Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

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

Search Results (7,835)

Search Parameters:
Keywords = optical sensors

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
36 pages, 5060 KiB  
Review
Two-Dimensional MoS2-Based Photodetectors
by Leilei Ye, Xiaorong Gan and Romana Schirhagl
Sustainability 2024, 16(22), 10137; https://doi.org/10.3390/su162210137 - 20 Nov 2024
Abstract
Nanomaterials can significantly improve the analytical performance of optical sensors for environmental pollutants. Two-dimensional (2D) molybdenum sulfide (MoS2) exhibits some unique physicochemical properties, such as strong light–matter interactions, bandgap tunability, and high carrier mobility, which are beneficial for constructing flexible optoelectronic [...] Read more.
Nanomaterials can significantly improve the analytical performance of optical sensors for environmental pollutants. Two-dimensional (2D) molybdenum sulfide (MoS2) exhibits some unique physicochemical properties, such as strong light–matter interactions, bandgap tunability, and high carrier mobility, which are beneficial for constructing flexible optoelectronic devices. In this review, the principle and classification of 2D MoS2-based photodetectors (PDs) are introduced, followed by a discussion about the physicochemical properties of 2D MoS2, as well as the structure–property relationships of 2D MoS2-based photoactive materials for PDs to understand the modulation strategies for enhancing the photodetection performance. Furthermore, we discuss significant advances in the surface modification and functionalization of 2D MoS2 for developing high-performance PDs, particularly focusing on synthesis pathways, modification strategies, and underlying physiochemical mechanisms for enhanced photodetection capabilities. Finally, conclusions and research perspectives on resolving significant bottlenecks or remaining challenges are offered based on recent developments in 2D MoS2-based PDs. Full article
20 pages, 7387 KiB  
Article
Skin Phototype Classification with Machine Learning Based on Broadband Optical Measurements
by Xun Yu, Keat Ghee Ong and Michael Aaron McGeehan
Sensors 2024, 24(22), 7397; https://doi.org/10.3390/s24227397 - 20 Nov 2024
Abstract
The Fitzpatrick Skin Phototype Classification (FSPC) scale is widely used to categorize skin types but has limitations such as the underrepresentation of darker skin phototypes, low classification resolution, and subjectivity. These limitations may contribute to dermatological care disparities in patients with darker skin [...] Read more.
The Fitzpatrick Skin Phototype Classification (FSPC) scale is widely used to categorize skin types but has limitations such as the underrepresentation of darker skin phototypes, low classification resolution, and subjectivity. These limitations may contribute to dermatological care disparities in patients with darker skin phototypes, including the misdiagnosis of wound healing progression and escalated dermatological disease severity. This study introduces (1) an optical sensor measuring reflected light across 410–940 nm, (2) an unsupervised K-means algorithm for skin phototype classification using broadband optical data, and (3) methods to optimize classification across the Near-ultraviolet-A, Visible, and Near-infrared spectra. The differentiation capability of the algorithm was compared to human assessment based on FSPC in a diverse participant population (n = 30) spanning an even distribution of the full FSPC scale. The FSPC assessment distinguished between light and dark skin phototypes (e.g., FSPC I vs. VI) at 560, 585, and 645 nm but struggled with more similar phototypes (e.g., I vs. II). The K-means algorithm demonstrated stronger differentiation across a broader range of wavelengths, resulting in better classification resolution and supporting its use as a quantifiable and reproducible method for skin type classification. We also demonstrate the optimization of this method for specific bandwidths of interest and their associated clinical implications. Full article
(This article belongs to the Special Issue Novel Optical Sensors for Biomedical Applications—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Fitzpatrick Skin Type Scale (I–VI) and (<b>b</b>) Generalized penetration depths of various wavelengths of light through tissue structures of interest [<a href="#B11-sensors-24-07397" class="html-bibr">11</a>].</p>
Full article ">Figure 2
<p>Block diagram of experimental procedures.</p>
Full article ">Figure 3
<p>Sensor outside of packaging showing electronics, LEDs, and photodiodes.</p>
Full article ">Figure 4
<p>K-means classification workflow diagram.</p>
Full article ">Figure 5
<p>Normalized intensity of (<b>a</b>) human evaluation skin classification method vs. (<b>b</b>) K-means<sub>410–940</sub> across a broad spectrum bandwidth; Significant main effects (α = 0.05) of the group on irradiance intensity are reported. NS: no statistical difference, *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001. All group-level statistical values across different wavelengths can be found in <a href="#app1-sensors-24-07397" class="html-app">Appendix A</a>, <a href="#sensors-24-07397-f0A1" class="html-fig">Figure A1</a> and <a href="#sensors-24-07397-f0A2" class="html-fig">Figure A2</a>.</p>
Full article ">Figure 6
<p>Normalized intensity of optimized (<b>a</b>) K-means<sub>410–535</sub>, (<b>b</b>) K-means<sub>560–705</sub>, and (<b>c</b>) K-means<sub>730–940</sub> across a 410–940 nm bandwidth. Green shading denotes optimized bandwidths in the K-means classification approach, whereas grey shading denotes neglected bandwidths. Significant main effects (α = 0.05) of the group on irradiance intensity are reported. NS: no statistical differences, *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001. All group-level statistical values across different wavelengths can be found in <a href="#app1-sensors-24-07397" class="html-app">Appendix A</a>, <a href="#sensors-24-07397-f0A3" class="html-fig">Figure A3</a>, <a href="#sensors-24-07397-f0A4" class="html-fig">Figure A4</a> and <a href="#sensors-24-07397-f0A5" class="html-fig">Figure A5</a>.</p>
Full article ">Figure 7
<p>Table of statistical differences for intra-grouping pairwise comparison results at various wavelengths under different classification methods. (<b>a</b>) human FSPC classification method, (<b>b</b>) K-means<sub>410–940</sub>, (<b>c</b>) K-means<sub>410–535</sub>, (<b>d</b>) K-means<sub>560–705</sub>, (<b>e</b>) K-means<sub>730–940</sub>. Colors in (<b>a</b>) represent Fitzpatrick skin type scales I–VI. NS: no statistical difference, *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001. All statistical tests excluded single-participant groupings. All intra-group level statistical values across different wavelengths can be found in <a href="#app1-sensors-24-07397" class="html-app">Appendix A</a>, <a href="#sensors-24-07397-f0A1" class="html-fig">Figure A1</a>, <a href="#sensors-24-07397-f0A2" class="html-fig">Figure A2</a>, <a href="#sensors-24-07397-f0A3" class="html-fig">Figure A3</a>, <a href="#sensors-24-07397-f0A4" class="html-fig">Figure A4</a> and <a href="#sensors-24-07397-f0A5" class="html-fig">Figure A5</a>.</p>
Full article ">Figure A1
<p>Groupwise (leftmost column) and pairwise (columns 2–8) comparison analysis results with <span class="html-italic">p</span>-value listed of human evaluation grouping results. (K–W: Kruskal–Wallis test). Red font denotes statistical significance.</p>
Full article ">Figure A2
<p>Groupwise (leftmost column) and pairwise (columns 2–8) comparison analysis results with <span class="html-italic">p</span>-value listed of K-means<sub>410–940</sub> grouping results. (K–W: Kruskal–Wallis test). Red font denotes statistical significance.</p>
Full article ">Figure A3
<p>Groupwise (leftmost column) and pairwise (columns 2–8) comparison analysis results with <span class="html-italic">p</span>-value listed of K-means<sub>410–535</sub> grouping results. (K–W: Kruskal–Wallis test). Red font denotes statistical significance.</p>
Full article ">Figure A4
<p>Groupwise (leftmost column) and pairwise (columns 2–8) comparison analysis results with <span class="html-italic">p</span>-value listed of K-means<sub>560–705</sub> grouping results. (K–W: Kruskal–Wallis test). Red font denotes statistical significance.</p>
Full article ">Figure A5
<p>Groupwise (leftmost column) and pairwise (columns 2–8) comparison analysis results with <span class="html-italic">p</span>-value listed of K-means<sub>730–940</sub> grouping results. (K–W: Kruskal–Wallis test). Red font denotes statistical significance.</p>
Full article ">
45 pages, 14894 KiB  
Review
Advances and Challenges in WO3 Nanostructures’ Synthesis
by Odeilson Morais Pinto, Rosimara Passos Toledo, Herick Ematne da Silva Barros, Rosana Alves Gonçalves, Ronaldo Spezia Nunes, Nirav Joshi and Olivia Maria Berengue
Processes 2024, 12(11), 2605; https://doi.org/10.3390/pr12112605 - 20 Nov 2024
Viewed by 65
Abstract
In recent decades, nanoscience has experienced rapid global advancements due to its focus on materials and compounds at the nanoscale with high efficiency across diverse applications. WO3 nanostructures have proven to be a unique material in the development of new technologies due [...] Read more.
In recent decades, nanoscience has experienced rapid global advancements due to its focus on materials and compounds at the nanoscale with high efficiency across diverse applications. WO3 nanostructures have proven to be a unique material in the development of new technologies due to their electrical, optical, and catalytic properties. They have been used as raw materials for the fabrication of electrochromic devices, optoelectronic devices, hydrogenation catalysts, gas sensors, adsorbents, lithium-ion batteries, solar driven-catalysts, and phototherapy. One of the most striking characteristics of WO3 is its morphological diversity, spanning from 0D to 2D, encompassing nanoparticles, nanowires, nanofibers, nanorods, nanosheets, and nanoplates. This review discusses common synthesis methods for WO3 nanostructures, including hydrothermal and solvothermal methods, microwave-assisted synthesis, sol-gel, electrospinning, co-precipitation, and solution combustion, with emphasis on the advantages and challenges of each of them. The processes involved, the obtained morphologies, and the resulting applications are also presented. As evidenced here, the fine control of the synthesis parameters allows the production of nanostructures with controlled phase, morphology, and size, essential aspects for the production of high-performance WO3-based devices. Full article
(This article belongs to the Section Materials Processes)
Show Figures

