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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,887)

Search Parameters:
Keywords = microsphere

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
11 pages, 702 KiB  
Article
Stretching the Limits of Refractometric Sensing in Water Using Whispering-Gallery-Mode Resonators
by Kevin Soler-Carracedo, Antonia Ruiz, Susana Ríos, Sergio de Armas-Rillo, Leopoldo L. Martín, Martin Hohmann, Inocencio R. Martín and Fernando Lahoz
Chemosensors 2025, 13(2), 33; https://doi.org/10.3390/chemosensors13020033 - 24 Jan 2025
Viewed by 349
Abstract
A novel application of microresonators for refractometric sensing in aqueous media is presented. To carry out this approach, microspheres of different materials and sizes were fabricated and doped with Nd3+ ions. Under 532 nm excitation, the microspheres presented typical NIR Nd3+ [...] Read more.
A novel application of microresonators for refractometric sensing in aqueous media is presented. To carry out this approach, microspheres of different materials and sizes were fabricated and doped with Nd3+ ions. Under 532 nm excitation, the microspheres presented typical NIR Nd3+ emission bands with superimposed sharp peaks, related to the Whispering Gallery Modes (WGMs), due to the geometry of the microspheres. When the microspheres were submerged in water with increasing concentrations of glycerol, spectral shifts for the WGMs were observed as a function of the glycerol concentration. These spectral shifts were studied and calibrated for three different microspheres and validated with the theoretical shifts, obtained by solving the Helmholtz equations for the electromagnetic field, considering the geometry of the system, and also by calculating the extinction cross-section. WGM shifts strongly depend on the diameter of the microspheres and their refractive index (RI) difference compared with the external medium, and are greater for decreasing values of the diameter and lower values of RI difference. Experimental sensitivities ranging from 2.18 to 113.36 nm/RIU (refractive index unit) were obtained for different microspheres. Furthermore, reproducibility measurements were carried out, leading to a repeatability of 2.3 pm and a limit of detection of 5 × 10−4 RIU. The proposed sensors, taking advantage of confocal microscopy for excitation and detection, offer a robust, reliable, and contactless alternative for environmental water analysis. Full article
13 pages, 12021 KiB  
Article
Production of Monodisperse Oil-in-Water Droplets and Polymeric Microspheres Below 20 μm Using a PDMS-Based Step Emulsification Device
by Naotomo Tottori, Seungman Choi and Takasi Nisisako
Micromachines 2025, 16(2), 132; https://doi.org/10.3390/mi16020132 - 24 Jan 2025
Viewed by 355
Abstract
Step emulsification (SE) is renowned for its robustness in generating monodisperse emulsion droplets at arrayed nozzles. However, few studies have explored poly(dimethylsiloxane) (PDMS)-based SE devices for producing monodisperse oil-in-water (O/W) droplets and polymeric microspheres with diameters below 20 µm—materials with broad applicability. In [...] Read more.
Step emulsification (SE) is renowned for its robustness in generating monodisperse emulsion droplets at arrayed nozzles. However, few studies have explored poly(dimethylsiloxane) (PDMS)-based SE devices for producing monodisperse oil-in-water (O/W) droplets and polymeric microspheres with diameters below 20 µm—materials with broad applicability. In this study, we present a PDMS-based microfluidic SE device designed to achieve this goal. Two devices with 264 nozzles each were fabricated, featuring straight and triangular nozzle configurations, both with a height of 4 µm and a minimum width of 10 µm. The devices were rendered hydrophilic via oxygen plasma treatment. A photocurable acrylate monomer served as the dispersed phase, while an aqueous polyvinyl alcohol solution acted as the continuous phase. The straight nozzles produced polydisperse droplets with diameters exceeding 30 µm and coefficient-of-variation (CV) values above 10%. In contrast, the triangular nozzles, with an opening width of 38 µm, consistently generated monodisperse droplets with diameters below 20 µm, CVs below 4%, and a maximum throughput of 0.5 mL h−1. Off-chip photopolymerization of these droplets yielded monodisperse acrylic microspheres. The low-cost, disposable, and scalable PDMS-based SE device offers significant potential for applications spanning from laboratory-scale research to industrial-scale particle manufacturing. Full article
(This article belongs to the Special Issue Recent Advances in Droplet Microfluidics)
Show Figures

Figure 1

Figure 1
<p>Polydimethylsiloxane (PDMS)-based step-emulsification (SE) devices for generating oil-in-water (O/W) droplets with diameters below 20 µm. (<b>a</b>) Schematic representation of the overall channel layout, showing: (1) a central channel for introducing the dispersed oil phase, (2) two side channels for supplying the continuous aqueous phase and collecting the produced droplets, and (3) two arrays of 132 shallow SE nozzles (264 nozzles in total). (<b>b</b>) Schematic cross-sectional view of the nozzle and channels, highlighting their respective heights. (<b>c</b>) Top-view illustrations of nozzle configurations, depicting straight nozzles with an upstream plateau (left) and triangular nozzles (right) along with their geometric parameters.</p>
Full article ">Figure 2
<p>Scanning electron microscopy (SEM) images of (<b>a</b>) master molds and (<b>b</b>) microchannels replicated in PDMS chips for the SE devices with straight nozzles (<b>left</b>) and triangular nozzles (<b>right</b>). Scale bar: 200 μm.</p>
Full article ">Figure 3
<p>Generation of polydisperse O/W droplets in the straight-nozzle device. (<b>a</b>) Droplet formation at a dispersed phase flow rate (<span class="html-italic">Q</span><sub>d</sub>) of 0.1 mL h<sup>−1</sup> and a continuous phase flow rate (<span class="html-italic">Q</span><sub>c</sub>) of 1.0 mL h<sup>−1</sup>. (<b>b</b>) Droplet formation under the same <span class="html-italic">Q</span><sub>d</sub> but at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup>. Arrows indicate the continuous phase flow direction. Scale bars: 100 μm.</p>
Full article ">Figure 4
<p>O/W droplet generation using the device with triangular nozzles. (<b>a</b>) Optical micrograph of the nozzle arrays operating at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>d</sub> = 0.1 mL h<sup>−1</sup>. Scale bar: 200 µm. (<b>b</b>) Magnified views of the nozzles in (<b>a</b>), demonstrating operation in the ‘small drop’ (SD) mode. (<b>c</b>) Magnified view of the nozzles operating at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>d</sub> = 0.7 mL h<sup>−1</sup>, showing two nozzles on the left operating in the SD mode, while the remaining nozzles operate in the ‘large drop’ (LD) mode. Scale bars: 20 µm.</p>
Full article ">Figure 5
<p>Micrographs and size distributions of O/W droplets collected from the device with triangular nozzles. Droplets were produced at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>d</sub> = (<b>a</b>) 0.1 mL h<sup>−1</sup>, (<b>b</b>) 0.7 mL h<sup>−1</sup>. Scale bars: 20 μm.</p>
Full article ">Figure 6
<p>Evolution of the average droplet diameter (<span class="html-italic">D</span><sub>avg</sub>) and CV values across the SD and LD regimes for O/W droplets generated using the device with triangular nozzles, with <span class="html-italic">Q</span><sub>d</sub> varied from 0.1 to 1.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>c</sub> fixed at 5.0 mL h<sup>−1</sup>.</p>
Full article ">Figure 7
<p>Monodisperse polymeric microspheres obtained through off-chip photopolymerization. (<b>a</b>) SEM images of polymeric microspheres derived from O/W droplets generated by the device with edge-shaped nozzles at <span class="html-italic">Q</span><sub>c</sub> = 5.0 mL h<sup>−1</sup> and <span class="html-italic">Q</span><sub>d</sub> = 0.1 mL h<sup>−1</sup>. (<b>b</b>) Size distribution of the microspheres. Scale bar: 10 µm.</p>
Full article ">
16 pages, 5756 KiB  
Article
Performance Evaluation of Fluorescent Polymer Gel Microspheres as a Reservoir Conformance Control Agent
by Saya Shagymgereyeva, Bauyrzhan Sarsenbekuly, Wanli Kang and Sarsenbek Turtabayev
Gels 2025, 11(2), 85; https://doi.org/10.3390/gels11020085 - 22 Jan 2025
Viewed by 528
Abstract
This study introduces fluorescent polymer gel microspheres (FPMs) as a novel approach to enhance conformance control in oil reservoirs. Designed to address the challenges of high-permeability zones, FPMs were synthesized via inverse suspension polymerization, incorporating 2-acrylamido-2-methylpropane sulfonic acid (AMPS) to improve thermal stability [...] Read more.
This study introduces fluorescent polymer gel microspheres (FPMs) as a novel approach to enhance conformance control in oil reservoirs. Designed to address the challenges of high-permeability zones, FPMs were synthesized via inverse suspension polymerization, incorporating 2-acrylamido-2-methylpropane sulfonic acid (AMPS) to improve thermal stability and swelling and fluorescein to enable fluorescence. Characterization using FT-IR, SEM, fluorescence spectroscopy, and thermal analysis revealed that FPMs swell significantly in brine, with diameters increasing from 46 μm to 210 μm, and maintain thermal stability up to 110 °C. These advanced properties make FPMs highly effective in reducing permeability and facilitating real-time tracking, offering a promising solution for improved oil recovery and efficient reservoir management. Full article
(This article belongs to the Special Issue Chemical and Gels for Oil Drilling and Enhanced Recovery)
Show Figures

