures

Figure 1

Figure 1
<p>The geometry for MMW radar observations.</p>
Full article ">Figure 2
<p>Experimental set-up for sea surface observations. The device in the blue circle is the UAV with the MMW radar. The red circle is the site of the wave buoy.</p>
Full article ">Figure 3
<p>Experimental set-up in the wave tank. The radar was fixed on the stationary red bridge at a height of about <math display="inline"><semantics> <mrow> <mn>13.5</mn> </mrow> </semantics></math> m.</p>
Full article ">Figure 4
<p>The range spectrum of single-chirp signal for the sea surface observation.</p>
Full article ">Figure 5
<p>The attitude of the UAV platform.</p>
Full article ">Figure 6
<p>Experimental set-up for the external calibration. (<b>a</b>) The geometry for the external calibration. The UAV with the MMW radar hovered in the air, and the trihedral corner reflector was placed on the ground. (<b>b</b>) The measurement environment from the view of the camera in the UAV.</p>
Full article ">Figure 7
<p>The simulated NRCS from two theoretical electromagnetic scattering models versus incidence angle with four sea states (<math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math>) in the 77 GHz frequency band. (<b>a</b>) Results from the Gaussian QS (QS-G) model in the upwind direction. (<b>b</b>) Results from the Gaussian GO4 (GO4-G) model in the upwind direction. (<b>c</b>) Results from the QS model in the crosswind direction. (<b>d</b>) Results from the GO4 mode in the crosswind direction.</p>
Full article ">Figure 8
<p>The simulated NRCS from two scattering models versus incidence angle with four sea states (<math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math>) in the 77 GHz frequency band. (<b>a</b>) Results from the QS model in the upwind direction. (<b>b</b>) Results from the GO4 model in the upwind direction. (<b>c</b>) Results from the QS model in the crosswind direction. (<b>d</b>) Results from the GO4 mode in the crosswind direction.</p>
Full article ">Figure 9
<p>The NRCS versus the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> under three incidence angles (<math display="inline"><semantics> <msup> <mn>5</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>, and <math display="inline"><semantics> <msup> <mn>15</mn> <mo>∘</mo> </msup> </semantics></math> incidence angles) for two sea spectrum models. (<b>a</b>) Results from the H spectrum in upwind directions. (<b>b</b>) Results from the K spectrum in upwind directions. (<b>c</b>) Results from the H spectrum in crosswind directions. (<b>d</b>) Results from the K spectrum in crosswind directions.</p>
Full article ">Figure 10
<p>The error <math display="inline"><semantics> <mrow> <mi>e</mi> <mi>r</mi> <mi>r</mi> </mrow> </semantics></math> as a function of <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> using the K spectrum. Curves with circles and squares represent the results in upwind and crosswind directions.</p>
Full article ">Figure 11
<p>The simulated NRCS versus azimuth angles with two sea states in the 77 GHz frequency band by the H spectrum and K spectrum, at different incidence angles. (<b>a</b>–<b>c</b>) Results from the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>0.45</mn> </mrow> </semantics></math> m at <math display="inline"><semantics> <msup> <mn>5</mn> <mo>∘</mo> </msup> </semantics></math> incidence, <math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math> incidence, and <math display="inline"><semantics> <msup> <mn>15</mn> <mo>∘</mo> </msup> </semantics></math> incidence. (<b>d</b>–<b>f</b>) Results from the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>1.08</mn> </mrow> </semantics></math> m.</p>
Full article ">Figure 12