Figure 1

Figure 1
<p>WO<sub>3</sub> structures with (<b>a</b>) nanoplates [<a href="#B40-processes-12-02605" class="html-bibr">40</a>], (<b>b</b>) nanoflowers [<a href="#B41-processes-12-02605" class="html-bibr">41</a>], (<b>c</b>) nanosheets [<a href="#B42-processes-12-02605" class="html-bibr">42</a>], (<b>d</b>) polyhedral NPs [<a href="#B43-processes-12-02605" class="html-bibr">43</a>], (<b>e</b>) quasi-spherical NPs [<a href="#B44-processes-12-02605" class="html-bibr">44</a>], and (<b>f</b>) nanorods [<a href="#B45-processes-12-02605" class="html-bibr">45</a>] morphology obtained from the co-precipitation method. Figure adapted with permission from References [<a href="#B40-processes-12-02605" class="html-bibr">40</a>,<a href="#B41-processes-12-02605" class="html-bibr">41</a>,<a href="#B42-processes-12-02605" class="html-bibr">42</a>,<a href="#B43-processes-12-02605" class="html-bibr">43</a>,<a href="#B44-processes-12-02605" class="html-bibr">44</a>,<a href="#B45-processes-12-02605" class="html-bibr">45</a>]. Copyright Elsevier.</p>
Full article ">Figure 2
<p>(<b>a</b>) TEM, (<b>b</b>) HRTEM, and (<b>c</b>) adsorption scheme for MB adsorption of the WO<sub>3</sub> nanoadsorbent. Figure adapted with permission from Reference [<a href="#B47-processes-12-02605" class="html-bibr">47</a>]. Copyright 2017 Elsevier.</p>
Full article ">Figure 3
<p>TEM and HR-TEM images of (<b>a</b>,<b>b</b>) WO<sub>3</sub> nanorods and (<b>c</b>,<b>d</b>) 2 wt% Pd-loaded WO<sub>3</sub> nanorods and their corresponding SAED patterns in the inserts. Figure printed with permission from Reference [<a href="#B57-processes-12-02605" class="html-bibr">57</a>]. Copyright 2016 Elsevier.</p>
Full article ">Figure 4
<p>(<b>a</b>) TEM image of the hierarchical nanoflowers WO<sub>3</sub>, (<b>b</b>) the image of an individual nanoflower, (<b>c</b>) HRTEM image of a nanosheet of the nanoflower, and (<b>d</b>) the SAED pattern of an individual WO<sub>3</sub> nanosheet. Figure adapted with permission from Reference [<a href="#B41-processes-12-02605" class="html-bibr">41</a>]. Copyright 2015 Elsevier.</p>
Full article ">Figure 5
<p>(<b>a</b>) SEM image depicting DNA-like double helical tripodal microfiber superstructure along with a representative DNA image shown in the inset for comparison. (<b>b</b>–<b>d</b>) represents HRTEM and SAED analysis of this structure. (<b>b</b>) The yellow dashed lines indicate the position of WO<sub>3−x</sub> nanorods. (<b>c</b>) Carbon encapsulation around WO<sub>3−x</sub> nanorods. (<b>d</b>) The yellow dashed line indicates the WO<sub>3−x</sub>/C heterostructure. Figure reprinted with permission from Reference [<a href="#B101-processes-12-02605" class="html-bibr">101</a>]. Copyright 2021 ACS.</p>
Full article ">Figure 6
<p>Schematic illustration of step-by-step sol-gel method. Adapted with permission from Reference [<a href="#B103-processes-12-02605" class="html-bibr">103</a>].</p>
Full article ">Figure 7
<p>FESEM and SEM images for (<b>a</b>) 0D, (<b>b</b>) 1D, (<b>c</b>) 2D, and (<b>d</b>) 3D WO<sub>3</sub> nanostructures. Adapted with permission from References [<a href="#B118-processes-12-02605" class="html-bibr">118</a>,<a href="#B119-processes-12-02605" class="html-bibr">119</a>,<a href="#B120-processes-12-02605" class="html-bibr">120</a>,<a href="#B121-processes-12-02605" class="html-bibr">121</a>].</p>
Full article ">Figure 8
<p>SEM images of a-WO<sub>3</sub>/OP-c-WO<sub>3</sub> films obtained with electrodeposition times of (<b>a</b>) 150 s, (<b>b</b>) 200 s, (<b>c</b>) 250 s, and (<b>d</b>) 300 s. The insets are AFM and SEM images of the cross-section of the samples. Figure reprinted with permission from Reference [<a href="#B134-processes-12-02605" class="html-bibr">134</a>]. Copyright 2024 Elsevier.</p>
Full article ">Figure 9
<p>Schematic illustration of a typical electrospinning setup. A composite solution is ejected under high voltage, forming a stable Taylor cone that transitions to an unstable whipping jet, resulting in the deposition of nanofibers on the collector, indicated by the purple arrow. Adapted with permission from Reference [<a href="#B153-processes-12-02605" class="html-bibr">153</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) XRD spectra of the WO<sub>3</sub> NFs calcined at different heating rates SEM images of the (<b>b</b>) as-prepared nanofibers before the calcination and (<b>c</b>) calcined NF550 with heating rates (5 °C/min). (<b>d</b>) Relative selectivity for 100 ppm of NO<sub>2</sub>, H<sub>2</sub>, and CO. Reprinted with permission from Reference [<a href="#B158-processes-12-02605" class="html-bibr">158</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>−<b>c</b>) SEM images of WO<sub>3</sub>, WO<sub>3</sub>/rGO and WO<sub>3−x</sub>/rGO, respectively. (<b>d</b>,<b>e</b>) TEM images, (<b>f</b>) HRTEM image, and (<b>g</b>) SAED pattern of WO<sub>3−x</sub>/rGO sample. Reprinted with permission from Reference [<a href="#B170-processes-12-02605" class="html-bibr">170</a>]. Copyright 2020 ACS.</p>
Full article ">Figure 12
<p>(<b>a</b>) Schematic diagram of WO<sub>3</sub> nanowire synthesis processing. (<b>b</b>) SEM images of WO<sub>3</sub> nanowires synthesized with different C<sub>W</sub> with different magnifications. Reprinted with permission from Reference [<a href="#B214-processes-12-02605" class="html-bibr">214</a>]. Copyright 2020 ACS.</p>
Full article ">Figure 13
<p>Illustration of the growth of WO<sub>3</sub> nanowires by the solvothermal method. Reprinted with permission from Reference [<a href="#B215-processes-12-02605" class="html-bibr">215</a>]. Copyright 2019 Elsevier.</p>
Full article ">Figure 14
<p>SEM image of WO<sub>3</sub> nanosheets with loosely (<b>a</b>–<b>c</b>) and tightly (<b>d</b>–<b>f</b>) arranged. The image was artificially colored to differentiate between the loosely and tightly arranged nanosheets. Reprinted with permission from Reference [<a href="#B231-processes-12-02605" class="html-bibr">231</a>]. Copyright 2019 Elsevier.</p>
Full article ">Figure 15
<p>SEM images of HMT<sub>x</sub>/WCl<sub>6</sub> with different molar ratios, from 0 to 2, and annealing at 500  °C: (<b>a</b>) HMT<sub>0</sub>−WO<sub>3</sub>, (<b>b</b>) HMT<sub>0.5</sub>−WO<sub>3</sub>, (<b>c</b>) HMT<sub>1</sub>−WO<sub>3</sub>, and (<b>d</b>) HMT<sub>2</sub>−WO<sub>3</sub>. (<b>e</b>) Illustration of the morphology-controlled process. Reprinted with permission from Reference [<a href="#B250-processes-12-02605" class="html-bibr">250</a>]. Copyright 2019 Elsevier.</p>
Full article ">Figure 16
<p>(<b>a</b>) SEM images of WO<sub>3</sub> nanosheets, (<b>b</b>) enlargement of the white-circled area in (<b>a</b>), revealing nanorods dispersed among the nanosheets, circled in red, (<b>c</b>) TEM, and (<b>d</b>) HARTEM images of WO<sub>3</sub> nanorods. Reprinted with permission from Reference [<a href="#B257-processes-12-02605" class="html-bibr">257</a>]. Copyright 2016 Elsevier.</p>
Full article ">Figure 17
<p>Colloidal synthesis of NPs via (<b>a</b>) chemical reduction, (<b>b</b>) traditional thermal decomposition, and (<b>c</b>) thermal decomposition via hot injection. Reprinted with permission from Reference [<a href="#B276-processes-12-02605" class="html-bibr">276</a>]. Copyright 2020 MDPI.</p>
Full article ">
14 pages, 2522 KiB  
Article
Quantitative Investigation of Layer-by-Layer Deposition and Dissolution Kinetics by New Label-Free Analytics Based on Low-Q-Whispering Gallery Modes