Figure 1

Figure 1
<p>The FTIR spectrum of terpolymer microspheres.</p>
Full article ">Figure 2
<p>Particle size of FPMs: (<b>A</b>) SEM images, (<b>B</b>) optical microscope images.</p>
Full article ">Figure 3
<p>Morphology of FPMs: (<b>A</b>) under the ordinary light, (<b>B</b>) under the UV light.</p>
Full article ">Figure 4
<p>SEM images of terpolymer microspheres: (<b>A</b>) before swelling, (<b>B</b>) after swelling.</p>
Full article ">Figure 5
<p>SEM images of terpolymer microspheres: (<b>A</b>) the 8th day, (<b>B</b>) the 14th day 110 °C temperature exposure.</p>
Full article ">Figure 6
<p>SEM images of terpolymer microspheres: (<b>A</b>) before temperature exposure, (<b>B</b>) the 15th day 110 °C temperature exposure.</p>
Full article ">Figure 7
<p>The TG curves of FPM.</p>
Full article ">Figure 8
<p>Schematic diagram of FPM synthesis.</p>
Full article ">Figure 9
<p>Synthesis scheme of (<b>A</b>). P(AM-AMPS-NVP), (<b>B</b>). P(AM-NVP-Flu), (<b>C</b>). P(AM-AMPS-NVP).</p>
Full article ">Figure 9 Cont.
<p>Synthesis scheme of (<b>A</b>). P(AM-AMPS-NVP), (<b>B</b>). P(AM-NVP-Flu), (<b>C</b>). P(AM-AMPS-NVP).</p>
Full article ">
14 pages, 6740 KiB  
Article
Facile Preparation of Flexible Phenolic-Silicone Aerogels with Good Thermal Stability and Fire Resistance
by Zengyue Su, Zhenrong Zheng, Xiaobiao Zuo, Lijuan Luo and Yaxin Guo
Molecules 2025, 30(3), 464; https://doi.org/10.3390/molecules30030464 - 21 Jan 2025
Viewed by 392
Abstract
A huge challenge is how to prepare flexible silicone aerogel materials with good flame retardancy, thermal stability, and hydrophobic properties. In this paper, resorcinol–formaldehyde was introduced into the silicone network composed of methyltrimethoxysilane (MTMS), phenyltriethoxysilane (PTES), and dimethyldimethoxysilane (DMDMS). Flexible hybrid aerogels with [...] Read more.
A huge challenge is how to prepare flexible silicone aerogel materials with good flame retardancy, thermal stability, and hydrophobic properties. In this paper, resorcinol–formaldehyde was introduced into the silicone network composed of methyltrimethoxysilane (MTMS), phenyltriethoxysilane (PTES), and dimethyldimethoxysilane (DMDMS). Flexible hybrid aerogels with excellent thermal insulation, flame retardant, and hydrophobic properties were prepared by the sol–gel method and ambient pressure drying (APD), and the preparation process does not require long-term solvent exchange, only about 3 h of soaking and washing of the wet gel. The results show that the prepared phenolic-silicone aerogel has low density (0.093 g/cm3), low thermal conductivity (0.041 W/m·K), high flexibility, and compression fatigue resistance. The phenolic microspheres are bonded to the silicone skeleton to maintain the original flexibility. After 50% compression deformation, it returns to the original size normally, and there is no significant change in the stress of the sample after 50 compression cycles. Compared with pure silicone aerogels, the hybrid aerogels doped with phenolic have better char yield (65.28%) and higher decomposition temperature (609 °C). The hybrid aerogel sample has good flame-retardant properties, which can withstand alcohol lamp burning without being ignited. The micron-sized phenolic beads give the hybrid aerogels better hydrophobic properties, showing a higher static water contact angle (152°). The excellent thermal and mechanical properties mean that the hybrid aerogels prepared in this paper have good application prospects for aerospace, outdoor equipment, and other fields. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Polymerization reaction of phenol and aldehyde in alkaline environment. (<b>b</b>) Hydrolysis of siloxane in acidic environment. (<b>c</b>) Drying the sol after fully stirring.</p>
Full article ">Figure 2
<p>(<b>a</b>) Changes from phenolic aqueous solution to emulsion. (<b>b</b>) Macroscopic morphology of aerogels with different molar additions. (<b>c</b>) The cross sections of samples S1–S5.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>e</b>) SEM of S1-S5. (<b>f</b>,<b>g</b>) SEM and EDS of sample S3. (<b>h</b>,<b>i</b>) The particle size of silicon microspheres and phenolic microspheres.</p>
Full article ">Figure 4
<p>(<b>a</b>−<b>c</b>) C1s, Si2p, and O high-resolution XPS spectra of S3. (<b>d</b>) FTIR spectra of S3.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>e</b>) TGA and DTGA curves of S1–S5 under nitrogen atmosphere. (<b>f</b>) Thermal conductivity and bulk density of sample S1–S5.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>e</b>) The infrared temperature diagram of samples S1–S5 heated in a heating plate for 10 s. (<b>f</b>–<b>j</b>) The infrared temperature diagram of samples S1–S5 heated in a heating plate for 300 s.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>c</b>) Flexibility test of S1 samples after 280 °C heating plate test. (<b>d</b>–<b>f</b>) Flexibility test of S3 samples after 280 °C heating plate test. (<b>g</b>–<b>k</b>) the macroscopic morphology of sample S1–S5 after heating plate test at 600 °C for 30 s.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>c</b>) Photographs of S1 aerogels burned with alcohol lamp. (<b>d</b>) Photograph of sample S1 after burning. (<b>e</b>–<b>g</b>) Photographs of S3 aerogels burned with alcohol lamp. (<b>h</b>) Photograph of sample S3 after burning.</p>
Full article ">Figure 9
<p>Nitrogen adsorption–desorption hysteresis curve and static contact angle.</p>
Full article ">Figure 10
<p>(<b>a</b>) Sample S3 finger compression rebound test. (<b>b</b>) Stress comparison of samples S1–S5 at 50% strain degree. (<b>c</b>–<b>g</b>) Stress–strain comparison of samples S1–S5 at 1, 10, and 50 cycles.</p>
Full article ">
12 pages, 2231 KiB  
Article
A Time-Resolved Fluorescent Microsphere Immunochromatographic Assay for Determination of Vitamin B12 in Infant Formula Milk Powder
by Qianqian Lu, Yongwei Feng, Qi Zhou, Ting Yang, Hua Kuang, Chuanlai Xu and Lingling Guo
Biosensors 2025, 15(2), 65; https://doi.org/10.3390/bios15020065 - 21 Jan 2025
Viewed by 391
Abstract
Vitamin B12 (VB12) is an important nutrient, and its quality control in food is crucial. In this study, based on the principle of specific recognition of target analyte by monoclonal antibodies (mAbs), a time-resolved fluorescent microsphere immunochromatographic assay (TRFM-ICA) was developed to detect [...] Read more.
Vitamin B12 (VB12) is an important nutrient, and its quality control in food is crucial. In this study, based on the principle of specific recognition of target analyte by monoclonal antibodies (mAbs), a time-resolved fluorescent microsphere immunochromatographic assay (TRFM-ICA) was developed to detect the content of VB12 in infant formula milk powder. First, the performance of the anti-VB12 mAb was evaluated, revealing a half-maximal inhibitory concentration of 0.370 ng/mL, an affinity constant of 2.604 × 109 L/mol and no cross-reactivity with other vitamins. Then, a highly sensitive TRFM-ICA was developed, with a visual limit of detection of 10 μg/kg and a cut-off value of 100 μg/kg for qualitative detection and a detection range of 4.125–82.397 μg/kg for quantitative detection. In addition, the test results of real samples were consistent with the results of quantification using microbiological methods, with a coefficient of variation of less than 10%, showing good accuracy and stability, and confirming that the TRFM-ICA is suitable for the analysis of VB12 in real infant formula milk powder samples. In this study, based on the principle of specific recognition of VB12 by monoclonal antibodies (mAbs) against VB12, a time-resolved fluorescence microsphere immunochromatographic assay (TRFM-ICA) was developed to detect the content of VB12 in infant formula by converting biological signals into optical signals. Full article
(This article belongs to the Special Issue Feature Papers of Biosensors)
Show Figures