<p>The error <math display="inline"><semantics> <mrow> <mi>e</mi> <mi>r</mi> <mi>r</mi> </mrow> </semantics></math> as a function of azimuth angles using the H spectrum. The blue solid line, the green dotted line, and red dotted line represent the results obtained from the <math display="inline"><semantics> <msup> <mn>5</mn> <mo>∘</mo> </msup> </semantics></math>, <math display="inline"><semantics> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math>, and <math display="inline"><semantics> <msup> <mn>15</mn> <mo>∘</mo> </msup> </semantics></math> incidences, respectively.</p>
Full article ">Figure 13
<p>The obtained NRCS versus incidence angles with six sea states in upwind directions. (<b>a</b>) Ocean surfaces with the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>0.33</mn> </mrow> </semantics></math> m. (<b>b</b>) Ocean surfaces with the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>0.40</mn> </mrow> </semantics></math> m. (<b>c</b>) Ocean surfaces with the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>0.43</mn> </mrow> </semantics></math> m. (<b>d</b>) Ocean surfaces with the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>0.48</mn> </mrow> </semantics></math> m. (<b>e</b>) Ocean surfaces with the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>0.52</mn> </mrow> </semantics></math> m. (<b>f</b>) Ocean surfaces with the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>0.63</mn> </mrow> </semantics></math> m. Errorbars are <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>1</mn> </mrow> </semantics></math> standard deviation.</p>
Full article ">Figure 14
<p>Comparison of the NRCS (in dB) obtained from the upwind directions in near-nadir-radar sea surface observations. The <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> of six datasets are <math display="inline"><semantics> <mrow> <mn>0.33</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <mn>0.40</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <mn>0.43</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <mn>0.48</mn> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <mn>0.52</mn> </mrow> </semantics></math> m, and <math display="inline"><semantics> <mrow> <mn>0.63</mn> </mrow> </semantics></math> m, respectively. Errorbars are <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>1</mn> </mrow> </semantics></math> standard deviation.</p>
Full article ">Figure 15
<p>The obtained NRCS from the wave tank observation in the upwind direction. (<b>a</b>) The irregular wave water surfaces. (<b>b</b>) The wave tank experiments with regular waves.</p>
Full article ">Figure 16
<p>The relative NRCS as a function of <math display="inline"><semantics> <mi>θ</mi> </semantics></math> for the results obtained from water and ocean surfaces in the 77 GHz frequency band. The results from two observations are shown in the red triangle line and green rhombus line, respectively. Curves with circles represent the relative difference (in dB) between the results from water and sea surfaces.</p>
Full article ">Figure 17
<p>Wave profile from regular waves in the wave tank.</p>
Full article ">
17 pages, 9384 KiB  
Article
Multi-Spectral Point Cloud Constructed with Advanced UAV Technique for Anisotropic Reflectance Analysis of Maize Leaves
by Kaiyi Bi, Yifang Niu, Hao Yang, Zheng Niu, Yishuo Hao and Li Wang
Remote Sens. 2025, 17(1), 93; https://doi.org/10.3390/rs17010093 - 30 Dec 2024
Viewed by 305
Abstract
Reflectance anisotropy in remote sensing images can complicate the interpretation of spectral signature, and extracting precise structural information under these pixels is a promising approach. Low-altitude unmanned aerial vehicle (UAV) systems can capture high-resolution imagery even to centimeter-level detail, potentially simplifying the characterization [...] Read more.