by Mateusz Olszyna, Algi Domac, Jasmine Zimmer and Lars Dähne
Photonics 2024, 11(11), 1087; https://doi.org/10.3390/photonics11111087 - 19 Nov 2024
Viewed by 280
Abstract
A new instrument for label-free measurements based on optical Low-Q Whispering Gallery Modes (WGMs) for various applications is used for a detailed study of the deposition and release of Layer-by-Layer polymer coatings. The two selected coating pairs interact either via hydrogen bonding or [...] Read more.
A new instrument for label-free measurements based on optical Low-Q Whispering Gallery Modes (WGMs) for various applications is used for a detailed study of the deposition and release of Layer-by-Layer polymer coatings. The two selected coating pairs interact either via hydrogen bonding or electrostatic interactions. Their assembly was followed by common Quartz Crystal Microbalance (QCM) technology and the Low-Q WGMs. In contrast to planar QCM sensor chips of 1 cm, the WGM sensors are fluorescent spherical beads with diameters of 10.2 µm, enabling the detection of analyte quantities in the femtogram range in tiny volumes. The beads, with a very smooth surface and high refractive index, act as resonators for circular light waves that can revolve up to 10,000 times within the bead. The WGM frequencies are highly sensitive to changes in particle diameter and the refractive index of the surrounding medium. Hence, the adsorption of molecules shifts the resonance frequency, which is detected by a robust instrument with a high-resolution spectrometer. The results demonstrate the high potential of the new photonic measurement and its advantages over QCM technology, such as cheap sensors (billions in one Eppendorf tube), simple pre-functionalization, much higher statistic safety by hundreds of sensors for one measurement, 5–10 times faster analysis, and that approx. 25, 000 fewer analyte molecules are needed for one sensor. In addition, the deposited molecule amount is not superposed by hydrated water as for QCM. A connection between sensors and instruments does not exist, enabling application in any transparent environment, like microfluidics, drop-on slides, Petri dishes, well plates, cell culture vasculature, etc. Full article
(This article belongs to the Special Issue Fundamentals, Advances, and Applications in Optical Sensing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Process of LbL coating of planar surfaces by dip coating; (<b>b</b>) LbL coating of colloidal templates by washing under centrifugation or filtration processes. Subsequent removal of the core leads to hollow capsules.</p>
Full article ">Figure 2
<p>(<b>a</b>) Whispering Gallery Modes detection system; (<b>b</b>) WGM chip with a microarray of 6000 wells (magnification 300×).</p>
Full article ">Figure 3
<p>(<b>a</b>) Principle of WGM measurement: (1) 405 nm laser for excitation of the fluorescence molecules; (2) some molecules in the neighborhood of the surface emit light by chance, which hits the surface by an angle larger than the critical angle and is totally reflected back; if their wavelength is in resonance after several reflections, they form the circulating WGM (3); after up to 10,000 circulations, they are scattered out (4), and their wavelengths are measured by the spectrometer (5); the spectra are sent for signal evaluation to the software (WhisperSense v1.1.3) (6). (<b>b</b>) Typical WGM spectrum of 10.2 µm polystyrene fluorescent microbeads before (black) and after adsorption of molecules (red, shifted). (<b>c</b>) Calculated mass of one polyelectrolyte layer adsorption over time.</p>
Full article ">Figure 4
<p>Deposition of one double layer of hydrogen-bonded PVPon/PMAA and electrostatically bonded PAH/PSS on a basic PVPon/PMAA or PAH/PSS coating, respectively (<b>a</b>) followed by WGM analytics; (<b>b</b>) followed by QCM analytics.</p>
Full article ">Figure 5
<p>Deposition of eight double layers of hydrogen-bonded PVPon/PMAA: (<b>a</b>) followed by WGM analytics; (<b>b</b>) followed by QCM analytics.</p>
Full article ">Figure 6
<p>Dissolution kinetics for (PVPon/PMAA)<sub>2</sub> at increasing pH values of different buffer solutions followed by WGM: the zero value corresponds to the sensor beads pre-coated with PAH/PSS/PAH/PMAA.</p>
Full article ">Figure 7
<p>(<b>a</b>) Assembling and degradation of the (PVPON/PMAA)<sub>3</sub>/(PAH/PSS)<sub>3.5</sub> salt-free assembling, followed by WGM; (<b>b</b>) assembling and degradation of the (PVPON/PMAA)<sub>3</sub>/(PAH/PSS)<sub>3.5</sub> assembled with 0.2 M salt, followed by WGM; (<b>c</b>) assembling and degradation of the (PVPON/PMAA)<sub>3</sub>/(PAH/PSS)<sub>3.5</sub> assembled with 0.2 M salt, followed by QCM; (<b>d</b>) degradation of the (PVPON/PMAA)<sub>3</sub>/(PAH-Rho/PSS)<sub>3.5</sub> assembled with 0.2 M salt on a glass slide, followed by Confocal Laser Scanning Microscopy (CLSM).</p>
Full article ">
21 pages, 12271 KiB  
Article
Detection of Marine Oil Spill from PlanetScope Images Using CNN and Transformer Models
by Jonggu Kang, Chansu Yang, Jonghyuk Yi and Yangwon Lee
J. Mar. Sci. Eng. 2024, 12(11), 2095; https://doi.org/10.3390/jmse12112095 - 19 Nov 2024
Viewed by 265
Abstract
The contamination of marine ecosystems by oil spills poses a significant threat to the marine environment, necessitating the prompt and effective implementation of measures to mitigate the associated damage. Satellites offer a spatial and temporal advantage over aircraft and unmanned aerial vehicles (UAVs) [...] Read more.
The contamination of marine ecosystems by oil spills poses a significant threat to the marine environment, necessitating the prompt and effective implementation of measures to mitigate the associated damage. Satellites offer a spatial and temporal advantage over aircraft and unmanned aerial vehicles (UAVs) in oil spill detection due to their wide-area monitoring capabilities. While oil spill detection has traditionally relied on synthetic aperture radar (SAR) images, the combined use of optical satellite sensors alongside SAR can significantly enhance monitoring capabilities, providing improved spatial and temporal coverage. The advent of deep learning methodologies, particularly convolutional neural networks (CNNs) and Transformer models, has generated considerable interest in their potential for oil spill detection. In this study, we conducted a comprehensive and objective comparison to evaluate the suitability of CNN and Transformer models for marine oil spill detection. High-resolution optical satellite images were used to optimize DeepLabV3+, a widely utilized CNN model; Swin-UPerNet, a representative Transformer model; and Mask2Former, which employs a Transformer-based architecture for both encoding and decoding. The results of cross-validation demonstrate a mean Intersection over Union (mIoU) of 0.740, 0.840 and 0.804 for all the models, respectively, indicating their potential for detecting oil spills in the ocean. Additionally, we performed a histogram analysis on the predicted oil spill pixels, which allowed us to classify the types of oil. These findings highlight the considerable promise of the Swin Transformer models for oil spill detection in the context of future marine disaster monitoring. Full article
(This article belongs to the Special Issue Remote Sensing Applications in Marine Environmental Monitoring)
Show Figures