Figure 1

Figure 1
<p>Principles (<b>a</b>) and test results (<b>b</b>) of TRFM-ICA.</p>
Full article ">Figure 2
<p>Performance testing of the anti-VB<sub>12</sub> mAb. (<b>a</b>) Isotype measurement. (<b>b</b>) Affinity measurement. (<b>c</b>) Sensitivity optimization. (<b>d</b>) Sensitivity measurement.</p>
Full article ">Figure 3
<p>Optimization of TRFM-ICA. (<b>a</b>) Optimization of coating antigen concentrations in T line. (<b>b</b>) Optimization of dilution ratio of microspheres. (<b>c</b>) Optimization of surfactant.</p>
Full article ">Figure 4
<p>Sensitivity measurement of TRFM-ICA. (<b>a</b>) Images of sensitivity testing. (<b>b</b>) Standard curve of VB<sub>12</sub> detection.</p>
Full article ">Figure 5
<p>Test results of TRFM-ICA in infant formula milk powder. (<b>a</b>) Optimization of dilution ratio of microspheres. (<b>b</b>) Standard curves for different dilution ratios of sample extraction solutions. (<b>c</b>) Image of measurement result of 6-fold dilution of sample extraction solution.</p>
Full article ">
20 pages, 4307 KiB  
Article
Preparation and Biochemical Characterization of Penicillium crustosum Thom P22 Lipase Immobilization Using Adsorption, Encapsulation, and Adsorption–Encapsulation Approaches
by Ismail Hasnaoui, Sondes Mechri, Ahlem Dab, Nour Eddine Bentouhami, Houssam Abouloifa, Reda Bellaouchi, Fawzi Allala, Ennouamane Saalaoui, Bassem Jaouadi, Alexandre Noiriel, Abdeslam Asehraou and Abdelkarim Abousalham
Molecules 2025, 30(3), 434; https://doi.org/10.3390/molecules30030434 - 21 Jan 2025
Viewed by 532
Abstract
This work describes the immobilization and the characterization of purified Penicillium crustosum Thom P22 lipase (PCrL) using adsorption, encapsulation, and adsorption–encapsulation approaches. The maximum activity of the immobilized PCrL on CaCO3 microspheres and sodium alginate beads was shifted from 37 to 45 [...] Read more.
This work describes the immobilization and the characterization of purified Penicillium crustosum Thom P22 lipase (PCrL) using adsorption, encapsulation, and adsorption–encapsulation approaches. The maximum activity of the immobilized PCrL on CaCO3 microspheres and sodium alginate beads was shifted from 37 to 45 °C, compared with that of the free enzyme. When sodium alginate was coupled with zeolite or chitosan, the immobilization yield reached 100% and the immobilized PCrL showed improved stability over a wide temperature range, retaining all of its initial activity after a one-hour incubation at 60 °C. The immobilization of PCrL significantly improves its catalytic performance in organic solvents, its pH tolerance value, and its thermal stability. Interestingly, 95% and almost 50% of PCrL’s initial activity was retained after 6 and 12 cycles, respectively. The characteristics of all PCrL forms were analyzed by X-ray diffraction and scanning electron microscopy combined with energy dispersive spectroscopy. The maximum conversion efficiency of oleic acid and methanol to methyl esters (biodiesel), by PCrL immobilized on CaCO3, was 65% after a 12 h incubation at 40 °C, while free PCrL generated only 30% conversion, under the same conditions. Full article
(This article belongs to the Section Macromolecular Chemistry)
Show Figures

Figure 1

Figure 1
<p>Adsorption kinetics of PCrL on CaCO<sub>3</sub> and Celite 545. PCrL adsorbed on CaCO<sub>3</sub> leads to the highest yield of 90% after 30 min of incubation (4500 U) at 4 °C. The lipase activity was measured with the pH-STAT technique using TC8 as the substrate.</p>
Full article ">Figure 2
<p>Identification of the PCrL-SA-ZE, PCrL-SA-CS, and PCrL-CaCO<sub>3</sub> beads. (<b>a</b>) The XRD patterns of the support beads alone (SA, SA-ZE, or SA-CS) or complexed with PCrL (PCrL-SA, PCrL-SA-ZE, and PCrL-SA-CS). The XRD patterns were generated using Match! software (version 3.10.2.173); (<b>b</b>) Identification of the support beads (SA, SA-CS, and SA-ZE) and those complexed with PCrL (PCrL-SA, PCrL-SA-CS, and PCrL-SA-ZE) in the FE-SEM images.</p>
Full article ">Figure 3
<p>The effect of temperature on PCrL activity and stability. (<b>a</b>) The temperature–activity profile. This graph shows the activity of the free and immobilized PCrL at different temperatures. The activity of the immobilized PCrL at its optimal temperature (45 °C) is set at 100%. (<b>b</b>) Temperature stability. This graph illustrates the stability of the free and immobilized PCrL after incubation at different temperatures for 60 min. The residual activity was measured at pH value 9 using TC8 as a substrate. Each data point on the graph represents the average of three independent experiments.</p>
Full article ">Figure 4
<p>The effect of pH on PCrL activity and stability. (<b>a</b>) The pH activity profile of the free and immobilized PCrL at different pH values. The maximum activity at pH value 9 is set at 100%. (<b>b</b>) pH value stability of the free and immobilized PCrL after incubation at different pH values for 1 h at 4 °C. The residual activity was measured at pH value 9 and 37 °C using TC8 as the substrate. Each data point represents the average of three independent experiments.</p>
Full article ">Figure 5
<p>The effect of organic solvents on PCrL-CaCO<sub>3</sub> activity and stability. The enzyme was incubated with 25% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) of each solvent for 24 h. The residual activity was measured under standard conditions, using TC8 as the substrate at 37 °C and pH 9, as described in Material and Methods, and then expressed as a percentage of the activity without any solvents. Each data point represents the average of three independent experiments, with the error bars indicating standard deviation.</p>
Full article ">Figure 6
<p>Performance evaluation of PCrL-CaCO<sub>3</sub>. (<b>a</b>) The kinetics of oleic acid esterification catalyzed by the free PCrL and immobilized PCrL-CaCO<sub>3</sub>. The reaction was performed at 40 °C, with stirring for 24 h, using 500 U of enzyme in hexane with a 3:1 molar ratio of methanol to oleic acid. (<b>b</b>) The reusability of immobilized PCrL-CaCO<sub>3</sub> in multiple reaction cycles. The enzyme was reused for 12 cycles, with each cycle lasting 12 h. The conversion yield of oleic acid to esters was monitored for each cycle.</p>
Full article ">Figure 7
<p>The thermodynamic parameters of the target PCrL. An Arrhenius diagram of Ln (<span class="html-italic">k</span><sub>d</sub>) vs. 1/temperature to calculate the activation energy (E<sub>a</sub>).</p>
Full article ">
14 pages, 6246 KiB  
Article
Design and Preparation of ZnIn2S4/g-C3N4 Z-Scheme Heterojunction for Enhanced Photocatalytic CO2 Reduction
by Jinghong Fang, Min Wang, Xiaotong Yang, Qiong Sun and Liyan Yu
Catalysts 2025, 15(1), 95; https://doi.org/10.3390/catal15010095 - 20 Jan 2025
Viewed by 760
Abstract