Reflectance anisotropy in remote sensing images can complicate the interpretation of spectral signature, and extracting precise structural information under these pixels is a promising approach. Low-altitude unmanned aerial vehicle (UAV) systems can capture high-resolution imagery even to centimeter-level detail, potentially simplifying the characterization of leaf anisotropic reflectance. We proposed a novel maize point cloud generation method that combines an advanced UAV cross-circling oblique (CCO) photography route with the Structure from the Motion-Multi-View Stereo (SfM-MVS) algorithm. A multi-spectral point cloud was then generated by fusing multi-spectral imagery with the point cloud using a DSM-based approach. The Rahman–Pinty–Verstraete (RPV) model was finally applied to establish maize leaf-level anisotropic reflectance models. Our results indicated a high degree of similarity between measured and estimated maize structural parameters (R2 = 0.89 for leaf length and 0.96 for plant height) based on accurate point cloud data obtained from the CCO route. Most data points clustered around the principal plane due to a constant angle between the sun and view vectors, resulting in a limited range of view azimuths. Leaf reflectance anisotropy was characterized by the RPV model with R2 ranging from 0.38 to 0.75 for five wavelength bands. These findings hold significant promise for promoting the decoupling of plant structural information and leaf optical characteristics within remote sensing data. Full article
Show Figures

Figure 1

Figure 1
<p>The location and imagery of the study site. (<b>a</b>) The position of the National Experiment Station for Precision Agriculture in Beijing; (<b>b</b>) the RGB image of a single plot; (<b>c</b>) the digital surface model (DSM) of 25 target maize plots.</p>
Full article ">Figure 2
<p>The two UAV systems (DJI Phantom4 Multispectral UAV, DJI Phantom 4 RTK UAV) used in the study.</p>
Full article ">Figure 3
<p>(<b>a</b>) Top view of a CCO route consisting of four single circles with a 50% inter-circle overlap; (<b>b</b>) the orthomosaic image acquired on 30 June 2023 and actual CCO flight paths for 25 plots. The blue arrow signifies the drone’s waypoints, and the white digits show the waypoint number.</p>
Full article ">Figure 4
<p>The analysis framework of the study.</p>
Full article ">Figure 5
<p>The fusion of spectral imagery and point cloud to generate multi-spectral point cloud. The voxelated grid of point cloud matches the multi-spectral image pixel (1.12 cm). (<b>a</b>) CCO multi-view images. (<b>b</b>) 3-D point cloud. (<b>c</b>) Multi-spectral imagery. (<b>d</b>) Ground-filtered point cloud. (<b>e</b>) Multi-spectral imagery registration. (<b>f</b>) Point cloud voxelization. (<b>g</b>) Multi-spectral point cloud.</p>
Full article ">Figure 6
<p>Reconstructed accurate point cloud from CCO oblique photography. (<b>a</b>) The side of RGB point cloud; (<b>b</b>) the side view of point cloud rendered by height.</p>
Full article ">Figure 7
<p>The accuracy of structural parameters, namely, (<b>a</b>) leaf length and (<b>b</b>) plant height, between extracted values from CCO-based point cloud and manual measurements.</p>
Full article ">Figure 8
<p>The sun-view geometries of point cloud and their statistics values. (<b>a</b>–<b>c</b>) The top view of observation geometries (view-angle, sun-angle, view-azimuth) calculated from the accurate point cloud of a single plot; (<b>d</b>–<b>f</b>) the frequency statistics of the three parameters. The unit of view-angle, sun-angle and view-azimuth is degree (°).</p>
Full article ">Figure 9
<p>(<b>a</b>) Polar distribution of the view-angle <span class="html-italic">θ</span><sub>v</sub> and view-azimuth <span class="html-italic">θ</span><sub>φ</sub>; (<b>b</b>) polar distribution of the sun-angle <span class="html-italic">θ</span><sub>v</sub> and view-azimuth <span class="html-italic">θ</span><sub>φ</sub> of points within the upper leaf layer of a single plot.</p>
Full article ">Figure 10
<p>The fitting between measured and estimated reflectance obtained by the RPV model. (<b>a</b>) The parameters of physical-based leaf BRDF; (<b>b</b>) the band of 450 nm; (<b>c</b>) the band of 560 nm; (<b>d</b>) the band of 650 nm; (<b>e</b>) the band of 730 nm; (<b>f</b>) the band of 840 nm. The red line represents the 1:1 line.</p>
Full article ">Figure 11
<p>Distribution of (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>) <math display="inline"><semantics> <mrow> <mo>Θ</mo> </mrow> </semantics></math> (<b>c</b>) k of the RPV model for 25 target maize plots. The ‘+’ symbol indicates outliers.</p>
Full article ">
10 pages, 3849 KiB  
Communication
Tunable Single-Longitudinal-Mode Thulium–Holmium Co-Doped Fiber Laser with an Ultra-Narrow Linewidth by Utilizing a Triple-Ring Passive Sub-Ring Resonator
by Pengfei Wang, Fengping Yan, Qi Qin, Dandan Yang, Ting Feng, Peng Liu, Ting Li, Chenhao Yu, Xiangdong Wang, Hao Guo, Yuezhi Cai, Wenjie Ji and Youchao Jiang
Photonics 2025, 12(1), 19; https://doi.org/10.3390/photonics12010019 - 28 Dec 2024
Viewed by 417
Abstract
A low-cost, wavelength-tunable single-longitudinal-mode (SLM) thulium–holmium co-doped fiber laser (THDFL) in a 2 μm band with a simple structure is described in the present paper. To obtain a stable SLM and narrow laser linewidth, a five-coupler-based three-ring (FCTR) filter is utilized in the [...] Read more.