Figure 1

Figure 1
<p>Examples of image processing steps: (<b>a</b>) original satellite images, (<b>b</b>) images after gamma correction and histogram adjustment, and (<b>c</b>) labeled images.</p>
Full article ">Figure 2
<p>Flowchart of this study, illustrating the processes of labeling, modeling, optimization, and evaluation using the DeepLabV3+, Swin-UPerNet, and Mask2Former models [<a href="#B23-jmse-12-02095" class="html-bibr">23</a>,<a href="#B24-jmse-12-02095" class="html-bibr">24</a>,<a href="#B25-jmse-12-02095" class="html-bibr">25</a>].</p>
Full article ">Figure 3
<p>Concept of the 5-fold cross-validation in this study.</p>
Full article ">Figure 4
<p>Examples of image data augmentation using the Albumentations library. The example images include random 90-degree rotation, horizontal flip, vertical flip, optical distortion, grid distortion, RGB shift, and random brightness/contrast adjustment.</p>
Full article ">Figure 5
<p>Randomly selected examples from fold 1, including PlanetScope RGB images, segmentation labels, and predictions from DeepLabV3+ (DL), Swin-UPerNet (Swin), and Mask2Former (M2F).</p>
Full article ">Figure 6
<p>Randomly selected examples from fold 2, including PlanetScope RGB images, segmentation labels, and predictions from DeepLabV3+ (DL), Swin-UPerNet (Swin), and Mask2Former (M2F).</p>
Full article ">Figure 7
<p>Randomly selected examples from fold 3, including PlanetScope RGB images, segmentation labels, and predictions from DeepLabV3+ (DL), Swin-UPerNet (Swin), and Mask2Former (M2F).</p>
Full article ">Figure 8
<p>Randomly selected examples from fold 4, including PlanetScope RGB images, segmentation labels, and predictions from DeepLabV3+ (DL), Swin-UPerNet (Swin), and Mask2Former (M2F).</p>
Full article ">Figure 9
<p>Randomly selected examples from fold 5, including PlanetScope RGB images, segmentation labels, and predictions from DeepLabV3+ (DL), Swin-UPerNet (Swin), and Mask2Former (M2F).</p>
Full article ">Figure 10
<p>Thick oil layers with a dark black tone: histogram distribution graph and box plot of oil spill pixels extracted from the labels, DeepLabV3+, Swin-UPerNet, and Mask2Former. The <span class="html-italic">x</span>-axis values represent the digital numbers (DNs) from PlanetScope images. (<b>a</b>) Oil mask, (<b>b</b>) histogram, and (<b>c</b>) box plot.</p>
Full article ">Figure 11
<p>Thin oil layers with a bright silver tone: histogram distribution graph and box plot of oil spill pixels extracted from the labels, DeepLabV3+, Swin-UPerNet, and Mask2Former. The <span class="html-italic">x</span>-axis values represent the digital numbers (DNs) from PlanetScope images. (<b>a</b>) Oil mask, (<b>b</b>) histogram, and (<b>c</b>) box plot.</p>
Full article ">Figure 12
<p>Thin oil layers with a bright rainbow tone: histogram distribution graph and box plot of oil spill pixels extracted from the labels, DeepLabV3+, Swin-UPerNet, and Mask2Former. The <span class="html-italic">x</span>-axis values represent the digital numbers (DNs) from PlanetScope images. (<b>a</b>) Oil mask, (<b>b</b>) histogram, and (<b>c</b>) box plot.</p>
Full article ">
1 pages, 143 KiB  
Correction
Correction: Kaur, M.; Menon, C. Submillimeter Sized 2D Electrothermal Optical Fiber Scanner. Sensors 2023, 23, 404
by Mandeep Kaur and Carlo Menon
Sensors 2024, 24(22), 7346; https://doi.org/10.3390/s24227346 - 18 Nov 2024
Viewed by 131
Abstract
The Editorial Office and Editorial Board of Sensors are jointly issuing a resolution and removal of the Journal Notice linked to this article [...] Full article
(This article belongs to the Section Physical Sensors)
17 pages, 3996 KiB  
Article
The Influence of Relative Humidity and Pollution on the Meteorological Optical Range During Rainy and Dry Months in Mexico City
by Blanca Adilen Miranda-Claudes and Guillermo Montero-Martínez
Atmosphere 2024, 15(11), 1382; https://doi.org/10.3390/atmos15111382 - 16 Nov 2024
Viewed by 296
Abstract
The Meteorological Optical Range (MOR) is a measurement of atmospheric visibility. Visibility impairment has been linked to increased aerosol levels in the air. This study conducted statistical analyses using meteorological, air pollutant concentration, and MOR data collected in Mexico City from [...] Read more.
The Meteorological Optical Range (MOR) is a measurement of atmospheric visibility. Visibility impairment has been linked to increased aerosol levels in the air. This study conducted statistical analyses using meteorological, air pollutant concentration, and MOR data collected in Mexico City from August 2014 to December 2015 to determine the factors contributing to haze occurrence (periods when MOR < 10,000 m), defined using a light scatter sensor (PWS100). The outcomes revealed seasonal patterns in PM2.5 and relative humidity (RH) for haze occurrence along the year. PM2.5 levels during hazy periods in the dry season were higher compared to the wet season, aligning with periods of poor air quality (PM2.5 > 45 μg/m3). Pollutant-to-CO ratios suggested that secondary aerosols’ production, led by SO2 conversion to sulfate particles, mainly impacts haze occurrence during the dry season. Meanwhile, during the rainy season, the PWS100 registered haze events even with PM2.5 values close to 15 μg/m3 (considered good air quality). The broadened distribution of extinction efficiency during the wet period and its correlation with RH suggest that aerosol water vapor uptake significantly impacts visibility during this season. Therefore, attributing poor visibility strictly to poor air quality may not be appropriate for all times and locations. Full article
(This article belongs to the Section Meteorology)
Show Figures

Figure 1

Figure 1
<p>The research methodology overview. Blue boxes represent the main phases/sections of the study, green boxes represent how the analysis was carried out, and the yellow box leads to the discussion of results.</p>
Full article ">Figure 2
<p>Time series for Meteorological Optical Range (<span class="html-italic">MOR</span>, black lines), meteorological, and pollutant (PM<sub>2.5</sub>, NO<sub>x</sub>, SO<sub>2</sub>, and CO) measurements from 22 to 23 November 2015. <span class="html-italic">MOR</span> data show a haze event on 23 November 2015. The upper panel (<b>a</b>) shows a comparison between PM<sub>2.5</sub>, NO<sub>x</sub>, and <span class="html-italic">RH</span> (red, blue, and yellow lines, respectively) measurements correlated with <span class="html-italic">MOR</span> data. The bottom panel (<b>b</b>) displays the SO<sub>2</sub>, CO, and <span class="html-italic">WS</span> (orange, blue, and green lines, respectively) estimates during the same period. It is observed that pollutant concentrations show higher levels during the haze occurrence. See more details in the text.</p>
Full article ">Figure 3
<p>The correlation matrix showing the relationship between <span class="html-italic">MOR</span> and meteorological and pollutants variables. Bold numbers in the green-colored cells indicate statistically significant results.</p>
Full article ">Figure 4
<p>The series of monthly averages of <span class="html-italic">MOR</span>, meteorological, and pollutant measurements obtained for haze (orange) and control (blue) periods. The information is displayed for the months when haze events occurred, so November 2014 and January, March, and October 2015 are missing. The open symbols indicate results obtained for the dry season. Each subfigure shows the comparison for the variables as: (<b>a</b>) <span class="html-italic">MOR</span>, (<b>b</b>) PM<sub>2.5</sub>, (<b>c</b>) <span class="html-italic">RH</span>, (<b>d</b>) NO<sub>x</sub>, (<b>e</b>) <span class="html-italic">WS</span>, (<b>f</b>) SO<sub>2</sub>, and (<b>g</b>) <span class="html-italic">WDIR</span>.</p>
Full article ">Figure 5
<p>The dispersion of <span class="html-italic">MOR</span> values, categorized into haze (<span class="html-italic">MOR</span> &lt; 10,000 m, blue points) and non-haze (<span class="html-italic">MOR</span> &gt; 10,000 m, orange points) classes, as a function of <span class="html-italic">RH</span> and PM<sub>2.5</sub> for the dry (<b>left panel</b>) and the precipitating (<b>right panel</b>) seasons.</p>
Full article ">Figure 6
<p>The contribution of particulate (PM<sub>2.5</sub>) pollution levels in four visibility ranges during the two chosen precipitation periods. The upper panel shows that bad air quality conditions contribute significantly (up to 60%) to haze occurrence during the low precipitation period.</p>
Full article ">Figure 7
<p>Estimates of (<b>a</b>) PM<sub>2.5</sub>/CO (μg/m<sup>3</sup>/ppmv), (<b>b</b>) SO<sub>2</sub>/CO (ppbv/ppmv), and (<b>c</b>) NO<sub>x</sub>/CO (ppbv/ppmv) ratios for two <span class="html-italic">MOR</span> ranges (shown in the <span class="html-italic">x</span>-axis of the bottom panel). Orange and blue bars show the mean values for each ratio during the representative periods of haze and good <span class="html-italic">MOR</span> estimates, respectively. The vertical bars correspond to the standard deviation of the mean values. Under different visibility conditions, these ratios are useful as a proxy for the contribution of gas–particle conversion processes. See details in the text.</p>
Full article ">Figure 8
<p>Frequency distributions of the extinction capacity of PM<sub>2.5</sub> per unit mass under diverse <span class="html-italic">RH</span> ranges: (<b>a</b>) 40 % &lt; <span class="html-italic">RH</span> &lt; 60 %, (<b>b</b>) 60 % &lt; <span class="html-italic">RH</span> &lt; 80 %, and (<b>c</b>) 80 % ≤ <span class="html-italic">RH.</span> The obtained distributions are displayed for the dry and rainy seasons.</p>
Full article ">Figure 9
<p>Cumulative curves of haze periods as a function of the PM<sub>2.5</sub> levels (<b>a</b>) and <span class="html-italic">RH</span> (<b>b</b>) during the two chosen seasons. The 50% frequency level was used to determine the particulate and moisture threshold values for haze incidence at the sampling site during the rainy and low precipitation seasons.</p>
Full article ">
21 pages, 2882 KiB  
Review
Gold Nanoprobes for Robust Colorimetric Detection of Nucleic Acid Sequences Related to Disease Diagnostics
by Maria Enea, Andreia Leite, Ricardo Franco and Eulália Pereira
Nanomaterials 2024, 14(22), 1833; https://doi.org/10.3390/nano14221833 - 16 Nov 2024
Viewed by 356
Abstract
Gold nanoparticles (AuNPs) are highly attractive for applications in the field of biosensing, particularly for colorimetric nucleic acid detection. Their unique optical properties, which are highly sensitive to changes in their environment, make them ideal candidates for developing simple, rapid, and cost-effective assays. [...] Read more.
Gold nanoparticles (AuNPs) are highly attractive for applications in the field of biosensing, particularly for colorimetric nucleic acid detection. Their unique optical properties, which are highly sensitive to changes in their environment, make them ideal candidates for developing simple, rapid, and cost-effective assays. When functionalized with oligonucleotides (Au-nanoprobes), they can undergo aggregation or dispersion in the presence of complementary sequences, leading to distinct color changes that serve as a visual signal for detection. Aggregation-based assays offer significant advantages over other homogeneous assays, such as fluorescence-based methods, namely, label-free protocols, rapid interactions in homogeneous solutions, and detection by the naked eye or using low-cost instruments. Despite promising results, the application of Au-nanoprobe-based colorimetric assays in complex biological matrices faces several challenges. The most significant are related to the colloidal stability and oligonucleotide functionalization of the Au-nanoprobes but also to the mode of detection. The type of functionalization method, type of spacer, the oligo–AuNPs ratio, changes in pH, temperature, or ionic strength influence the Au-nanoprobe colloidal stability and thus the performance of the assay. This review elucidates characteristics of the Au-nanoprobes that are determined for colorimetric gold nanoparticles (AuNPs)-based nucleic acid detection, and how they influence the sensitivity and specificity of the colorimetric assay. These characteristics of the assay are fundamental to developing low-cost, robust biomedical sensors that perform effectively in biological fluids. Full article
(This article belongs to the Special Issue Noble Metal-Based Nanostructures: Optical Properties and Applications)
Show Figures

