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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,056)

Search Parameters:
Keywords = biomedical sensor

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
27 pages, 5720 KiB  
Review
MXene-Based Electrochemical Biosensors: Advancing Detection Strategies for Biosensing (2020–2024)
by Joydip Sengupta and Chaudhery Mustansar Hussain
Biosensors 2025, 15(3), 127; https://doi.org/10.3390/bios15030127 - 20 Feb 2025
Abstract
MXenes, a class of two-dimensional materials, have emerged as promising candidates for developing advanced electrochemical biosensors due to their exceptional electrical conductivity, large surface area, and rich surface chemistry. These unique properties enable high sensitivity, rapid response, and versatile functionalization, making MXene-based biosensors [...] Read more.
MXenes, a class of two-dimensional materials, have emerged as promising candidates for developing advanced electrochemical biosensors due to their exceptional electrical conductivity, large surface area, and rich surface chemistry. These unique properties enable high sensitivity, rapid response, and versatile functionalization, making MXene-based biosensors highly suitable for detecting biomolecules and pathogens in biomedical applications. This review explores recent advancements in MXene-based electrochemical biosensors from 2020 to 2024, focusing on their design principles, fabrication strategies, and integration with microfluidic platforms for enhanced performance. The potential of MXene sensors to achieve real-time and multiplexed detection is highlighted, alongside the associated challenges. Emphasis is placed on the role of MXenes in addressing critical needs in disease diagnostics, personalized medicine, and point-of-care testing, providing insights into future trends and transformative possibilities in the field of biomedical sensing technologies. Full article
(This article belongs to the Special Issue Nano/Micro Biosensors for Biomedical Applications (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Trend of research publications on electrochemical biosensors for biomedical applications, retrieved from Scopus (Elsevier) using the search term “TITLE-ABS-KEY (biosensor AND electrochemical AND biomedical)”.</p>
Full article ">Figure 2
<p>Representative structures and compositions of MXenes. MXenes are characterized by the general formula M<sub>n+1</sub>X<sub>n</sub>T<sub>x</sub>, where M denotes an early transition metal, X represents carbon (C) or nitrogen (N), and T<sub>x</sub> indicates surface terminations such as -OH, -O, or -F. The n value, which determines the structural composition, can range from 1 to 4. The M-sites may be occupied by one, two, or more transition metals, leading to the formation of solid solutions or distinct ordered structures. Notably, double-transition-metal MXenes can exhibit unique arrangements, including in-plane ordered structures (i-MXenes), in-plane vacancy configurations, and out-of-plane ordered structures (o-MXenes), where one or two layers of a secondary transition metal M<sup>II</sup> are intercalated between layers of the primary transition metal M<sup>I</sup> (reproduced with permission from [<a href="#B28-biosensors-15-00127" class="html-bibr">28</a>]).</p>
Full article ">Figure 3
<p>Synthesis of MXenes (reproduced with permission from [<a href="#B38-biosensors-15-00127" class="html-bibr">38</a>]).</p>
Full article ">Figure 4
<p>A schematic illustration of the PAMAM@MXene synthesis process and the sensing mechanism for FR detection (reproduced with permission from [<a href="#B79-biosensors-15-00127" class="html-bibr">79</a>]).</p>
Full article ">Figure 5