In this study, a novel Z-scheme heterojunction photocatalyst was developed by integrating g-C3N4 nanoplates into ZnIn2S4 microspheres. X-ray photoelectron spectroscopy analysis revealed a directional electron transfer from g-C3N4 to ZnIn2S4 upon [...] Read more.
In this study, a novel Z-scheme heterojunction photocatalyst was developed by integrating g-C3N4 nanoplates into ZnIn2S4 microspheres. X-ray photoelectron spectroscopy analysis revealed a directional electron transfer from g-C3N4 to ZnIn2S4 upon heterojunction formation. Under irradiation, electrochemical tests and electron paramagnetic resonance spectroscopy demonstrated significantly enhanced charge generation and separation efficiencies in the ZnIn2S4/g-C3N4 composite, accompanied by reduced charge transfer resistance. In photocatalytic CO2 reduction, the ZnIn2S4/g-C3N4 composite achieved the highest CO yield, 1.92 and 5.83 times higher than those of pristine g-C3N4 and ZnIn2S4, respectively, with a notable CO selectivity of 91.3% compared to H2 (8.7%). The Z-scheme heterojunction mechanism, confirmed in this work, effectively preserved the strong redox capabilities of the photoinduced charge carriers, leading to superior photocatalytic performance and excellent long-term stability. This study offers valuable insights into the design and development of g-C3N4-based heterojunctions for efficient solar-driven CO2 reduction. Full article
(This article belongs to the Special Issue Functional Nanomaterials in Catalysis and Sensing)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of g-C<sub>3</sub>N<sub>4</sub>, ZnIn<sub>2</sub>S<sub>4</sub> and 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>.</p>
Full article ">Figure 2
<p>FESEM images of (<b>a</b>) ZnIn<sub>2</sub>S<sub>4</sub> and (<b>b</b>) 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>, along with (<b>c</b>–<b>h</b>) the corresponding elemental mappings of Zn, In, S, C, and N in 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>.</p>
Full article ">Figure 3
<p>XPS spectra of g-C<sub>3</sub>N<sub>4</sub>, ZnIn<sub>2</sub>S<sub>4</sub>, and 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>: (<b>a</b>) survey spectra, (<b>b</b>) C 1s, (<b>c</b>) N 1s, (<b>d</b>) Zn 2p, (<b>e</b>) In 3d, and (<b>f</b>) S 2p spectra.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV-vis absorption spectra, (<b>b</b>) Tauc plots, and (<b>c</b>) PL spectra of g-C<sub>3</sub>N<sub>4</sub>, ZnIn<sub>2</sub>S<sub>4</sub>, and 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>.</p>
Full article ">Figure 5
<p>Mott–Schottky plots of (<b>a</b>) g-C<sub>3</sub>N<sub>4</sub>, (<b>b</b>) ZnIn<sub>2</sub>S<sub>4</sub>, and (<b>c</b>) 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub> measured at different frequencies.</p>
Full article ">Figure 6
<p>(<b>a</b>) Transient photocurrent response curves of g-C<sub>3</sub>N<sub>4</sub>, ZnIn<sub>2</sub>S<sub>4</sub>, and 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub> under light on/off cycles. (<b>b</b>) Nyquist plots of g-C<sub>3</sub>N<sub>4</sub>, ZnIn<sub>2</sub>S<sub>4</sub>, and 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>.</p>
Full article ">Figure 7
<p>CO<sub>2</sub> reduction performance under various conditions: (1) photocatalytic CO<sub>2</sub> reduction using a 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub> composite under light irradiation; (2) reaction with light and CO<sub>2</sub> but without the catalyst; (3) reaction without light, but with the catalyst and CO<sub>2</sub>; (4) reaction with light and the catalyst, where CO<sub>2</sub> is replaced by Ar.</p>
Full article ">Figure 8
<p>(<b>a</b>) Photocatalytic performance of the catalyst with varying mass ratios of ZnIn<sub>2</sub>S<sub>4</sub> to g-C<sub>3</sub>N<sub>4</sub>; (<b>b</b>) time-dependent gas yield for the 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>; (<b>c</b>) photocatalytic performance of the composite materials at different dosages; (<b>d</b>) results of the cyclic stability experiments.</p>
Full article ">Figure 9
<p>TEMPO spin-trapping EPR spectra of g-C<sub>3</sub>N<sub>4</sub> and 1:2 ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>, showing photoinduced electron signals recorded in water.</p>
Full article ">Figure 10
<p>Proposed charge transfer mechanism for the efficient photocatalytic CO<sub>2</sub> reduction process facilitated by the ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub> heterojunction.</p>
Full article ">Scheme 1
<p>The preparation process of ZnIn<sub>2</sub>S<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>.</p>
Full article ">Scheme 2
<p>Detailed flow chart of the photocatalytic CO<sub>2</sub> reduction process.</p>
Full article ">
18 pages, 5599 KiB  
Article
The Essential Role of Monte Carlo Simulations for Lung Dosimetry in Liver Radioembolization—Part B: 166Ho Microspheres
by Edoardo d’Andrea, Andrea Politano, Bartolomeo Cassano, Nico Lanconelli, Marta Cremonesi, Vincenzo Patera and Massimiliano Pacilio
Appl. Sci. 2025, 15(2), 958; https://doi.org/10.3390/app15020958 - 19 Jan 2025
Viewed by 569
Abstract
This study compares dosimetric approaches for lung dosimetry in 166 radioembolization (Ho-TARE) with direct Monte Carlo (MC) simulations on a voxelized anthropomorphic phantom derived from a real patient’s CT scan, preserving the patient’s lung density distribution. Lung dosimetry was assessed for five lung [...] Read more.
This study compares dosimetric approaches for lung dosimetry in 166 radioembolization (Ho-TARE) with direct Monte Carlo (MC) simulations on a voxelized anthropomorphic phantom derived from a real patient’s CT scan, preserving the patient’s lung density distribution. Lung dosimetry was assessed for five lung shunt (LS) scenarios with conventional methods: the mono-compartmental organ-level approach (MIRD), voxel S-value convolution for soft tissue (kST, ICRU soft tissue with 1.04 g/cm3) and lung tissue (kLT, ICRU lung tissue with 0.296 g/cm3), local density rescaling (kSTL and kLTL, respectively, for soft tissue and lung tissue), or global rescaling for a lung mean density of 0.221 g/cm3 (kLT221). Significant underestimations in the mean absorbed dose (AD) were observed, with relative differences with respect to the reference (MC) of −64% for MIRD, −93% for kST, −56% for kSTL, −76% for kLT, −68% for kLT221, and −60% for kLTL. Given the high heterogeneity of lung tissue, standard dosimetric approaches cannot accurately estimate the AD. Additionally, MC results for 166Ho showed notable spatial absorbed dose inhomogeneity, highlighting the need for tailored lung dosimetry in Ho-TARE accounting for the patient-specific lung density distribution. MC-based dosimetry thus proves to be essential for safe and effective radioembolization treatment planning in the presence of LS. Full article
Show Figures