Figure 1

Figure 1
<p>Timeline of AuNPs use for nucleic acid detection.</p>
Full article ">Figure 2
<p>Dependence of LSPR on spherical gold nanoparticles diameter and aggregation state.</p>
Full article ">Figure 3
<p>Colorimetric detection methods using spherical AuNPs: (Top panel) Cross-linking assay—a color change occurs as nucleic acid sequence strands specifically hybridize with complementary sequences, reducing the distance between particles, and resulting in a blue solution (positive test). In the absence of complementary sequences, the solution stays red (negative test). (Middle panel) Non-cross-linking assay—an increase in ionic strength induces AuNP aggregation, resulting in a blue solution (negative test). When complementary targets are present, the solution stays red (positive test). (Bottom panel) Colorimetric assay using unmodified AuNPs: In the absence of complementary sequences, only single-stranded DNA (ssDNA) is present, stabilizing AuNPs against salt-induced aggregation, and the solution stays red (negative result). Conversely, when hybridization occurs in the presence of a complementary sequence, double-stranded DNA (dsDNA) forms, and aggregation occurs (blue solution is a positive result). UV/vis spectra and Nanoparticle Tracking analysis (NTA) profiles are shown with blue lines corresponding to aggregated AuNPs samples and red lines to non-aggregated ones. Also indicated are the positive (green check) and negative (red cross) results for each test.</p>
Full article ">Figure 4
<p>Published successful functionalization methods of AuNPs with HS-oligos, resulting in Au-nanoprobes.</p>
Full article ">Figure 5
<p>Examples of Au nanoparticle interaction with (i) ssDNA, (ii) PolyA-ssDNA and PolyT-ssDNA, (iii) PEG-ssDNA, and (iv) thiolated-(CH2)6-ssDNA.</p>
Full article ">
17 pages, 5463 KiB  
Article
Geographically-Informed Modeling and Analysis of Platform Attitude Jitter in GF-7 Sub-Meter Stereo Mapping Satellite
by Haoran Xia, Xinming Tang, Fan Mo, Junfeng Xie and Xiang Li
ISPRS Int. J. Geo-Inf. 2024, 13(11), 413; https://doi.org/10.3390/ijgi13110413 - 15 Nov 2024
Viewed by 362
Abstract
The GF-7 satellite, China’s inaugural sub-meter-level stereoscopic mapping satellite, has been deployed for a wide range of applications, including natural resource investigation, environmental monitoring, fundamental surveying, and the development of global geospatial information resources. The satellite’s stable platform and reliable imaging systems are [...] Read more.
The GF-7 satellite, China’s inaugural sub-meter-level stereoscopic mapping satellite, has been deployed for a wide range of applications, including natural resource investigation, environmental monitoring, fundamental surveying, and the development of global geospatial information resources. The satellite’s stable platform and reliable imaging systems are crucial for achieving high-quality imaging and precise attitude measurements. However, the satellite’s operation is affected by both internal and external factors, which induce vibrations in the satellite platform, thereby affecting image quality and mapping accuracy. To address this challenge, this paper proposes a novel method for constructing a satellite platform vibration model based on geographic location information. The model is developed by integrating composite data from star sensors and gyroscopes (gyro) with subsatellite point location data. The experimental methodology involves the composite processing of gyro data and star sensor optical axis angles, integration of the processed data through time-matching and normalization, and denoising of the integrated data, followed by trigonometric fitting to capture the periodic characteristics of platform vibrations. The positions of the satellite substellar points are determined from the satellite orbit data. A rigorous geometric imaging model is then used to construct a vibration model with geographic location correlation in combination with the satellite subsatellite point positions. The experimental results demonstrate the following: (1) Over the same temporal range, there is a significant convergence in the waveform similarities between the gyro data and the star sensor optical axis angles, indicating a strong correlation in the jitter information; (2) The platform vibration exhibits a robust correlation with the satellite’s geographic location along its orbit. Specifically, the model reveals that the GF-7 satellite experiences the maximum vibration amplitude between 5° S and 20° S latitude during its ascending phase, and the minimum vibration amplitude between 5° N and 20° N latitude during the descending phase. The model established in this study offers theoretical support for optimizing satellite attitude and mitigating platform vibrations. Full article
Show Figures

Figure 1

Figure 1
<p>Comparative analysis of real and ideal vibration information from the star sensor and the gyro.</p>
Full article ">Figure 2
<p>Flow chart for vibration modeling method of the GF-7 satellite.</p>
Full article ">Figure 3
<p>Time consistency analysis of gyro data and star sensor data.</p>
Full article ">Figure 4
<p>Time offset analysis in composite data from the gyro and star sensor. Time Offset is the temporal difference between the two datasets. T1 is a specific time point.</p>
Full article ">Figure 5
<p>Schematic of the moving average filter.</p>
Full article ">Figure 6
<p>Satellite orbital data for calculating geographic locations of subsatellite points.</p>
Full article ">Figure 7
<p>Filtered gyro data across multiple satellite tracks: (<b>a</b>) Track 016396; (<b>b</b>) Track 016416; (<b>c</b>) Track 016431; (<b>d</b>) Track 016446.</p>
Full article ">Figure 8
<p>Optical axis clamping angle measurements from star sensor.</p>
Full article ">Figure 9
<p>Composite analysis of gyro and star sensor pinch angle data.</p>
Full article ">Figure 10
<p>Denoising results using moving average filter.</p>
Full article ">Figure 11
<p>Results of trigonometric function fitting for vibration data analysis. The blue color indicates composite data. The red curve represents the fitted result.</p>
Full article ">Figure 12
<p>Geographic distribution of satellite orbital paths.</p>
Full article ">Figure 13
<p>Variations in satellite flutter amplitude relative to geographic location.</p>
Full article ">
32 pages, 11087 KiB  
Article
Path Planning and Motion Control of Robot Dog Through Rough Terrain Based on Vision Navigation
by Tianxiang Chen, Yipeng Huangfu, Sutthiphong Srigrarom and Boo Cheong Khoo
Sensors 2024, 24(22), 7306; https://doi.org/10.3390/s24227306 - 15 Nov 2024
Viewed by 578
Abstract
This article delineates the enhancement of an autonomous navigation and obstacle avoidance system for a quadruped robot dog. Part one of this paper presents the integration of a sophisticated multi-level dynamic control framework, utilizing Model Predictive Control (MPC) and Whole-Body Control (WBC) from [...] Read more.
This article delineates the enhancement of an autonomous navigation and obstacle avoidance system for a quadruped robot dog. Part one of this paper presents the integration of a sophisticated multi-level dynamic control framework, utilizing Model Predictive Control (MPC) and Whole-Body Control (WBC) from MIT Cheetah. The system employs an Intel RealSense D435i depth camera for depth vision-based navigation, which enables high-fidelity 3D environmental mapping and real-time path planning. A significant innovation is the customization of the EGO-Planner to optimize trajectory planning in dynamically changing terrains, coupled with the implementation of a multi-body dynamics model that significantly improves the robot’s stability and maneuverability across various surfaces. The experimental results show that the RGB-D system exhibits superior velocity stability and trajectory accuracy to the SLAM system, with a 20% reduction in the cumulative velocity error and a 10% improvement in path tracking precision. The experimental results also show that the RGB-D system achieves smoother navigation, requiring 15% fewer iterations for path planning, and a 30% faster success rate recovery in challenging environments. The successful application of these technologies in simulated urban disaster scenarios suggests promising future applications in emergency response and complex urban environments. Part two of this paper presents the development of a robust path planning algorithm for a robot dog on a rough terrain based on attached binocular vision navigation. We use a commercial-of-the-shelf (COTS) robot dog. An optical CCD binocular vision dynamic tracking system is used to provide environment information. Likewise, the pose and posture of the robot dog are obtained from the robot’s own sensors, and a kinematics model is established. Then, a binocular vision tracking method is developed to determine the optimal path, provide a proposal (commands to actuators) of the position and posture of the bionic robot, and achieve stable motion on tough terrains. The terrain is assumed to be a gentle uneven terrain to begin with and subsequently proceeds to a more rough surface. This work consists of four steps: (1) pose and position data are acquired from the robot dog’s own inertial sensors, (2) terrain and environment information is input from onboard cameras, (3) information is fused (integrated), and (4) path planning and motion control proposals are made. Ultimately, this work provides a robust framework for future developments in the vision-based navigation and control of quadruped robots, offering potential solutions for navigating complex and dynamic terrains. Full article
Show Figures