<p>(<b>A</b>) The DPV curves of the MXene-rGO-Au-based biosensor were recorded for miRNA-21 solutions with concentrations of 0, 1, 5, 10, and 100 fM, as well as 1, 10, 100, and 1 nM, respectively. (<b>B</b>) The DPV current responses were plotted against the logarithmic concentrations of miRNA-21 for the developed biosensor. (<b>C</b>) The effect of hybridization with different 100 fM miRNA-21 concentrations on the performance of the MXene-rGO-Au-based biosensor was assessed. (<b>D</b>) The peak currents of the MXene-rGO-Au-based biosensor were measured across ten repeated measurements at 100 fM concentration (reproduced with permission from [<a href="#B20-biosensors-15-00127" class="html-bibr">20</a>]).</p>
Full article ">Figure 6
<p>A schematic representation of the fabrication process and detection mechanism of the AuNPs&amp;MXene-SPE sensor is presented. (<b>A</b>) The architectural layout and sequential development of the screen-printed electrodes are detailed. (<b>B</b>) The step-by-step fabrication of the AuNPs&amp;MXene-SPE sensor is systematically carried out on the working electrodes. (<b>C</b>) The electrode functionality is achieved by labelling L-cysteine (L-Cys) on the electrode surface through the formation of Au-S bonds, followed by the immobilization of antibodies using amine coupling facilitated by EDC/NHS chemistry. (<b>D</b>) The label-free electrochemical immunobiosensor detects target interactions by monitoring changes in the interfacial electron-transfer kinetics of the redox probe [Fe(CN)<sub>6</sub>]<sup>4−/3−</sup> during antibody–antigen binding events (reproduced with permission from [<a href="#B22-biosensors-15-00127" class="html-bibr">22</a>]).</p>
Full article ">Figure 7
<p>A schematic representation of the biosensor designed for rapid electrochemical detection of West Nile Virus (WNV), utilizing MXene/Tr-WNV aptamer and fabricated through ACEF technology (reproduced with permission from [<a href="#B83-biosensors-15-00127" class="html-bibr">83</a>]).</p>
Full article ">Figure 8
<p>(<b>a</b>) Diagram showcasing the components of a flexible, wearable PyTS@Ti<sub>3</sub>C<sub>2</sub>T<sub>x</sub>-based microfluidic chip; (<b>b</b>) cross-sectional depiction of sweat collection through the chip; (<b>c</b>) schematic representation of the chip’s structure alongside images illustrating the sequential flow of black ink fluid through the inlet into the chamber over time; (<b>d</b>) photograph of a wearable device for on-body monitoring of sweat uric acid; and (<b>e</b>) correlation between on-body uric acid concentration measurements and blood uric acid levels (reproduced with permission from [<a href="#B21-biosensors-15-00127" class="html-bibr">21</a>]).</p>
Full article ">Figure 9
<p>Schematic representation of the fabrication process of the FeVO<sub>4</sub>@Ti<sub>3</sub>C<sub>2</sub> MXene-based biosensor and its application in 5-HT detection (reproduced with permission from [<a href="#B86-biosensors-15-00127" class="html-bibr">86</a>]).</p>
Full article ">Figure 10
<p>Schematic representation of the working principle of an electrochemical biosensor for detecting the BCR-ABL1 fusion gene (reproduced with permission from [<a href="#B88-biosensors-15-00127" class="html-bibr">88</a>]).</p>
Full article ">Scheme 1
<p>Key attributes highlighting the integration potential of MXene in biosensor applications.</p>
Full article ">
15 pages, 5107 KiB  
Article
Feasibility Study of Photoelectrochemical Sensing of Glucose and Urea Using BiVO4 and BiVO4/BiOCl Photoanodes
by Monika Skruodiene, Jelena Kovger-Jarosevic, Irena Savickaja, Jurga Juodkazyte and Milda Petruleviciene
Sensors 2025, 25(4), 1260; https://doi.org/10.3390/s25041260 - 19 Feb 2025
Abstract
This study investigates the photoelectrochemical (PEC) performance of molybdenum-doped bismuth vanadate (Mo-doped BiVO4) and its heterojunction with the BiOCl layer in glucose and urea sensing. Photoelectrochemical analyses, including cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS), revealed that the formation of [...] Read more.