Figure 1

Figure 1
<p>VSV kernels (available as Supporting Material of this paper) represented as a plot of the AD per unit decay to the target voxel (y-axis) versus the source–target voxel distance (x-axis) for the soft tissue (<b>a</b>) and lung tissue (<b>b</b>) for <sup>166</sup>Ho on a square voxel of 2.21 mm side.</p>
Full article ">Figure 2
<p>Correlation plot of the mean absorbed dose in the lungs (<math display="inline"><semantics> <mover> <mrow> <mi>A</mi> <mi>D</mi> </mrow> <mo>¯</mo> </mover> </semantics></math>) per GBq of administered activity, obtained from MC simulations with the reference phantom (x-axis), compared with those obtained using the methods listed in <a href="#applsci-15-00958-t002" class="html-table">Table 2</a> (y-axis). Each point in the data series represents an increasing LS value (10%, 20%, 30%, and 40%), with a line representing the linear interpolation of each dataset, provided as a qualitative visual guide only.</p>
Full article ">Figure 3
<p>Example slices in the coronal view of the AD spatial distributions (left) for <math display="inline"><semantics> <msub> <mi>kLT</mi> <mi>L</mi> </msub> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <msub> <mi>kST</mi> <mi>L</mi> </msub> </semantics></math> (<b>b</b>), and MC (<b>c</b>), along with their respective color scales, are shown, whereas the plot (right) reports the corresponding DVH (<b>d</b>). All dosimetric approaches show significant heterogeneity in the AD spatial distribution. The values of the AD maps and the DVH are given in Gy per GBq of administered activity.</p>
Full article ">Figure 4
<p>Example slices in the coronal view of the ADr maps (left) of <sup>166</sup>Ho (<b>a</b>) and <sup>90</sup>Y (<b>b</b>), along with the corresponding DrVHs (<b>c</b>), are shown for LS = 10%. The data demonstrate a different degree of inhomogeneity between the two radionuclides, due to the distinct physical characteristics of their decay spectra.</p>
Full article ">Figure 5
<p>Cumulative DVHs of <math display="inline"><semantics> <mrow> <mi>E</mi> <mi>Q</mi> <msub> <mi>D</mi> <mn>2</mn> </msub> </mrow> </semantics></math> for <sup>90</sup>Y (blue band) and <sup>166</sup>Ho (yellow band) for the 0.5–1 h range of <math display="inline"><semantics> <msub> <mi mathvariant="normal">T</mi> <mi>μ</mi> </msub> </semantics></math> along with the volumetric constraints (red dots) listed in <a href="#applsci-15-00958-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 6
<p><math display="inline"><semantics> <mover> <mrow> <mi>A</mi> <mi>D</mi> </mrow> <mo>¯</mo> </mover> </semantics></math> NTCP model for RP incidence in partial lung irradiation treatments from EBRT as reported in QUANTEC [<a href="#B42-applsci-15-00958" class="html-bibr">42</a>]. The black solid line is the logistic model according to Equation (<a href="#FD6-applsci-15-00958" class="html-disp-formula">6</a>) with parameters <math display="inline"><semantics> <mrow> <msub> <mi>b</mi> <mn>0</mn> </msub> <mo>=</mo> <mo>−</mo> <mn>3.87</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>b</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.126</mn> <mspace width="4.pt"/> <msup> <mi>Gy</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>, the blue dots refer to <sup>90</sup>Y cases, and the yellow ones to <sup>166</sup>Ho cases, each for the LS = 10% case and for the labeled <math display="inline"><semantics> <msub> <mi mathvariant="normal">T</mi> <mi>μ</mi> </msub> </semantics></math>.</p>
Full article ">
29 pages, 6394 KiB  
Review
Preparation of Colored Polymer Microspheres
by Lei Wang, Weiting Ma, Shuheng Zhang, Mengke He, Ping Song, Hongying Wang, Xianxiao Song and Botian Li
Molecules 2025, 30(2), 375; https://doi.org/10.3390/molecules30020375 - 17 Jan 2025
Viewed by 393
Abstract
Colored polymer microspheres have attracted significant attention in both academia and industry due to their unique optical properties and extensive application potential. However, achieving a uniform distribution of dyes within these microspheres remains a challenge, particularly when heavy concentrations of dye are used, [...] Read more.
Colored polymer microspheres have attracted significant attention in both academia and industry due to their unique optical properties and extensive application potential. However, achieving a uniform distribution of dyes within these microspheres remains a challenge, particularly when heavy concentrations of dye are used, as this can lead to aggregation or delamination, adversely affecting their application. Additionally, many dyes are prone to degradation or fading when exposed to light, heat, or chemicals, which compromises the long-term color stability of the microspheres. Consequently, the preparation of colored polymer microspheres with high stability continues to be a significant challenge. This review offers a comprehensive overview of the preparation techniques for colored polymer microspheres and their dyeing mechanisms, introducing the fundamental concepts of these microspheres and their applications in various fields, such as biomedicine, optical devices, and electronic display technologies. It further presents a detailed discussion of the different preparation methods, including physical adsorption, chemical bonding, and copolymerization. The advantages, limitations, and potential improvements of each method are explored, along with an analysis of the interactions between dyes and the polymer matrix, and how these interactions influence the properties of the microspheres, including their color uniformity, stability, and durability. Finally, the review discusses future perspectives on the development of colored polymer microspheres, highlighting the advancement of novel materials, innovations in preparation technology, and the exploration of potential new application areas. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Applications of colored polymer microspheres [<a href="#B2-molecules-30-00375" class="html-bibr">2</a>,<a href="#B3-molecules-30-00375" class="html-bibr">3</a>,<a href="#B4-molecules-30-00375" class="html-bibr">4</a>,<a href="#B5-molecules-30-00375" class="html-bibr">5</a>].</p>
Full article ">Figure 2
<p>Chemical structure of (<b>a</b>) methylene blue, (<b>b</b>) Rhodamine B acrylate [<a href="#B52-molecules-30-00375" class="html-bibr">52</a>], (<b>c</b>) ZnAOTPP [<a href="#B55-molecules-30-00375" class="html-bibr">55</a>], (<b>d</b>) Sudan III [<a href="#B56-molecules-30-00375" class="html-bibr">56</a>], (<b>e</b>) RB2 [<a href="#B60-molecules-30-00375" class="html-bibr">60</a>], and (<b>f</b>) pyrene acrylate [<a href="#B61-molecules-30-00375" class="html-bibr">61</a>].</p>
Full article ">Figure 3
<p>Synthesis routes of (<b>a</b>) P(AM-BA-AMCO), (<b>b</b>) P(AM-BA-Ac-Flu), and (<b>c</b>) P(AM-BA-RhB); (<b>d</b>) fluorescent images of the three polymer microspheres (A–C: Fluorescent images; A<sub>1</sub>–C<sub>1</sub>: Optical microscope images) [<a href="#B3-molecules-30-00375" class="html-bibr">3</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Synthesis routes and fluorescence properties of a fluorescent polymer with AIE units [<a href="#B59-molecules-30-00375" class="html-bibr">59</a>]; (<b>b</b>) synthesis routes of a fluorescent polymer with Rhodamine B groups and their fluorescence microscopy (b1: blue light, b2: green light) [<a href="#B53-molecules-30-00375" class="html-bibr">53</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Chemical structure of PPE; (<b>b</b>) synthesis route of APGMA-PPE-NR fluorescent microspheres; (<b>c</b>) encoding strategy for microspheres that comprises varying the concentrations of PPE and NR in THF solution, as well as their combination [<a href="#B2-molecules-30-00375" class="html-bibr">2</a>].</p>
Full article ">Figure 6
<p>Illustration of the preparation of APGMA-CP fluorescent microspheres and the bio-conjugation of BSA-FITC onto the microspheres [<a href="#B66-molecules-30-00375" class="html-bibr">66</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Synthesis route of polysulfide zinc (II); (<b>b</b>) synthesis route of compound 1 [<a href="#B75-molecules-30-00375" class="html-bibr">75</a>]; (<b>c</b>) illustration of crosslinked shell-coated fluorescent PPV microspheres; (<b>d</b>) chemical structures [<a href="#B76-molecules-30-00375" class="html-bibr">76</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) Preparation of fluorescent nanoparticles via the swelling method [<a href="#B58-molecules-30-00375" class="html-bibr">58</a>]; (<b>b</b>) illustration of quantum dots being embedded into PSDM microspheres [<a href="#B82-molecules-30-00375" class="html-bibr">82</a>]; (<b>c</b>) SB@PS bead preparation process [<a href="#B84-molecules-30-00375" class="html-bibr">84</a>]; (<b>d</b>) preparation process of QD-encoded polymer microspheres, including (I) swelling, (II) temperature increase, and (III) deswelling and rapidly cooling. a1–a2 Dissolution: QDs are mixed with polymer microspheres, which dissolve in chloroform due to hydrophobicity. a2–a3 warming impregnation: warming allows the microspheres to continue to swell and the QDs to penetrate into the microspheres due to the concentration difference. a3–a4 Encapsulation: Remove chloroform and cool quickly to immobilize the molecular chains and encapsulate the QDs [<a href="#B85-molecules-30-00375" class="html-bibr">85</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) Chemical structures of the organic dyes for adsorption; (<b>b</b>) equilibrium adsorption capacity of PDA microspheres for different dyes [<a href="#B91-molecules-30-00375" class="html-bibr">91</a>]; (<b>c</b>) synthesis of polymer microspheres; (<b>d</b>) UV–Vis absorption spectra of dyed microspheres with different amounts of VBC; (<b>e</b>) variation in the dye content incorporated into dyed microspheres [<a href="#B92-molecules-30-00375" class="html-bibr">92</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) Synthesis route of RB2, (<b>b</b>) absorption spectra of RB1 and RB2, and (<b>c</b>) emission spectra [<a href="#B60-molecules-30-00375" class="html-bibr">60</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Synthesis route of rosin-derived CCRM via the suspension polymerization strategy; (<b>b</b>) adsorption capacity of CCRM for different dyes [<a href="#B96-molecules-30-00375" class="html-bibr">96</a>]; (<b>c</b>) synthesis route of CMPSF; (<b>d</b>) synthesis route of porous hollow microspheres; (<b>e</b>) synthesis route of porous hollow carboxylated polysulfone microspheres; (<b>f</b>) purification performance; and (<b>g</b>) UV–Vis spectra of PH-PSF and PH-CPSF microspheres [<a href="#B97-molecules-30-00375" class="html-bibr">97</a>].</p>
Full article ">Figure 12
<p>Bright-field (<b>a</b>–<b>d</b>) and fluorescence (green channel) (<b>e</b>–<b>h</b>) microscope images of PS microspheres loaded with chlorophyll a at different dye loading levels: (<b>a</b>,<b>e</b>) 1 wt%, (<b>b</b>,<b>f</b>) 2.5 wt%, (<b>c</b>,<b>g</b>) 4 wt%, and (<b>d</b>,<b>h</b>) 10 wt% [<a href="#B99-molecules-30-00375" class="html-bibr">99</a>].</p>
Full article ">Figure 13
<p>(<b>a</b>) Effect of pH on MB adsorption using PAM/SA [<a href="#B96-molecules-30-00375" class="html-bibr">96</a>]; (<b>b</b>) effect of temperature on MB adsorption using PAM/SA [<a href="#B103-molecules-30-00375" class="html-bibr">103</a>]; (<b>c</b>) change in relative concentration of RhB in solution 1 under different irradiation times [<a href="#B108-molecules-30-00375" class="html-bibr">108</a>].</p>
Full article ">
15 pages, 1672 KiB  
Article
Colostrum-Derived Melatonin Plus PEG Microspheres Modulate the Oxidative Metabolism of Human Colostrum Phagocytes
by Caroline G. Silva, Viviane F. Luz, Victor L. Nunes, Ana B. M. Verzoto, Aron C. de M. Cotrim, Wagner B. dos Santos, Eduardo L. França and Adenilda C. Honorio-França
Metabolites 2025, 15(1), 57; https://doi.org/10.3390/metabo15010057 - 16 Jan 2025
Viewed by 428
Abstract
Background/Objectives: Exogenous melatonin adsorbed onto PEG microspheres can modulate the functional activity of phagocytes in colostrum, but no data are available on the activity of melatonin found in colostrum. Therefore, the objective of this study was to extract melatonin from human colostrum, develop [...] Read more.
Background/Objectives: Exogenous melatonin adsorbed onto PEG microspheres can modulate the functional activity of phagocytes in colostrum, but no data are available on the activity of melatonin found in colostrum. Therefore, the objective of this study was to extract melatonin from human colostrum, develop and characterize PEG microspheres with the extracted melatonin adsorbed onto them, and evaluate the effects of this system on the oxidative metabolism of colostrum phagocytes. Methods: Thirty colostrum samples were collected; ten were used for melatonin extraction, while twenty were used to obtain phagocytes. Melatonin was extracted from the colostrum supernatant through affinity chromatography and quantified by ELISA. The polyethylene glycol microspheres produced were analyzed using fluorescence microscopy and flow cytometry. Oxidative metabolism was assessed by measuring the release of the superoxide anion and superoxide enzymes. A control was conducted using commercial melatonin. Results: The fluorescence microscopy and flow cytometry analyses demonstrated that PEG microspheres can adsorb melatonin. There was an increase in superoxide release in phagocytes incubated with colostrum-derived or synthetic melatonin. When exposed to bacteria, colostrum phagocytes treated with colostrum melatonin adsorbed to PEG microspheres exhibited increased superoxide, accompanied by a decrease in the release of superoxide dismutase (SOD) and a lower SOD-to-superoxide ratio. In contrast, synthetic melatonin reduced the release of superoxide and increased the release of the enzyme and the SOD-to-superoxide ratio. Conclusions: These data highlight the importance of melatonin on cellular metabolism and suggest that colostrum-derived melatonin may be a more effective option for controlling oxidative metabolism, particularly during infectious processes. Full article
(This article belongs to the Section Cell Metabolism)
Show Figures