Figure 1

Figure 1
<p>Simplified box model of the Lite3P quadruped robotic dog.</p>
Full article ">Figure 2
<p>Internal sensor arrangement of the quadruped robotic dog.</p>
Full article ">Figure 3
<p>Dynamic control flowchart.</p>
Full article ">Figure 4
<p>MPC flowchart.</p>
Full article ">Figure 5
<p>WBC flowchart [<a href="#B30-sensors-24-07306" class="html-bibr">30</a>].</p>
Full article ">Figure 6
<p>Robot coordinates and joint point settings [<a href="#B30-sensors-24-07306" class="html-bibr">30</a>].</p>
Full article ">Figure 7
<p>Intel D435i and velodyne LIDAR.</p>
Full article ">Figure 8
<p>ICP diagram.</p>
Full article ">Figure 9
<p>Comparison of before and after modifying the perception region.</p>
Full article ">Figure 10
<p>Point cloud processing flowchart.</p>
Full article ">Figure 11
<p>{p, v} generation: (<b>a</b>) the creation of {p, v} pairs for collision points; (<b>b</b>) the process of generating anchor points and repulsive vectors for dynamic obstacle avoidance [<a href="#B41-sensors-24-07306" class="html-bibr">41</a>].</p>
Full article ">Figure 12
<p>Overall framework of 2D EGO-Planner.</p>
Full article ">Figure 13
<p>Robot initialization and control process in Gazebo simulation: (<b>a</b>) Gazebo environment creation, (<b>b</b>) robot model import, (<b>c</b>) torque balance mode activation, and (<b>d</b>) robot stepping and rotation in simulation.</p>
Full article ">Figure 14
<p>Joint rotational angles of FL and RL legs.</p>
Full article ">Figure 15
<p>Joint angular velocities of FL and RL legs.</p>
Full article ">Figure 16
<p>Torque applied to FL and RL joints during the gait cycle.</p>
Full article ">Figure 17
<p>The robot navigating in a simple environment using a camera.</p>
Full article ">Figure 18
<p>The robot navigating in a complex environment using a camera.</p>
Full article ">Figure 19
<p>A 2D trajectory showing start and goal positions, obstacles, and rough path.</p>
Full article ">Figure 20
<p>Initial environment setup.</p>
Full article ">Figure 21
<p>The robot starts navigating in a simple environment with a static obstacle (brown box).</p>
Full article ">Figure 22
<p>Dynamic Obstacle 1 introduced: the robot detects a new obstacle and recalculates its path.</p>
Full article ">Figure 23
<p>Dynamic Obstacle 2 introduced: after avoiding the first obstacle, a second obstacle is introduced and detected by the planner.</p>
Full article ">Figure 24
<p>Approaching the target: the robot adjusts its path to approach the target point as the distance shortens.</p>
Full article ">Figure 25
<p>Reaching the target: the robot completes its path and reaches the designated target point.</p>
Full article ">Figure 26
<p>Real-time B-spline trajectory updates in response to dynamic obstacles. Set 1 (orange) shows the initial path avoiding static obstacles. When the first dynamic obstacle is detected, the EGO-Planner updates the path (Set 2, blue) using local optimization. A second obstacle prompts another adjustment (Set 3, green), guiding the robot smoothly towards the target as trajectory updates become more frequent.</p>
Full article ">Figure 27
<p>The robot navigating a simple environment using SLAM.</p>
Full article ">Figure 28
<p>The robot navigating a complex environment using SLAM.</p>
Full article ">Figure 29
<p>A 2D trajectory showing start and goal positions, obstacles, and the planned path in a complex environment using SLAM.</p>
Full article ">Figure 30
<p>Navigation based on RGB-D camera.</p>
Full article ">Figure 31
<p>Navigation based on SLAM.</p>
Full article ">Figure 32
<p>Velocity deviation based on RGB-D camera.</p>
Full article ">Figure 33
<p>Velocity deviation based on SLAM.</p>
Full article ">Figure 34
<p>Cumulative average iterations.</p>
Full article ">Figure 35
<p>Cumulative success rate.</p>
Full article ">
22 pages, 5816 KiB  
Article
Causality-Driven Feature Selection for Calibrating Low-Cost Airborne Particulate Sensors Using Machine Learning
by Vinu Sooriyaarachchi, David J. Lary, Lakitha O. H. Wijeratne and John Waczak
Sensors 2024, 24(22), 7304; https://doi.org/10.3390/s24227304 - 15 Nov 2024
Viewed by 408
Abstract
With escalating global environmental challenges and worsening air quality, there is an urgent need for enhanced environmental monitoring capabilities. Low-cost sensor networks are emerging as a vital solution, enabling widespread and affordable deployment at fine spatial resolutions. In this context, machine learning for [...] Read more.
With escalating global environmental challenges and worsening air quality, there is an urgent need for enhanced environmental monitoring capabilities. Low-cost sensor networks are emerging as a vital solution, enabling widespread and affordable deployment at fine spatial resolutions. In this context, machine learning for the calibration of low-cost sensors is particularly valuable. However, traditional machine learning models often lack interpretability and generalizability when applied to complex, dynamic environmental data. To address this, we propose a causal feature selection approach based on convergent cross mapping within the machine learning pipeline to build more robustly calibrated sensor networks. This approach is applied in the calibration of a low-cost optical particle counter OPC-N3, effectively reproducing the measurements of PM1 and PM2.5 as recorded by research-grade spectrometers. We evaluated the predictive performance and generalizability of these causally optimized models, observing improvements in both while reducing the number of input features, thus adhering to the Occam’s razor principle. For the PM1 calibration model, the proposed feature selection reduced the mean squared error on the test set by 43.2% compared to the model with all input features, while the SHAP value-based selection only achieved a reduction of 29.6%. Similarly, for the PM2.5 model, the proposed feature selection led to a 33.2% reduction in the mean squared error, outperforming the 30.2% reduction achieved by the SHAP value-based selection. By integrating sensors with advanced machine learning techniques, this approach advances urban air quality monitoring, fostering a deeper scientific understanding of microenvironments. Beyond the current test cases, this feature selection method holds potential for broader applications in other environmental monitoring applications, contributing to the development of interpretable and robust environmental models. Full article
(This article belongs to the Section Sensor Networks)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Attractor manifold of the canonical Lorenz system (<span class="html-italic">M</span>) plotted in 3D space, showing the trajectory of the original system in the state space with variables <span class="html-italic">X</span>, <span class="html-italic">Y</span>, and <span class="html-italic">Z</span>. (<b>b</b>) Reconstructed manifold <math display="inline"><semantics> <msub> <mi>M</mi> <mi>X</mi> </msub> </semantics></math> using delay-coordinate embedding of the <span class="html-italic">X</span> variable. The coordinates <math display="inline"><semantics> <mrow> <mi>X</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>X</mi> <mo>(</mo> <mi>t</mi> <mo>−</mo> <mi>τ</mi> <mo>)</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>X</mi> <mo>(</mo> <mi>t</mi> <mo>−</mo> <mn>2</mn> <mi>τ</mi> <mo>)</mo> </mrow> </semantics></math> approximate the original attractor dynamics, capturing the structure of the system dynamics based only on the <span class="html-italic">X</span> time series. (<b>c</b>) Reconstructed manifold <math display="inline"><semantics> <msub> <mi>M</mi> <mi>Y</mi> </msub> </semantics></math> using delay-coordinate embedding of the <span class="html-italic">Y</span> variable. The coordinates <math display="inline"><semantics> <mrow> <mi>Y</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>(</mo> <mi>t</mi> <mo>−</mo> <mi>τ</mi> <mo>)</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>Y</mi> <mo>(</mo> <mi>t</mi> <mo>−</mo> <mn>2</mn> <mi>τ</mi> <mo>)</mo> </mrow> </semantics></math> again form an attractor diffeomorphic to the original manifold, illustrating how the <span class="html-italic">Y</span> time series alone, through lagged coordinates, captures the dynamics of the system.</p>
Full article ">Figure 2
<p>Proposed causality-driven feature selection pipeline.</p>
Full article ">Figure 3
<p>Input features to the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mn>1</mn> </msub> </mrow> </semantics></math> calibration model ranked in descending order of mean absolute SHAP values. The 10 highest-ranked features are highlighted in red.</p>
Full article ">Figure 4
<p>Potential input features to the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mn>1</mn> </msub> </mrow> </semantics></math> calibration model ranked in descending order of strength of causal influence after eliminating features with <span class="html-italic">p</span>-value <math display="inline"><semantics> <mrow> <mo>≥</mo> <mn>0.05</mn> </mrow> </semantics></math>. The 10 highest-ranked features are highlighted in red.</p>
Full article ">Figure 5