This study investigates the photoelectrochemical (PEC) performance of molybdenum-doped bismuth vanadate (Mo-doped BiVO4) and its heterojunction with the BiOCl layer in glucose and urea sensing. Photoelectrochemical analyses, including cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS), revealed that the formation of a heterojunction enhanced charge carrier separation. The impact of the interaction between the surface of the photoanode and analytes on sensing performance was systematically evaluated. Among the tested configurations, Mo-doped BiVO4 exhibited superior glucose sensing with a limit of detection (LOD) of 0.173 µM, while BiVO4/BiOCl demonstrated an LOD of 2.474 µM. In the context of urea sensing, Mo-doped BiVO4 demonstrated an LOD of 0.656 µM, while BiVO4/BiOCl exhibited an LOD of 0.918 µM. Notably, despite the enhanced PEC activity observed in heterostructured samples, Mo-doped BiVO4 exhibited superior sensing performance, attributable to good interaction with analytes. The photocurrent response trends—an increase with glucose concentration and a decrease with urea concentration—were attributed to oxidation and adsorption phenomena on the photoanode surface. These findings underscore the critical role of photoanode surface engineering in advancing PEC sensor technology, paving the way for more efficient environmental and biomedical applications. Full article
(This article belongs to the Special Issue Recent Advances in Photo(electro)chemical Sensing and Sensors)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD of BiVO<sub>4</sub> (<b>a</b>), BiVO<sub>4</sub>/BiOCl (<b>b</b>) coatings. Green columns correspond to indicated references of COD; *—Bi<sub>2</sub>O<sub>3</sub> (PDF: 96-153-7329).</p>
Full article ">Figure 2
<p>SEM of BiVO<sub>4</sub> (<b>a</b>,<b>c</b>) and BiVO<sub>4</sub>/BiOCl (<b>b</b>,<b>d</b>).</p>
Full article ">Figure 3
<p>CVs (<b>a</b>) and Nyquist plots (<b>b</b>) of BiVO<sub>4</sub> and BiVO<sub>4</sub>/BiOCl photoelectrodes recorded in borate buffer electrolyte under illumination; inset in (<b>b</b>) shows equivalent circuit used for fitting the EIS data.</p>
Full article ">Figure 4
<p>I/I<sub>0</sub> vs. glucose concentration plots for BiVO<sub>4</sub> (<b>a</b>) and BiVO<sub>4</sub>/BiOCl (<b>b</b>) photoelectrodes. Number of replicates: n = 3; the limits of the experimental errors are indicated at the experimental points.</p>
Full article ">Figure 5
<p>I/I<sub>0</sub> vs. urea concentration plots for BiVO<sub>4</sub> (<b>a</b>) and BiVO<sub>4</sub>/BiOCl (<b>b</b>) photoelectrodes in sodium borate buffer. Number of replicates: n = 3.</p>
Full article ">Figure 6
<p>Chronoamperograms of BiVO<sub>4</sub> (<b>a</b>) and BiVO<sub>4</sub>/BiOCl (<b>b</b>) photoelectrodes recorded in borate buffer without or with urea at 1.2 V; linear relationship between I/I<sub>0</sub> and urea concentration for BiVO<sub>4</sub> and BiVO<sub>4</sub>/BiOCl (<b>c</b>).</p>
Full article ">Figure 7
<p>CVs of BiVO4 (<b>a</b>,<b>b</b>) and BiVO<sub>4</sub>/BiOCl (<b>c</b>,<b>d</b>) photoelectrodes in sodium borate buffer containing glucose (<b>a</b>,<b>c</b>) and urea (<b>b</b>,<b>d</b>).</p>
Full article ">Figure 8
<p>Nyquist plots of BiVO<sub>4</sub> (<b>a</b>) and BiVO<sub>4</sub>/BiOCl (<b>b</b>) photoelectrodes recorded in sodium borate buffer in the absence or presence of 30 mM of glucose or urea.</p>
Full article ">Figure 9
<p>ΔOCP vs. concentration of analyte for BiVO<sub>4</sub> and BiVO<sub>4</sub>/BiOCl photoelectrodes measured in sodium borate buffer in the absence or presence of glucose or urea.</p>
Full article ">Figure 10
<p>Selectivity performance of BiVO<sub>4</sub> and BiVO<sub>4</sub>/BiOCl photoanodes in 0.2 M sodium borate buffer, evaluated using sequential additions of indicated analytes during chronoamperometric measurements at 1.2 V.</p>
Full article ">
33 pages, 14611 KiB  
Review
Silicon-Based Biosensors: A Critical Review of Silicon’s Role in Enhancing Biosensing Performance
by Waqar Muhammad, Jaeyoon Song, Sehyeon Kim, Falguni Ahmed, Eunseo Cho, Huiseop Lee and Jinsik Kim
Biosensors 2025, 15(2), 119; https://doi.org/10.3390/bios15020119 - 18 Feb 2025
Abstract
This review into recent advancements in silicon-based technology, with a particular emphasis on the biomedical applications of silicon sensors. Owing to their diminutive size, high sensitivity, and intrinsic compatibility with electronic systems, silicon-based sensors have found widespread utilization across healthcare, industrial, and environmental [...] Read more.