Figure 1

Figure 1
<p>Representative scheme for obtaining samples and experimental design.</p>
Full article ">Figure 2
<p>Fluorescence microscopy images of PEG microspheres stained with DyLight 488 (1000×): (<b>a</b>) represents the PEG microsphere without melatonin adsorption, (<b>b</b>) represents the microsphere incubated with melatonin extracted from colostrum, and (<b>c</b>) represents the microsphere incubated with synthetic melatonin.</p>
Full article ">Figure 3
<p>Three-dimensional image of size (panel (<b>a</b>)) and fluorescence intensity (FL-2; panel (<b>b</b>)) and of PEG microspheres adsorbed or not with melatonin extracted from colostrum stained with phycoerythrin (PE) as described in Materials and Methods. The standard PE-labeled polymethylmethacrylate microsphere (BD Microsphere, Becton Dickinson, San Jose, CA, USA) was used as a standard (FACScalibur, Becton Dickinson, San Jose, USA). Flow cytometry (FACScalibur, Becton Dickinson, USA) evaluated microsphere size and immunofluorescence analysis. Gray color—PEG microsphere; blue color—BD microsphere; green color—microsphere plus synthetic melatonin; red—microsphere plus colostrum melatonin.</p>
Full article ">Figure 4
<p>Melatonin release (pg/mL) from colostrum and synthetic melatonin in polyethylene glycol (PEG) microspheres during 24 h of incubation in RPMI 1640 medium (n = 3). Melatonin from colostrum (MLT-C); synthetic melatonin (MLT-S).</p>
Full article ">Figure 5
<p>Superoxide release from human colostrum MN phagocytes. Results represent the mean and standard deviation of 20 colostrum samples. Cells were treated with PEG microspheres, colostrum melatonin (MLT-C), or synthetic melatonin (MLT-S) (<b>a</b>). They were also incubated with enteropathogenic <span class="html-italic">Escherichia coli</span> (EPEC) and then treated with PEG microspheres containing either MLT-C or MLT-S (<b>b</b>). The letter “a” indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05) in superoxide anion release between the PBS-treated group and those treated with MLT-C and MLT-S. Letter “b” shows differences between MLT-C-PEG and MLT-S-PEG treatments.</p>
Full article ">Figure 6
<p>The concentration of CuZn-superoxide dismutase (SOD) in the culture supernatant of mononuclear (MN) phagocytes from human colostrum was analyzed using 20 samples. MN cells were treated with colostrum-derived and synthetic melatonin (<b>a</b>). Additionally, MN cells exposed to enteropathogenic <span class="html-italic">Escherichia coli</span> (EPEC) were treated with PEG microspheres containing colostrum-derived or synthetic melatonin (<b>b</b>). Results showed significance at <span class="html-italic">p</span> &lt; 0.05. The letter “a” highlights differences in enzyme levels between the PBS group and those treated with colostrum-derived melatonin (MLT-C), synthetic melatonin (MLT-S), or PEG microspheres (MLT-C-PEG and MLT-S-PEG). Letter “b” indicates the differences between the PEG microspheres with colostrum melatonin (MLT-C-PEG) and those with synthetic melatonin (MLT-S-PEG).</p>
Full article ">Figure 7
<p>SOD/superoxide ratio in human colostrum MN phagocytes with colostrum melatonin and synthetic melatonin (<b>a</b>). Results represent the mean and standard deviation of 20 colostrum samples. It also presents the ratio for MN phagocytes exposed to EPEC (enteropathogenic <span class="html-italic">Escherichia coli</span>) and treated with PEG microspheres containing either type of melatonin (<b>b</b>), with <span class="html-italic">p</span> &lt; 0.05. The letter “a” highlights differences in enzyme levels between the PBS group and those treated with colostrum-derived melatonin (MLT-C), synthetic melatonin (MLT-S), or PEG microspheres (MLT-C-PEG and MLT-S-PEG). Letter “b” indicates the differences between the PEG microspheres with colostrum melatonin (MLT-C-PEG) and those with synthetic melatonin (MLT-S-PEG).</p>
Full article ">
36 pages, 1986 KiB  
Review
Exploring Innovative Approaches for the Analysis of Micro- and Nanoplastics: Breakthroughs in (Bio)Sensing Techniques
by Denise Margarita Rivera-Rivera, Gabriela Elizabeth Quintanilla-Villanueva, Donato Luna-Moreno, Araceli Sánchez-Álvarez, José Manuel Rodríguez-Delgado, Erika Iveth Cedillo-González, Garima Kaushik, Juan Francisco Villarreal-Chiu and Melissa Marlene Rodríguez-Delgado
Biosensors 2025, 15(1), 44; https://doi.org/10.3390/bios15010044 - 13 Jan 2025
Viewed by 1024
Abstract
Plastic pollution, particularly from microplastics (MPs) and nanoplastics (NPs), has become a critical environmental and health concern due to their widespread distribution, persistence, and potential toxicity. MPs and NPs originate from primary sources, such as cosmetic microspheres or synthetic fibers, and secondary fragmentation [...] Read more.
Plastic pollution, particularly from microplastics (MPs) and nanoplastics (NPs), has become a critical environmental and health concern due to their widespread distribution, persistence, and potential toxicity. MPs and NPs originate from primary sources, such as cosmetic microspheres or synthetic fibers, and secondary fragmentation of larger plastics through environmental degradation. These particles, typically less than 5 mm, are found globally, from deep seabeds to human tissues, and are known to adsorb and release harmful pollutants, exacerbating ecological and health risks. Effective detection and quantification of MPs and NPs are essential for understanding and mitigating their impacts. Current analytical methods include physical and chemical techniques. Physical methods, such as optical and electron microscopy, provide morphological details but often lack specificity and are time-intensive. Chemical analyses, such as Fourier transform infrared (FTIR) and Raman spectroscopy, offer molecular specificity but face challenges with smaller particle sizes and complex matrices. Thermal analytical methods, including pyrolysis gas chromatography–mass spectrometry (Py-GC-MS), provide compositional insights but are destructive and limited in morphological analysis. Emerging (bio)sensing technologies show promise in addressing these challenges. Electrochemical biosensors offer cost-effective, portable, and sensitive platforms, leveraging principles such as voltammetry and impedance to detect MPs and their adsorbed pollutants. Plasmonic techniques, including surface plasmon resonance (SPR) and surface-enhanced Raman spectroscopy (SERS), provide high sensitivity and specificity through nanostructure-enhanced detection. Fluorescent biosensors utilizing microbial or enzymatic elements enable the real-time monitoring of plastic degradation products, such as terephthalic acid from polyethylene terephthalate (PET). Advancements in these innovative approaches pave the way for more accurate, scalable, and environmentally compatible detection solutions, contributing to improved monitoring and remediation strategies. This review highlights the potential of biosensors as advanced analytical methods, including a section on prospects that address the challenges that could lead to significant advancements in environmental monitoring, highlighting the necessity of testing the new sensing developments under real conditions (composition/matrix of the samples), which are often overlooked, as well as the study of peptides as a novel recognition element in microplastic sensing. Full article
(This article belongs to the Special Issue Micro-nano Optic-Based Biosensing Technology and Strategy)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of the most common polymers present in MPs and NPs.</p>
Full article ">Figure 2
<p>Scheme of the current methods for analyzing microplastics and nanoplastics.</p>
Full article ">Figure 3
<p>Scheme of electrochemical (bio)sensing approaches (with or without receptors) for the analysis of microplastics.</p>
Full article ">Figure 4
<p>Scheme of plasmonic and fluorescence sensing approaches: (<b>a</b>) colorimetric methods, (<b>b</b>) surface plasmon resonance, (<b>c</b>) localized surface plasmon resonance, (<b>d</b>) surface-enhanced Raman spectroscopy, and (<b>e</b>) fluorescence for the analysis of microplastics.</p>
Full article ">
10 pages, 2538 KiB  
Article
Rapid Acquisition of High-Pixel Fluorescence Lifetime Images of Living Cells via Image Reconstruction Based on Edge-Preserving Interpolation
by Yinru Zhu, Yong Guo, Xinwei Gao, Qinglin Chen, Yingying Chen, Ruijie Xiang, Baichang Lin, Luwei Wang, Yuan Lu and Wei Yan
Biosensors 2025, 15(1), 43; https://doi.org/10.3390/bios15010043 - 13 Jan 2025
Viewed by 541
Abstract
Fluorescence lifetime imaging (FLIM) has established itself as a pivotal tool for investigating biological processes within living cells. However, the extensive imaging duration necessary to accumulate sufficient photons for accurate fluorescence lifetime calculations poses a significant obstacle to achieving high-resolution monitoring of cellular [...] Read more.
Fluorescence lifetime imaging (FLIM) has established itself as a pivotal tool for investigating biological processes within living cells. However, the extensive imaging duration necessary to accumulate sufficient photons for accurate fluorescence lifetime calculations poses a significant obstacle to achieving high-resolution monitoring of cellular dynamics. In this study, we introduce an image reconstruction method based on the edge-preserving interpolation method (EPIM), which transforms rapidly acquired low-resolution FLIM data into high-pixel images, thereby eliminating the need for extended acquisition times. Specifically, we decouple the grayscale image and the fluorescence lifetime matrix and perform an individual interpolation on each. Following the interpolation of the intensity image, we apply wavelet transformation and adjust the wavelet coefficients according to the image gradients. After the inverse transformation, the original image is obtained and subjected to noise reduction to complete the image reconstruction process. Subsequently, each pixel is pseudo-color-coded based on its intensity and lifetime, preserving both structural and temporal information. We evaluated the performance of the bicubic interpolation method and our image reconstruction approach on fluorescence microspheres and fixed-cell samples, demonstrating their effectiveness in enhancing the quality of lifetime images. By applying these techniques to live-cell imaging, we can successfully obtain high-pixel FLIM images at shortened intervals, facilitating the capture of rapid cellular events. Full article
Show Figures