A low-cost, wavelength-tunable single-longitudinal-mode (SLM) thulium–holmium co-doped fiber laser (THDFL) in a 2 μm band with a simple structure is described in the present paper. To obtain a stable SLM and narrow laser linewidth, a five-coupler-based three-ring (FCTR) filter is utilized in the ring cavity of the fiber laser. Tunable SLM wavelength output from THDFLs with kHz linewidths can be achieved by designing the FCTR filter with an effective free-spectral range and a 3 dB bandwidth at the main resonant peak. The measurement results show that the laser is in the SLM lasing state, with a highly stabilized optical spectrum, a linewidth of approximately 9.45 kHz, an optical signal-to-noise ratio as high as 73.6 dB, and a relative intensity noise of less than −142.66 dB/Hz. Furthermore, the wavelength can be tuned in the range of 2.6 nm. The proposed fiber laser has a wide range of applications, including coherence optical communication, optical fiber sensing, and dense wavelength-division-multiplexing. Full article
(This article belongs to the Special Issue Advanced Fiber Laser Technology and Its Application)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Experimental configuration of THDFL. LD: laser diode; WDM: wavelength division multiplexer; THDF: thulium–holmium co-doped fiber; CIR: circulator; FBG: fiber Bragg grating; OC: optical coupler; (<b>b</b>) schematic diagram of the proposed FCTR filter; (<b>c</b>) transmission and reflection spectra of the FBG.</p>
Full article ">Figure 2
<p>Signal-flow graph representation of the sub-ring cavity.</p>
Full article ">Figure 3
<p>Simulated spectra of the proposed FCTR filter. The inset is a zoom-in of the main resonant peak of the FCTR filter.</p>
Full article ">Figure 4
<p>(<b>a</b>) Spectrum of the single-wavelength at 2048.39 nm; (<b>b</b>) fluctuations in wavelength and power at 2048.39 nm.</p>
Full article ">Figure 5
<p>(<b>a</b>) The self-homodyne RF spectrum measured using a signal analyzer with a range of 0–100 MHz; (<b>b</b>) 0–500 MHz; and (<b>c</b>) 0–1000 MHz; (<b>d</b>) the spectrum of the main cavity without an FCTR filter.</p>
Full article ">Figure 6
<p>RIN spectra of the proposed SLM THDFL, in 0–5 MHz, using a RBW of 10 kHz for the signal analyzer. Insets show the same measurements in the 0–200 kHz range using a RBW of 100 Hz with relaxation oscillation peaks.</p>
Full article ">Figure 7
<p>Frequency noise PSD of the constructed SLM THDFL, and the linewidths at different integration times.</p>
Full article ">Figure 8
<p>The spectrum of the THDFL with a tunable wavelength range of ~2.6 nm.</p>
Full article ">
24 pages, 6956 KiB  
Article
Tailoring the Preparation, Microstructure, FTIR, Optical Properties and Photocatalysis of (Fe/Co) Co-Doped ZnO Nanoparticles (Zn0.9FexCo0.1−xO)
by A. M. Faramawy, W. R. Agami and Mohamed A. Swillam
Ceramics 2025, 8(1), 2; https://doi.org/10.3390/ceramics8010002 - 28 Dec 2024
Viewed by 352
Abstract
In this work, Fe3+- and Co2+-doped ZnO NPs (zinc oxide nanoparticles), Zn0.9FexCo0.1−xO, with a hexagonal wurtzite phase (single-phase), were synthesized via a co-precipitation technique where the phase purity and elemental composition were confirmed [...] Read more.