This review into recent advancements in silicon-based technology, with a particular emphasis on the biomedical applications of silicon sensors. Owing to their diminutive size, high sensitivity, and intrinsic compatibility with electronic systems, silicon-based sensors have found widespread utilization across healthcare, industrial, and environmental monitoring domains. In the realm of biomedical sensing, silicon has demonstrated significant potential to enhance human health outcomes while simultaneously driving progress in microfabrication techniques for multifunctional device development. The review systematically examines the versatile roles of silicon in the fabrication of electrodes, sensing channels, and substrates. Silicon electrodes are widely used in electrochemical biosensors for glucose monitoring and neural activity recording, while sensing channels in field-effect transistor biosensors enable the detection of cancer biomarkers and small molecules. Porous silicon substrates are applied in optical biosensors for label-free protein and pathogen detection. Key challenges in this field, including the interaction of silicon with biomolecules, the economic barriers to miniaturization, and issues related to signal stability, are critically analyzed. Proposed strategies to address these challenges and improve sensor functionality and reliability are also discussed. Furthermore, the article explores emerging developments in silicon-based biosensors, particularly their integration into wearable technologies. The pivotal role of artificial intelligence (AI) in enhancing the performance, functionality, and real-time capabilities of these sensors is also highlighted. This review provides a comprehensive overview of the current state, challenges, and future directions in the field of silicon-based biomedical sensing technologies. Full article
Show Figures

Figure 1

Figure 1
<p>Classification of silicon-based biosensors categorized into three major types: electrochemical, optical, and mechanical biosensors. Each quadrant highlights key sensing mechanisms, including SiNWs FETs, Copyright 2016, Royal Society of Chemistry [<a href="#B54-biosensors-15-00119" class="html-bibr">54</a>]. ISFETs, reproduced from [<a href="#B55-biosensors-15-00119" class="html-bibr">55</a>] under Creative Commons Attribution License, 2022. Hydrogel-gated FETs, Copyright 2018, Elsevier [<a href="#B56-biosensors-15-00119" class="html-bibr">56</a>]. TFET, Copyright 2024, Elsevier [<a href="#B57-biosensors-15-00119" class="html-bibr">57</a>]. Capacitive, Copyright 2022, Elsevier [<a href="#B58-biosensors-15-00119" class="html-bibr">58</a>]. Optical, reproduced from [<a href="#B59-biosensors-15-00119" class="html-bibr">59</a>] under Creative Commons Attribution License, 2011. Photonic, reproduced from [<a href="#B60-biosensors-15-00119" class="html-bibr">60</a>] under Creative Commons Attribution License, 2021. Microcantilever, Copyright 2024, Springer Nature [<a href="#B61-biosensors-15-00119" class="html-bibr">61</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Image showing the packaged SiNW-FET device. (<b>b</b>) Scanning electron microscope (SEM) image of the SiNW-FET device. (<b>c</b>) SEM view of a region with an array of silicon nanowires. (<b>d</b>) SEM image of a silicon nanowire from the schematic diagram of the SiNW-FET device. (<b>e</b>) Schematic illustration of the biosensor detection mechanism of the SiNW-FET. (<b>f</b>) Normalized current response with the Ag85B concentration gradient. (<b>g</b>) Normalized current response of the SiNW-FET to 10 ng/mL, demonstrating the biosensor’s specificity in detecting six different proteins. (<b>a</b>–<b>g</b>) reproduced with permission from the American Chemical