<p>Input features to the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mrow> <mn>2.5</mn> </mrow> </msub> </mrow> </semantics></math> calibration model ranked in descending order of mean absolute SHAP values. The 10 highest-ranked features are highlighted in red.</p>
Full article ">Figure 6
<p>Potential input features to the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mrow> <mn>2.5</mn> </mrow> </msub> </mrow> </semantics></math> calibration model ranked in descending order of strength of causal influence after eliminating features with <span class="html-italic">p</span>-value <math display="inline"><semantics> <mrow> <mo>≥</mo> <mn>0.05</mn> </mrow> </semantics></math>. The 10 highest-ranked features are highlighted in red.</p>
Full article ">Figure 7
<p>Scatter diagram comparing the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mn>1</mn> </msub> </mrow> </semantics></math> measurements from the reference instrument on the x-axis against the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mn>1</mn> </msub> </mrow> </semantics></math> estimates from OPC-N3 on the y-axis prior to calibration.</p>
Full article ">Figure 8
<p>Density plots of the residuals for the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mn>1</mn> </msub> </mrow> </semantics></math> calibration models derived from each approach.</p>
Full article ">Figure 9
<p>Scatter diagrams for the calibration models with the x-axis showing the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mn>1</mn> </msub> </mrow> </semantics></math> count from the reference instrument and the y-axis showing the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mn>1</mn> </msub> </mrow> </semantics></math> count provided by calibrating the LCS: (<b>a</b>) Without any feature selection. (<b>b</b>) SHAP value-based feature selection. (<b>c</b>) Causality-based feature selection. (<b>d</b>) Comparison of true vs. predicted values for the test set across models.</p>
Full article ">Figure 10
<p>Scatter diagram comparing the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mrow> <mn>2.5</mn> </mrow> </msub> </mrow> </semantics></math> measurements from the reference instrument on the x-axis against the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mrow> <mn>2.5</mn> </mrow> </msub> </mrow> </semantics></math> estimates from OPC-N3 on the y-axis prior to calibration.</p>
Full article ">Figure 11
<p>Density plots of the residuals for the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mrow> <mn>2.5</mn> </mrow> </msub> </mrow> </semantics></math> calibration models derived from each approach.</p>
Full article ">Figure 12
<p>Scatter diagrams for the calibration models with the x-axis showing the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mrow> <mn>2.5</mn> </mrow> </msub> </mrow> </semantics></math> count from the reference instrument and the y-axis showing the <math display="inline"><semantics> <mrow> <msub> <mi>PM</mi> <mrow> <mn>2.5</mn> </mrow> </msub> </mrow> </semantics></math> count provided by calibrating the LCS: (<b>a</b>) Without any feature selection. (<b>b</b>) SHAP value-based feature selection. (<b>c</b>) Causality-based feature selection. (<b>d</b>) Comparison of true vs. predicted values for the test set across models.</p>
Full article ">
13 pages, 3048 KiB  
Article
Thermal Quenching of Intrinsic Photoluminescence in Amorphous and Monoclinic HfO2 Nanotubes
by Artem Shilov, Sergey Savchenko, Alexander Vokhmintsev, Kanat Zhusupov and Ilya Weinstein
Materials 2024, 17(22), 5587; https://doi.org/10.3390/ma17225587 - 15 Nov 2024
Viewed by 254
Abstract
Nanotubular hafnia arrays hold significant promise for advanced opto- and nanoelectronic applications. However, the known studies concern mostly the luminescent properties of doped HfO2-based nanostructures, while the optical properties of nominally pure hafnia with optically active centers of intrinsic origin are [...] Read more.
Nanotubular hafnia arrays hold significant promise for advanced opto- and nanoelectronic applications. However, the known studies concern mostly the luminescent properties of doped HfO2-based nanostructures, while the optical properties of nominally pure hafnia with optically active centers of intrinsic origin are far from being sufficiently investigated. In this work, for the first time we have conducted research on the wide-range temperature effects in the photoluminescence processes of anion-defective hafnia nanotubes with an amorphous and monoclinic structure, synthesized by the electrochemical oxidation method. It is shown that the spectral parameters, such as the position of the maximum and half-width of the band, remain almost unchanged in the range of 7–296 K. The experimental data obtained for the photoluminescence temperature quenching are quantitatively analyzed under the assumption made for two independent channels of non-radiative relaxation of excitations with calculating the appropriate energies of activation barriers—9 and 39 meV for amorphous hafnia nanotubes, 15 and 141 meV for monoclinic ones. The similar temperature behavior of photoluminescence spectra indicates close values of short-range order parameters in the local atomic surrounding of the active emission centers in hafnium dioxide with amorphous and monoclinic structure. Anion vacancies VO and VO2 appeared in the positions of three-coordinated oxygen and could be the main contributors to the spectral features of emission response and observed thermally stimulated processes. The recognized and clarified mechanisms occurring during thermal quenching of photoluminescence could be useful for the development of light-emitting devices and thermo-optical sensors with functional media based on oxygen-deficient hafnia nanotubes. Full article
(This article belongs to the Special Issue Advances in Luminescent Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scanning electron microscope (SEM) (<b>a</b>,<b>b</b>) and transmission electron microscope (TEM) (<b>c</b>,<b>d</b>) images obtained for the monoclinic HfO<sub>2</sub> nanotubes under study. The value marked in (<b>d</b>) corresponds to the interplanar distance <math display="inline"><semantics> <mrow> <mover accent="true"> <mn>1</mn> <mo>¯</mo> </mover> <mn>11</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Photoluminescence (PL) spectra of amorphous (<b>top</b>) and monoclinic (<b>bottom</b>) hafnia nanotubes measured at different temperatures.</p>
Full article ">Figure 3
<p>Temperature dependencies of the experimental values of the maximum position E<sub>max</sub> (blue color) and half-width FWHM (green color) of the measured PL bands. The circles indicate data for amorphous NTs, triangles—for monoclinic NTs. The dashed lines show the averaged values of E<sub>max</sub> and FWHM in the temperature range of 7–296 K.</p>
Full article ">Figure 4
<p>PL spectra of amorphous (<b>left</b>, circles) and monoclinic (<b>right</b>, triangles) nanotubes measured at a temperature of 10 K, with decomposition into Gaussian components (red lines).</p>
Full article ">Figure 5
<p>Dependence <span class="html-italic">I</span>(<span class="html-italic">T</span>) for amorphous (<b>top</b>) and monoclinic (<b>bottom</b>) NTs. The red and blue lines, see insets, are linear approximations.</p>
Full article ">
15 pages, 4260 KiB  
Article
Microwave-Assisted Synthesis of N, S Co-Doped Carbon Quantum Dots for Fluorescent Sensing of Fe(III) and Hydroquinone in Water and Cell Imaging
by Zhaochuan Yu, Chao Deng, Wenhui Ma, Yuqian Liu, Chao Liu, Tingwei Zhang and Huining Xiao
Nanomaterials 2024, 14(22), 1827; https://doi.org/10.3390/nano14221827 - 14 Nov 2024
Viewed by 515
Abstract
The detection of heavy metal ions and organic pollutants from water sources remains critical challenges due to their detrimental effects on human health and the environment. Herein, a nitrogen and sulfur co-doped carbon quantum dot (NS-CQDs) fluorescent sensor was developed using a microwave-assisted [...] Read more.
The detection of heavy metal ions and organic pollutants from water sources remains critical challenges due to their detrimental effects on human health and the environment. Herein, a nitrogen and sulfur co-doped carbon quantum dot (NS-CQDs) fluorescent sensor was developed using a microwave-assisted carbonization method for the detection of Fe3+ ions and hydroquinone (HQ) in aqueous solutions. NS-CQDs exhibit excellent optical properties, enabling sensitive detection of Fe3+ and HQ, with detection limits as low as 3.40 and 0.96 μM. Notably, with the alternating introduction of Fe3+ and HQ, NS-CQDs exhibit significant fluorescence (FL) quenching and recovery properties. Based on this property, a reliable “on-off-on” detection mechanism was established, enabling continuous and reversible detection of Fe3+ and HQ. Furthermore, the low cytotoxicity of NS-CQDs was confirmed through successful imaging of HeLa cells, indicating their potential for real-time intracellular detection of Fe3+ and HQ. This work not only provides a green and rapid synthesis strategy for CQDs but also highlights their versatility as fluorescent probes for environmental monitoring and bioimaging applications. Full article
(This article belongs to the Special Issue Nanomaterials in Electrochemical Electrode and Electrochemical Sensor)
Show Figures