Figure 1

Figure 1
<p>Workflow of image reconstruction based on edge-preserving interpolation. The green boxes represent data processing operations. The colored, black-and-white, and gray rounded rectangles denote the lifetime information, intensity information, and gradient orientation information of the fluorescence lifetime images, respectively. The dashed blue box presents the workflow of the adopted edge-preserving interpolation method. The blue squares contain the data from the original low-resolution image, which are directly assigned to the odd rows and columns of the high-resolution image. The pink squares contain the data from the first step of interpolation, their calculation involving the surrounding 4 × 4 pixels of the original image, and these data are interpolated to the even rows and columns of the high-pixel image. The yellow squares contain data from the second step of interpolation, calculated based on a 5 × 5 neighborhood in the high-pixel image, and these data are distributed on the odd rows and even columns or even rows and odd columns of the high-resolution image.</p>
Full article ">Figure 2
<p>Fluorescence lifetime images of microsphere captured at a low resolution (LP) of 256 × 256 pix-els, processed using bicubic interpolation, reconstructed through EPIM, and acquired at a high resolution serving as the ground truth (GT), respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) Fluorescence lifetime image comparison of the LP, bicubic-interpolated, EPIM reconstruction, and GT images, divided by dotted lines. The amplification factor was set to 2. (<b>b</b>–<b>d</b>) ROIs of (<b>a</b>). (<b>e</b>) Fluorescence lifetime distribution histogram of (<b>a</b>). (<b>f</b>–<b>h</b>) Spatially resolved analysis at the white dashed line in (<b>b</b>), (<b>c</b>), and (<b>e</b>), respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) Fluorescence lifetime image comparison of the LP (128 × 128), bicubic-interpolated, EPIM reconstruction, and GT images, divided by dotted lines. (<b>b</b>–<b>d</b>) ROIs of (<b>a</b>). (<b>e</b>) Fluorescence lifetime distribution histogram of (<b>a</b>).</p>
Full article ">Figure 5
<p>(<b>a</b>) Fluorescence lifetime image of live cells. The mitochondria are marked. (<b>b</b>) Regions of interest of (<b>a</b>). These images were collected continuously, with the first one starting at 0 min and the interval between each start time being 1 min. White arrows: rapid alterations in mitochondrial characteristics can be observed.</p>
Full article ">
14 pages, 6550 KiB  
Article
Rapid Degradation of Organic Dyes by Nanostructured Gd2O3 Microspheres
by Carlos R. Michel
Appl. Nano 2025, 6(1), 1; https://doi.org/10.3390/applnano6010001 - 13 Jan 2025
Viewed by 413
Abstract
Pollution of freshwater by synthetic organic dyes is a major concern due to their high toxicity and mutagenicity. In this study, the degradation of Congo red (CR) and malachite green (MG) dyes was investigated using nanostructured Gd2O3. It was [...] Read more.
Pollution of freshwater by synthetic organic dyes is a major concern due to their high toxicity and mutagenicity. In this study, the degradation of Congo red (CR) and malachite green (MG) dyes was investigated using nanostructured Gd2O3. It was prepared using the coprecipitation method, using gadolinium nitrate and concentrated formic acid, with subsequent calcination at 600 °C. Its morphology corresponds to hollow porous microspheres with a size between 0.5 and 7.5 μm. The optical bandgap energy was determined by using the Tauc method, giving 4.8 eV. The degradation of the dyes was evaluated by UV-vis spectroscopy, which revealed that dissociative adsorption (in the dark) played a key role. It is explained by the cleavage and fragmentation of the organic molecules by hydroxyl radicals (OH), superoxide radicals (O2) and other reactive oxygen species (ROS) produced on the surface of Gd2O3. For CR, the degradation percentage was ~56%, through dissociative adsorption, while UV light photocatalysis increased it to ~65%. For MG, these values were ~78% and ~91%, respectively. The difference in degradation percentages is explained in terms of the isoelectric point of solid (IEPS) of Gd2O3 and the electrical charge of the dyes. FTIR and XPS spectra provided evidence of the role of ROS in dye degradation. Full article
Show Figures

Figure 1

Figure 1
<p>XRD pattern of the powder calcined at 600 °C, in air, and the corresponding JCPDF pattern (at the bottom) used for its identification.</p>
Full article ">Figure 2
<p>FTIR spectrum of Gd<sub>2</sub>O<sub>3</sub> annealed at 600 °C.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) FESEM images of Gd<sub>2</sub>O<sub>3</sub> microspheres, and (<b>d</b>) TEM photograph of this sample.</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>–<b>c</b>) FESEM images of Gd<sub>2</sub>O<sub>3</sub> microspheres, and (<b>d</b>) TEM photograph of this sample.</p>
Full article ">Figure 4
<p>XPS spectra of a Gd<sub>2</sub>O<sub>3</sub> sample: (<b>a</b>) wide scan showing peaks of gadolinium, carbon, and oxygen ions; (<b>b</b>,<b>c</b>) narrow spectra of Gd 3d and 4d orbitals, respectively, and (<b>d</b>) O 1s spectrum.</p>
Full article ">Figure 4 Cont.
<p>XPS spectra of a Gd<sub>2</sub>O<sub>3</sub> sample: (<b>a</b>) wide scan showing peaks of gadolinium, carbon, and oxygen ions; (<b>b</b>,<b>c</b>) narrow spectra of Gd 3d and 4d orbitals, respectively, and (<b>d</b>) O 1s spectrum.</p>
Full article ">Figure 5
<p>(<b>a</b>) Absorbance vs. wavelength graph and (<b>b</b>) Tauc graph used to determine the optical bandgap energy of Gd<sub>2</sub>O<sub>3</sub> (<span class="html-italic">E<sub>g</sub></span> = ~4.8 eV).</p>
Full article ">Figure 6
<p>(<b>a</b>) Absorbance spectra collected during the photocatalytic degradation of CR in UV light; (<b>b</b>) degradation percentage curves obtained from photocatalysis, adsorption, and photolysis, (<b>c</b>,<b>d</b>) cycling test graphs for photocatalytic and adsorption degradation, respectively, and (<b>e</b>) scheme of dissociative adsorption of a CR molecule by Gd<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 7
<p>(<b>a</b>) UV-Vis spectra of the photocatalytic degradation of MG, and (<b>b</b>) degradation percentage curves.</p>
Full article ">Figure 8
<p>Narrow XPS profile of the O 1 s peak measured after CR degradation.</p>
Full article ">Figure 9
<p>FTIR spectra obtained from Gd<sub>2</sub>O<sub>3</sub> samples before and after CR degradation.</p>
Full article ">
11 pages, 2265 KiB  
Article
Graphene-Based, Flexible, Wearable Piezoresistive Sensors with High Sensitivity for Tiny Pressure Detection
by Rui Li, Jiahao Hu, Yalong Li, Yi Huang, Lin Wang, Mohan Huang, Zhikun Wang, Junlang Chen, Yan Fan and Liang Chen
Sensors 2025, 25(2), 423; https://doi.org/10.3390/s25020423 - 13 Jan 2025
Viewed by 598
Abstract
Flexible, wearable, piezoresistive sensors have significant potential for applications in wearable electronics and electronic skin fields due to their simple structure and durability. Highly sensitive, flexible, piezoresistive sensors with the ability to monitor laryngeal articulatory vibration supply a new, more comfortable and versatile [...] Read more.
Flexible, wearable, piezoresistive sensors have significant potential for applications in wearable electronics and electronic skin fields due to their simple structure and durability. Highly sensitive, flexible, piezoresistive sensors with the ability to monitor laryngeal articulatory vibration supply a new, more comfortable and versatile way to aid communication for people with speech disorders. Here, we present a piezoresistive sensor with a novel microstructure that combines insulating and conductive properties. The microstructure has insulating polystyrene (PS) microspheres sandwiched between a graphene oxide (GO) film and a metallic nanocopper-graphene oxide (n-Cu/GO) film. The piezoresistive performance of the sensor can be modulated by controlling the size of the PS microspheres and doping degree of the copper nanoparticles. The sensor demonstrates a high sensitivity of 232.5 kPa−1 in a low-pressure range of 0 to 0.2 kPa, with a fast response of 45 ms and a recovery time of 36 ms, while also exhibiting excellent stability. The piezoresistive performance converts subtle laryngeal articulatory vibration into a stable, regular electrical signal; in addition, there is excellent real-time monitoring capability of human joint movements. This work provides a new idea for the development of wearable electronic devices, healthcare, and other fields. Full article
(This article belongs to the Section Nanosensors)
Show Figures