Society, Copyright [2021]. (<b>h</b>) (Left to right) Photograph of the packaged SiNW Bio-FET device, SEM image of the Si-NW array, and TEM image of the cross-section of the SiNW. (<b>i</b>) Schematic diagram of the SiNW Bio-FET for exosomes detection. (<b>j</b>) I<sub>D</sub>-V<sub>G</sub> curves of the anti-CD81 functionalized SiNW Bio-FET after incubation with different concentrations of exosomes derived from 293F cells. (<b>k</b>) The Vth of Si-NW Bio-FET to 293F-derived exosomes at a series of concentrations. (<b>l</b>) The Vth of the SiNW Bio-FET in response to 293F-derived exosomes at various concentrations. (<b>h</b>–<b>l</b>) reproduced with permission from Elsevier, Copyright [2024]. (<b>m</b>) Schematic of the SiNW biosensor featuring a microfluidic channel, (<b>n</b>) (left to right) photograph of the SiNW biosensor, optical microscope image showing the partially amplified SiNW biosensor, SEM image of the SiNW array within the open channel, TEM image of the SiNW. (<b>o</b>) Transfer characteristics of the SiNW FET biosensor in response to different concentrations of Cys-C (the red box highlights the shift in the curve near the sub-threshold region). (<b>p</b>) The Vth of the SiNW FET biosensor in a varying Cys-C concentrations. (<b>q</b>) The modified antibody SiNW FET biosensor’s electrical performance in detecting various biomarkers of early kidney failure. (<b>m</b>–<b>q</b>) reproduced from [<a href="#B67-biosensors-15-00119" class="html-bibr">67</a>] under Creative Commons Attribution License, 2023.</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic representation of the detection mechanism implemented in the ISFET-based handheld testing device for cTnI measurement. (<b>b</b>) Photograph of the ISFET-based handheld testing device utilized for the detection of cTnI. (<b>c</b>) Electrode connection configuration of ISFET in the dual-gate mode. (<b>d</b>) IDS current response to varying cTnI concentrations. (<b>e</b>) Linear curve of varying cTnI concentrations ranging from 1 pg/mL to 1000 pg/mL. (<b>f</b>) Specificity test of ISFET biosensor with four different biomarkers. (<b>a</b>–<b>f</b>) reproduced with permission from Elsevier, Copyright [2021]. (<b>g</b>) Top view and cross-sectional view of Si<sub>3</sub>N<sub>4</sub>-ISFET biosensor. (<b>h</b>) Relationship between anti-HSA antibody at various concentrations and gate potential change in ISFET. (<b>i</b>) Specificity of has-modified ISFET biosensor for the detection of the HSA protein. (<b>g</b>–<b>i</b>) reproduced from [<a href="#B71-biosensors-15-00119" class="html-bibr">71</a>] under Creative Commons Attribution License, 2021. (<b>j</b>) Cross-sectional schematic of the Na<sup>+</sup> ISFET illustrating membrane deposition procedure and the biosensing mechanism. (<b>k</b>) Calibration plots of Na<sup>+</sup> ISFETs over the varying concentration range. (<b>l</b>) The sensitivity of four ISFETs periodically calibrated over a duration of 43 days. (<b>m</b>) Calibration plots of a Na<sup>+</sup> ISFET obtained in deionized water and in KCl solutions with varying concentrations. (<b>j</b>–<b>m</b>) reproduced with permission from Elsevier, Copyright [2023].