Figure 1

Figure 1
<p>Synthesis and characterization of NS-CQDs. (<b>a</b>) HRTEM image of NS-CQDs (the inset shows its lattice fringes). (<b>b</b>) Diameter distribution analysis, (<b>c</b>) XRD, (<b>d</b>) AFM image, (<b>e</b>) height distribution analysis, and (<b>f</b>) Raman spectroscopy of NS-CQDs.</p>
Full article ">Figure 2
<p>Chemical composition analysis of NS-CQDs. (<b>a</b>) FTIR and (<b>b</b>) XPS spectra of NS-CQDs. (<b>c</b>–<b>f</b>) High-resolution XPS of C 1s, N 1s, S 2p, and O 1s, respectively.</p>
Full article ">Figure 3
<p>Evaluation of FL properties of NS-MQDs. (<b>a</b>) UV-vis absorption spectra and FL excitation (Ex) and FL emission (Em) spectra of NS-MQDs in aqueous solution. (<b>b</b>) FL emission spectra of NS-MQDs at different excitation wavelengths (320–480 nm).</p>
Full article ">Figure 4
<p>“On-off-on” detection performance of NS-CQDs for Fe<sup>3+</sup> and HQ. (<b>a</b>) The concentration-dependent emission spectra of NS-CQDs at different concentrations of Fe<sup>3+</sup> (0–833.3 μM). (<b>b</b>) Stern-Volmer plot of F/F<sub>0</sub> versus Fe<sup>3+</sup> concentration for NS-CQDs. (<b>c</b>) FL intensity change profiles of NS-CQDs in the presence of potential competing ions (green) and Fe<sup>3+</sup> upon addition of competing anions (3-fold excess, purple) in DI water. (<b>d</b>) The concentration-dependent emission spectra of (NS-CQDs+Fe<sup>3+</sup>) for different concentrations of HQ (0–1333.3 μM). (<b>e</b>) Stern-Volmer plot of F/F<sub>0</sub> versus HQ concentration for (NS-CQDs+Fe<sup>3+</sup>). (<b>f</b>) FL intensity change profiles of NS-CQDs in the presence of potential competing ions (green) and HQ upon addition of competing anions (3-fold excess, purple) in DI water.</p>
Full article ">Figure 5
<p>Fluorescence “on-off-on” principle of NS-CQDs. (<b>a</b>) UV-vis absorption spectrum and (<b>b</b>) FL decay curves of NS-CQDs before and after the addition of Fe<sup>3+</sup>. (<b>c</b>) FL spectrum of NS-CQDs, NS-CQDs+HQ, NS-CQDs+Fe<sup>3+</sup>, NS-CQDs Fe<sup>2+</sup>, and NS-CQDs+Fe<sup>3+</sup>+HQ. (<b>d</b>) Schematic of the proposed “on-off-on” mechanism for the detection of Fe<sup>3+</sup> and HQ using NS-CQDs.</p>
Full article ">Figure 6
<p>Confocal FL images of living HeLa cells. The images after incubating NS-CQDs at 37 °C for 4 h are as follows: (<b>a</b>) bright field, (<b>d</b>) FL, and (<b>g</b>) merged image. The images of HeLa cells stained with NS-CQDs after treatment with Fe<sup>3+</sup> (300 μM) for 4 h are: (<b>b</b>) bright field, (<b>e</b>) FL, and (<b>h</b>) merged image. The images of HeLa cells stained with NS-CQDs after treatment with Fe<sup>3+</sup> (300 μM) followed by HQ (300 μM) for 4 h are: (<b>c</b>) bright field, (<b>f</b>) FL, and (<b>i</b>) merged image.</p>
Full article ">Scheme 1
<p>Schematic of the fabrication of co-doped NS-CQDs and its applications.</p>
Full article ">
19 pages, 6266 KiB  
Article
The Optical Parameter Optimization for Brain Implant Alzheimer Sensor Using Phototherapy Angle and Wavelength Simulation (PAWS) Methodology
by So-Hyun Cho, Chang-Hee Won, Chang-Hyun Kim and Jong-Ha Lee
Sensors 2024, 24(22), 7282; https://doi.org/10.3390/s24227282 - 14 Nov 2024
Viewed by 322
Abstract
Photonic therapy is emerging as a promising method in neuroscience for addressing Alzheimer’s disease (AD). This study uses computational simulations to investigate the impact of specific wavelengths emitted by photodiodes on the light absorption rates in brain tissue for brain implant sensors. Additionally, [...] Read more.
Photonic therapy is emerging as a promising method in neuroscience for addressing Alzheimer’s disease (AD). This study uses computational simulations to investigate the impact of specific wavelengths emitted by photodiodes on the light absorption rates in brain tissue for brain implant sensors. Additionally, it presents a novel methodology that enhances light absorption via multi-parameter optimization. By adjusting the angle and wavelength of the incident light, the absorption rate was significantly enhanced using four photodiodes, each emitting at 660 nm with a power input of 3 mW. Notably, an incident angle of 20 degrees optimized light absorption and minimized thermal effects on brain tissue. The findings indicate that photodiodes within the near-infrared spectrum are suitable for low-temperature therapeutic applications in brain tissues, affirming the viability of non-invasive and safe photonic therapy. This research contributes foundational data for advancing brain implant photonic sensor design and therapeutic strategies. Furthermore, it establishes conditions for achieving high light absorption rates with minimal heat generation, identifying optimal parameters for efficient energy transfer. Full article
(This article belongs to the Special Issue (Bio)sensors for Physiological Monitoring)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Design of an implantable photonic stimulator used for Alzheimer’s disease treatment, (<b>b</b>) Implantable photonic device.</p>
Full article ">Figure 2
<p>Diagram illustrating the experimental setup for varying the incident angle from 0 to 90 degrees in 10-degree increments. The diffusion of light is visualized as a colormap, with areas of higher intensity represented in red and lower intensity areas in blue.</p>
Full article ">Figure 3
<p>(<b>a</b>) Design based on sagittal section of an actual human brain, (<b>b</b>) Simulation modeling diagrams.</p>
Full article ">Figure 4
<p>Diagram showing the distribution of energy from the LED component: (<b>a</b>) Results of the overall, (<b>b</b>) Perspective of photonics, (<b>c</b>) Frontal view of photonics.</p>
Full article ">Figure 5
<p>Absorption rates of brain tissue across different wavelengths displayed on a 2D surface. The red areas indicate regions of high energy absorption by the medium: (<b>a</b>) 660 nm, (<b>b</b>) 690 nm, (<b>c</b>) 720 nm, (<b>d</b>) 750 nm, (<b>e</b>) 780 nm, (<b>f</b>) 810 nm, (<b>g</b>) 840 nm, (<b>h</b>) 870 nm.</p>
Full article ">Figure 6
<p>Graphical representation of varying light absorption rates across different wavelengths.</p>
Full article ">Figure 7
<p>Illustration of how light absorption rates in brain tissue increase with power on a 2D surface. The red areas indicate regions of high energy absorption by the medium.: (<b>a</b>) 1 W, (<b>b</b>) 2 W, (<b>c</b>) 3 W, (<b>d</b>) 4 W, (<b>e</b>) 5 W, (<b>f</b>) 6 W, (<b>g</b>) 7 W, (<b>h</b>) 8 W.</p>
Full article ">Figure 8
<p>Graphs depicting the changes in light absorption rates as power increases.</p>
Full article ">Figure 9
<p>Analysis of how light absorption rates vary with incident angle and wavelength: (<b>a</b>) Line graph, (<b>b</b>) 2D graph with color map.</p>
Full article ">Figure 10
<p>Comprehensive analysis of light absorption rates influenced by changes in incident angle and wavelength: (<b>a</b>) Dot plot with color map, (<b>b</b>) 3D graph with color map.</p>
Full article ">Figure 11
<p>Temporal variations in temperature across different brain tissues: (<b>a</b>) cerebellum, (<b>b</b>) gray–white matter, and (<b>c</b>) CSF.</p>
Full article ">Figure 12
<p>Mesh independence test; The arrows indicate the direction of light injection.</p>
Full article ">Figure 13
<p>Mesh independence test.</p>
Full article ">Figure 14
<p>Absorption and Scattering Coefficients Not Considered with Respect to Wavelength.</p>
Full article ">Figure 15
<p>(<b>a</b>) Absorption Coefficient of Tissue as a Function of Wavelength, (<b>b</b>) Scattering Coefficient of Tissue as a Function of Wavelength.</p>
Full article ">
15 pages, 1885 KiB  
Article
Innovative Peptide-Based Plasmonic Optical Biosensor for the Determination of Cholesterol
by Ana Lia Bernardo, Anne Parra, Virginia Cebrián, Óscar Ahumada, Sergio Oddi and Enrico Dainese
Biosensors 2024, 14(11), 551; https://doi.org/10.3390/bios14110551 - 13 Nov 2024
Viewed by 534
Abstract
Plasmonic-based biosensors have gained prominence as potent optical biosensing platforms in both scientific and medical research, attributable to their enhanced sensitivity and precision in detecting biomolecular and chemical interactions. However, the detection of low molecular weight analytes with high sensitivity and specificity remains [...] Read more.
Plasmonic-based biosensors have gained prominence as potent optical biosensing platforms in both scientific and medical research, attributable to their enhanced sensitivity and precision in detecting biomolecular and chemical interactions. However, the detection of low molecular weight analytes with high sensitivity and specificity remains a complex and unresolved issue, posing significant limitations for the advancement of clinical diagnostic tools and medical device technologies. Notably, abnormal cholesterol levels are a well-established indicator of various pathological conditions; yet, the quantitative detection of the free form of cholesterol is complicated by its small molecular size, pronounced hydrophobicity, and the necessity for mediator molecules to achieve efficient sensing. In the present study, a novel strategy for cholesterol quantification was developed, leveraging a plasmonic optical readout in conjunction with a highly specific cholesterol-binding peptide (C-pept) as a biorecognition element, anchored on a functionalized silica substrate. The resulting biosensor exhibited an exceptionally low detection limit of 21.95 µM and demonstrated a linear response in the 10–200 µM range. This peptide-integrated plasmonic sensor introduces a novel one-step competitive method for cholesterol quantification, positioning itself as a highly sensitive biosensing modality for implementation within the AVAC platform, which operates using reflective dark-field microscopy. Full article
(This article belongs to the Special Issue Nanotechnology-Enabled Biosensors)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>GNP-Chol counts of the cholesterol quantification bioassay conducted by the competitive interaction of GNPs functionalized with cholesterol and free cholesterol in a PBST (0.05%) solution. All data reported are presented as the mean ± standard deviation (SD) (<span class="html-italic">n</span> ≥ 3).</p>
Full article ">Figure 2
<p>Evaluation of the sensitivity of the system. Chart-box plot of the counts of GNPs after interaction with the functionalized Si substrate in the presence (green box) and in absence (red box) of C-pept as biorecognition element. Effect of different concentrations of free cholesterol in solution when incubating GNP-Chol-PEG for 1 h at 37 °C in PBST. All data reported are presented as the mean ± standard deviation (<span class="html-italic">n</span> ≥ 3).</p>
Full article ">Figure 3
<p>Quantification of free cholesterol in solution in the competitive bioassay implementing the system based on GNPs-PEG-C-pept and Chol_PEG immobilized on the Si-APTES-GA substrate at a fixed concentration. All data reported are presented as the mean ± standard deviation (<span class="html-italic">n</span> ≥ 3).</p>
Full article ">Figure 4
<p>Specificity of the competitive bioassay of GNPs-Chol interacting with C-pept (blue), bare surface (green), and Pept 4 as negative control (red) in the presence of increasing concentrations of free cholesterol in PBST solution. All experiments were conducted under the same conditions (<span class="html-italic">n</span> ≥ 3).</p>
Full article ">Scheme 1
<p>Schematic representation of surface modification steps involved in the construction of the cholesterol biosensor. (<b>a</b>) Surface activation and APTES deposition via CVD, (<b>b</b>) incorporation of GA crosslinker and (<b>c</b>) immobilization of the biorecognition element C-pept.</p>
Full article ">Scheme 2
<p>Synthetic route of the surface functionalization of GNPs with Cholesterol_PEG_NH<sub>2</sub>.</p>
Full article ">Scheme 3
<p>Schematic representation of the competitive assay proposed in this work. The presence of free cholesterol in solution competes with cholesterol-functionalized GNPs for binding to C-pept, the cholesterol sensing element, immobilized on the sensor surface.</p>
Full article ">
Back to TopTop