In this work, Fe3+- and Co2+-doped ZnO NPs (zinc oxide nanoparticles), Zn0.9FexCo0.1−xO, with a hexagonal wurtzite phase (single-phase), were synthesized via a co-precipitation technique where the phase purity and elemental composition were confirmed by XRD and EDX, respectively. Due to the substitution of Fe by Co, the cell parameters (a and c) were increased, alongside which a slight shift to higher diffracted angles appeared. FTIR was carried out to confirm the insertion of both the Fe3+ and Co2+ dopants into the ZnO hexagonal phase. Based on the experimental results, different numerical techniques were used to determine the optical gap and refractive index for the ZnO NP-doped samples, and when the concentration of Fe3+ ions was increased, the band gap value of ZnO decreased from 3.36 eV to 3.29 eV, accompanied by a decrease in the Urbach energy, while the refractive index increased. The doped ZnO NPs were later found to be effective UV photocatalysts which demonstrated a maximum reduction (84%) of methylene blue (MB) in a neutral environment for X = 0.05. The correlation between the Fe3+ concentration, structure, optical parameters, and photocatalytic efficacy is explained in detail. Full article
Show Figures

Figure 1

Figure 1
<p>An illustration of co-precipitation technique for Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O synthesis and the photodegradation steps of MB.</p>
Full article ">Figure 2
<p>(<b>a</b>) X-ray powder diffraction for Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles samples. (<b>b</b>) Shift in position of the (100), (002), and (101) diffraction peaks with increasing Fe<sup>3+</sup> substitution content.</p>
Full article ">Figure 3
<p>Variation in the texture coefficient (<span class="html-italic">TC</span>) of Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles.</p>
Full article ">Figure 4
<p>SEM and EDX spectra of Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles.</p>
Full article ">Figure 5
<p>FTIR spectra of Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles for different Fe concentrations.</p>
Full article ">Figure 6
<p>The plots of the derivative of <span class="html-italic">dF</span>(<span class="html-italic">R</span>)/<span class="html-italic">dλ</span> with respect to photon energy (<span class="html-italic">hν</span>) for the Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles.</p>
Full article ">Figure 7
<p>LD of (<span class="html-italic">αhν</span>) as a function of photon energy <span class="html-italic">hν</span> for Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles.</p>
Full article ">Figure 8
<p>The refractive index obtained from different numerical methods varies with Fe<sup>3+</sup> concentration.</p>
Full article ">Figure 9
<p>ln (<span class="html-italic">α</span>) vs. (<span class="html-italic">hν</span>) to determine the Urbach energy (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>u</mi> </mrow> </msub> </mrow> </semantics></math>) of Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles.</p>
Full article ">Figure 10
<p>UV–Vis spectra of MB as a function of time in presence of Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles under UV light irradiation.</p>
Full article ">Figure 11
<p>An illustration of charge carriers (electrons and holes) and mechanism of MB photodegradation, and images of MB discoloration in the presence of the Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O catalyst for sample X = 0.05.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>d</b>): The plots of (<b>a</b>) <span class="html-italic">C<sub>t</sub></span>/<span class="html-italic">C<sub>o</sub></span>, (<b>b</b>) <span class="html-italic">C<sub>t</sub></span>, (<b>c</b>) ln(<span class="html-italic">C<sub>o</sub></span>/<span class="html-italic">C<sub>t</sub></span>), and (<b>d</b>) (1/<span class="html-italic">C<sub>t</sub></span>) versus time along with their linear fits for MB dye with different Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles as photocatalysts with methylene blue at different Fe<sup>3+</sup> concentrations.</p>
Full article ">Figure 13
<p>Bar diagrams and data points for the influence of Fe<sup>3+</sup> concentrations on the photocatalytic degradation of Zn<sub>0.9</sub>Fe<sub>x</sub>Co<sub>0.1−x</sub>O nanoparticles. (The dotted lines are guides for the eye).</p>
Full article ">
Back to TopTop