</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic illustration of hydrogel-gated FET SARS-CoV-2 biosensor. (<b>b</b>) (Left to right) Optical microscopy of unmodified FETs, SEM magnification of the nanonet, and hydrogel-modified FET. (<b>c</b>) Biosensor response to SARS-CoV-2 RBD in PBS. (<b>d</b>) Response to RBD in saliva, with the calibration graph included as an inset. (<b>e</b>) Transfer characteristics of the biosensor after incubation with samples from COVID-19 negative (blue) and COVID-19 positive (red) cases. (<b>a</b>–<b>e</b>) reproduced from [<a href="#B75-biosensors-15-00119" class="html-bibr">75</a>] under Creative Commons Attribution License, 2023. (<b>f</b>) Schematic showing biosensing mechanism of hydrogel-modified SiNW FET. (<b>g</b>) Transfer characteristic curves of hydrogel-functionalized SiNW FET at varying pH concentrations. (<b>h</b>) Real-time response of IDS for a hydrogel-functionalized SiNW FET measured in solutions of varying pH levels. (<b>f</b>–<b>h</b>) reproduced from [<a href="#B76-biosensors-15-00119" class="html-bibr">76</a>] under Creative Commons Attribution License, 2022.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic diagram of InSb/Si TFET-based biomolecule sensor. (<b>b</b>) Impact of dielectric constant on sensor sensitivity. (<b>c</b>) Selectivity versus dielectric constant. (<b>a</b>–<b>c</b>) reproduced with permission from Elsevier, Copyright [2022]. (<b>d</b>) Cross-sectional view DM-DL-DGTFET biosensor. (<b>e</b>) Sensitivity of various step profiles for partly filled cavity at a 66% fill factor, and (<b>f</b>) ~41% fill factor for DL-DGTFET biosensor. (<b>e</b>,<b>f</b>) reproduced with permission from Elsevier, Copyright [2024].</p>
Full article ">Figure 6
<p>(<b>a</b>) Stepwise fabrication process and schematic illustration of the lectin biosensor: (anti-clockwise) bare Al-IDE; surface modification with linker 2-mercaptoacetate, forming a SAM layer with thiol and carboxylate functional groups; deposition of carboxylic acid-functionalized GNP; immobilization of Con A; detection of the target PSMA. (<b>b</b>) Capacitance as a function of logarithmic frequency for Al-IDE, modified with (i) linker, (ii) GNP, (iii) Con A, and (iv–viii) various concentrations of PSMA; (<b>c</b>) calibration curve utilizing capacitance data for varying concentrations; (<b>a-c</b>) reproduced with permission from Elsevier, Copyright [2021]. (<b>d</b>) Micrograph of 3D interdigital electrode arrays (top left), image of the microfluidic-integrated 3D capacitive biosensor (top right), and fluorescent analysis (before and after) of biofunctionalization on gold electrodes following biological modification and antibody incubation. (<b>e</b>) Capacitance responses to reference biomolecules and CRP. (<b>f</b>) Real-time monitoring measures the variation in capacitance value of the electric sensor as the CRP concentration increases. (<b>d</b>–<b>f</b>) reproduced with permission from Royal Society of Chemistry, Copyright [2021]. (<b>g</b>) Schematic of a CMUTs-based label-free biosensor with Al electrodes designed to detect ssDNA oligonucleotides utilizing a self-assembled monolayer (SAM). (<b>h</b>) Cross-sectional SEM micrograph of fabricated CMUTs cell. (<b>i</b>) Frequency responses with various concentrations of complementary ssDNA measured by the impedance analyzer, and (<b>j</b>) internal ASIC interface. (<b>g</b>–<b>j</b>) reproduced from [<a href="#B85-biosensors-15-00119" class="html-bibr">85</a>] under Creative Commons Attribution License, 2024.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic illustrating the fabrication steps for the half-buried and pedestal HCG structures. Cross-sectional SEM images of (<b>b</b>) the pedestal HCG and (<b>c</b>) the half-buried HCG are shown. (<b>d</b>) Immobilization of the myoglobin antibody and myoglobin binding to the biosensor surface. (<b>e</b>) Optical setup: the optical spectrum analyzer (OSA) acts as the detector, while a broadband supercontinuum laser serves as the light source. (<b>f</b>) Resonance wavelength of the pedestal (blue) and half-buried (red) HCGs as a function of refractive index (RI) changes caused by varying glycerol concentrations. (<b>g</b>) Resonance shifts measured between the myoglobin antibody (blank) and myoglobin molecules (antigen). (<b>a</b>–<b>g</b>) reproduced with permission from the American Chemical Society, Copyright [2023]. (<b>h</b>) Schematic of RIFTS measurement: the surface modification section includes analyte molecules that penetrate or alter surface chemistry, changing EOT, reflection spectra from PSi at each step, and FFT intensity from the spectra. The biosensing section includes bacterial suspension monitored with a fluidic system, reflection spectra during two steps, FFT intensity from the spectra, and a real-time biosensor response during washing, bacterial exposure, and final washing, reproduced from [<a href="#B89-biosensors-15-00119" class="html-bibr">89</a>] under Creative Commons Attribution License, 2020. (<b>i</b>) (left) Schematic of the photonic biochip (inset shows a photo of the biochip with a two Euro coin for scale comparison), (right) schematic layout of MZI biosensors, with insets showing the capture of pathogens and proteins on the Al) surface. (<b>j</b>) Absolute wavelength shift for each CRP concentration. (<b>k</b>) Bar graph of the wavelength shift for varying bacteria concentration. (<b>i</b>–<b>k</b>) reproduced with permission from Elsevier, Copyright [2024].</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic showing the biosensing principle of the microcantilever biosensor. (<b>b</b>) Time-dependent resonance frequency curves of aptamer-functionalized microcantilevers exposed to varying concentrations of vitamin D. (<b>c</b>) Calibration plot showing the frequency shift in the aptamer-functionalized microcantilever relative to varying concentrations of vitamin D. (<b>d</b>) Specificity evaluation of the aptamer-based assay. (<b>a</b>–<b>d</b>) reproduced from [<a href="#B92-biosensors-15-00119" class="html-bibr">92</a>] under Creative Commons Attribution License, 2023. (<b>e</b>) The monolithically integrated microcantilever includes a microcantilever array, with analog and digital circuit components. (<b>f</b>) Schematic of the U-shape (left) and piezoresistive microcantilever sensor (right). (<b>g</b>) Results obtained for human IgG measured at varying concentrations. (<b>h</b>) The relationship between saturated voltage changes and varying IgG concentrations. (<b>e</b>–<b>h</b>) reproduced from [<a href="#B93-biosensors-15-00119" class="html-bibr">93</a>] under Creative Commons Attribution License, 2021. (<b>i</b>) The schematic representation of the quarter-configuration Wheatstone bridge. (<b>j</b>) Cross-sectional view of PI/Si/SiO<sub>2</sub> microcantilevers. (<b>k</b>) Measurement results for aflatoxin B1 at concentrations of 1, 10, 20, 50 ng/mL. (<b>l</b>) Linear correlation of the measured output voltage response and aflatoxin B1 concentrations. (<b>m</b>) Detection results for ricin, abrin, and IgG at 100 ng/mL concentrations using microcantilever biosensors functionalized with aflatoxin B1 antibodies. (<b>i</b>–<b>m</b>) reproduced from [<a href="#B94-biosensors-15-00119" class="html-bibr">94</a>] under Creative Commons Attribution License, 2021.</p>
Full article ">
18 pages, 3277 KiB  
Article
Signal Differentiation of Moving Magnetic Nanoparticles for Enhanced Biodetection and Diagnostics
by Kee Young Hwang, Dakota Brown, Supun B. Attanayake, Dan Luu, Minh Dang Nguyen, T. Randall Lee and Manh-Huong Phan
Biosensors 2025, 15(2), 116; https://doi.org/10.3390/bios15020116 - 17 Feb 2025
Abstract
Magnetic nanoparticles are extensively utilized as markers/signal labelling in various biomedical applications. Detecting and distinguishing magnetic signals from similarly sized moving magnetic nanoparticles in microfluidic systems is crucial yet challenging for biosensing. In this study, we have developed an original method to detect [...] Read more.