Figure 1

Figure 1
<p>Scheme of the piezoresistive sensor preparation process.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) Working principle and corresponding equiv. circuits of the sensor operating at different pressures. (<b>d</b>) SEM image of the surface of GO film. (<b>e</b>) SEM image of the sandwich structure of the sensor. (<b>f</b>) SEM image of the surface of n-Cu/GO film. (<b>g</b>) SEM image of cross-section of GO film. (<b>h</b>) XRD pattern of GO film and n-Cu/GO film. (<b>i</b>) SEM image of cross-section of n-Cu/GO film.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Resistance change rate of sensors prepared of different particle sizes (50 μm, 100 μm and 200 μm); PS microspheres at different pressures. (<b>c</b>) Resistance change rate of sensors with copper nanoparticle dopings of 0 wt%, 5 wt%, 10 wt%, 15 wt% and 20 wt% at 1 kPa pressure. (<b>d</b>) Resistance change rate versus the pressure applied to the sensor. (<b>e</b>,<b>f</b>) Response time and recovery time of the sensor. (<b>g</b>) The durability of the sensor for loading and unloading pressure for 2800 cycles.</p>
Full article ">Figure 4
<p>(<b>a</b>) Signal responses caused by laryngeal articulatory vibration when the sensor was attached to the larynx. (<b>b</b>) Electrical signals of the sensor caused by wheat and rice. (<b>c</b>) Electrical signals generated by periodic finger presses. (<b>d</b>) Electrical signals generated by bending finger at different angles. (<b>e</b>) Electrical signals generated by bending wrist at different angles. (<b>f</b>) Electrical signals generated by bending knee at different angles.</p>
Full article ">
17 pages, 4378 KiB  
Article
Snapshot Imaging of Stokes Vector Polarization Speckle in Turbid Optical Phantoms and In Vivo Tissues
by Daniel C. Louie, Carla Kulcsar, Héctor A. Contreras-Sánchez, W. Jeffrey Zabel, Tim K. Lee and Alex Vitkin
Photonics 2025, 12(1), 59; https://doi.org/10.3390/photonics12010059 - 11 Jan 2025
Viewed by 571
Abstract
Significance: We present a system to measure and analyze the complete polarization state distribution of speckle patterns generated from in vivo tissue. Accurate measurement of polarization speckle requires both precise spatial registration and rapid polarization state acquisition. A unique measurement system must be [...] Read more.
Significance: We present a system to measure and analyze the complete polarization state distribution of speckle patterns generated from in vivo tissue. Accurate measurement of polarization speckle requires both precise spatial registration and rapid polarization state acquisition. A unique measurement system must be designed to achieve accurate images of polarization speckle patterns for detailed investigation of the scattering properties of biological tissues in vivo. Aim and approach: This system features a polarization state analyzer with no moving parts. Two pixel-polarizer cameras allow for the instantaneous acquisition of the spatial Stokes vector distribution of polarization speckle patterns. System design and calibration methods are presented, and representative images from measurements on liquid phantoms (microsphere suspensions) and in vivo healthy and tumor murine models are demonstrated and discussed. Results and Conclusions: Quantitative measurements of polarization speckle from microsphere suspensions with controlled scattering coefficients demonstrate differences in speckle contrast, speckle size, and the degree of polarization. Measurements on in vivo murine skin and xenograft tumor tissue demonstrate the ability of the system to acquire snapshot polarization speckle images in living systems. The developed system can thus rapidly and accurately acquire polarization speckle images from different media in dynamic conditions such as in vivo tissue. This capability opens the potential for future detailed investigation of polarization speckle for in vivo biomedical applications. Full article
(This article belongs to the Special Issue New Shining Spots in Biomedical Photonics)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The Poincaré sphere is a geometrical interpretation of polarization state. Note the use of white for linear polarization, blue for right-hand helicity, and yellow for left-hand helicity. This interpretation is reflected in the elliptical markers of (<b>B</b>), which indicate the polarization state. (<b>B</b>) An array of elliptical markers representing the polarization state distribution of polarization speckle, generated from a microsphere suspension (see <a href="#sec4dot2-photonics-12-00059" class="html-sec">Section 4.2</a>). Ellipses have size proportional to the degree of polarization (DOP) at that pixel, are oriented in the direction of linear polarization, and are color coded as in (<b>A</b>). Note that spatial DOP (DOP<sub>sp</sub>) as calculated over the entire image is much lower than the DOP at any individual location (pixel). A measure of depolarization thus strongly depends on the area used to calculate it.</p>
Full article ">Figure 2
<p>(<b>A</b>) Optical path: Incident laser is directed by a mirror through a polarization state generator comprised of a halfwave plate (HWP), linear polarizer (LP), and quarter-wave plate (QWP). Speckle is observed through a 10× magnifying objective (Olympus WHK 10×/20). Non-polarizing beam splitter (BS) divides signal between two pixel-polarization cameras. Each camera has a QWP oriented with fast axis at 67.5° to produce eight unique polarization analyzers. (<b>B</b>) The polarization analyzers set to eight elliptical states placed at points around the Poincaré sphere. These points inscribe a cube of maximum volume within the sphere, with each point being as far from the poles and equator as possible. These conditions minimize the influence of noise and random error in the calculation of a Stokes vector.</p>
Full article ">Figure 3
<p>Example of polarization speckle from a polished metal surface, generated with right-handed circular incident polarization, displaying (<b>A</b>) S<sub>0</sub> channel speckle pattern (polarization agnostic), (<b>B</b>) corresponding elliptical polarization indicators, and (<b>C</b>) overlay of (<b>A</b>,<b>B</b>). Red boxes indicate areas of low total intensity that distort DOP. Note the high contrast in the S<sub>0</sub> channel and uniform left-handed helicity (flipped from right-handed incidence due to reflection) across the whole image. As seen in (<b>C</b>), ellipse sizes are similar in both light and dark regions of the image, indicating similar DOP. Quantification of polarization speckle metrics is shown in <a href="#photonics-12-00059-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Examples of S<sub>0</sub> channel speckle patterns (left) with elliptical polarization indicators (right) from microsphere suspensions. Samples selected from (<b>A</b>,<b>B</b>) 1.07 µm microspheres dispersed in water at µ<sub>s</sub> = 100 cm<sup>−1</sup> and µ<sub>s</sub> = 300 cm<sup>−1</sup>, respectively; (<b>C</b>,<b>D</b>) 0.58 µm microspheres dispersed in water at µ<sub>s</sub> = 100 cm<sup>−1</sup> and µ<sub>s</sub> = 300 cm<sup>−1</sup>, respectively. Note the mixed regions of helicity that contribute to low <span class="html-italic">DOP<sub>sp</sub></span> relative to speckle from metal in <a href="#photonics-12-00059-f003" class="html-fig">Figure 3</a>. Quantification of polarization speckle metrics in <a href="#photonics-12-00059-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 5
<p>Examples of S<sub>0</sub> channel speckle patterns (left) with polarization markers (right) from (<b>A</b>) <span class="html-italic">in vivo</span> healthy mouse skin and (<b>B</b>) <span class="html-italic">in vivo</span> mouse tumor model. As in the volumetric phantoms of <a href="#photonics-12-00059-f004" class="html-fig">Figure 4</a>, mixed regions of helicity contribute to lower <span class="html-italic">DOP<sub>sp</sub></span>. Quantification of polarization speckle metrics in <a href="#photonics-12-00059-t001" class="html-table">Table 1</a>.</p>
Full article ">
Back to TopTop