Magnetic nanoparticles are extensively utilized as markers/signal labelling in various biomedical applications. Detecting and distinguishing magnetic signals from similarly sized moving magnetic nanoparticles in microfluidic systems is crucial yet challenging for biosensing. In this study, we have developed an original method to detect and differentiate magnetic signals from moving superparamagnetic (SPM) and ferrimagnetic (FM) nanoparticles of comparable sizes. Our approach utilizes a highly sensitive magnetic-coil-based sensor that harnesses the combined effects of giant magnetoimpedance (GMI) and an LC-resonance circuit, offering performance superior to that of conventional GMI sensors. Iron oxide nanoparticles, which have similar particle sizes but differing coercivities (zero for SPM and non-zero for FM) or similar zero coercivities but differing particle sizes, flow through the magnetic coil at controlled velocities. Their distinct effects are analyzed through changes in the complex impedance of the sensing system. Our findings provide a unique pathway for utilizing SPM and FM nanoparticles as innovative magnetic markers to identify specific biological entities, thereby expanding their potential applications. Full article
(This article belongs to the Special Issue Biosensing Technologies in Medical Diagnosis)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of nanoparticle detection using a magnetic-coil-based sensor. The sensor distinguishes between superparamagnetic (SPM) and ferrimagnetic (FM) nanoparticles of comparable sizes (<span class="html-italic">D<sub>SPM</sub></span> ~ <span class="html-italic">D<sub>FM</sub></span>) by measuring the impact of their coercivity differences (<span class="html-italic">H<sub>C</sub><sup>SPM</sup></span> = 0 vs. <span class="html-italic">H<sub>C</sub><sup>FM</sup></span> ≠ 0) on the sensor’s sensitivity, represented as changes in impedance (Δ<span class="html-italic">Z</span>), allowing for the identification of each nanoparticle type. An image of the magnetic coil used in the MLCR sensor is also displayed.</p>
Full article ">Figure 2
<p>Frequency dependence of the impedance (<span class="html-italic">Z</span>), resistance (<span class="html-italic">R</span>), and reactance (<span class="html-italic">X</span>) of the magnetic coil in the absence of a magnetic field.</p>
Full article ">Figure 3
<p>(<b>a</b>) The magnetic field dependence of the wire’s impedance at a frequency of 325 MHz, which is the operating frequency of the MLCR sensor. (<b>b</b>) A zoomed-in portion of the low-field Z(H) curve. Points A, B, and C represent the three distinct regimes where the presence of magnetic nanoparticles with different coercivities affects the complex impedance of the wire.</p>
Full article ">Figure 4
<p>SEM images of (<b>a</b>,<b>b</b>) the SPM superparticles (S1: <span class="html-italic">D</span> = 160 nm, <span class="html-italic">C<sub>s</sub></span> = 10 nm; S2: <span class="html-italic">D</span> = 160 nm, <span class="html-italic">C<sub>s</sub></span> = 12 nm) and (<b>c</b>,<b>d</b>) the FM superparticles (S3: <span class="html-italic">D</span> = 160 nm, <span class="html-italic">C<sub>s</sub></span> = 19 nm; S4: <span class="html-italic">D</span> = 160 nm, <span class="html-italic">C<sub>s</sub></span> = 26 nm).</p>
Full article ">Figure 5
<p>The room-temperature magnetic hysteresis (M-H) loops for (<b>a</b>,<b>b</b>) SPM superparticles (S1: <span class="html-italic">D</span> = 160 nm, <span class="html-italic">C<sub>s</sub></span> = 10 nm; S2: <span class="html-italic">D</span> = 160 nm, <span class="html-italic">C<sub>s</sub></span> = 12 nm) and (<b>c</b>,<b>d</b>) FM superparticles (S3: <span class="html-italic">D</span> = 160 nm, <span class="html-italic">C<sub>s</sub></span> = 19 nm; S4: <span class="html-italic">D</span> = 160 nm, <span class="html-italic">C<sub>s</sub></span> = 26 nm).</p>
Full article ">Figure 6
<p>The total impedance of the biosensor varied over time when the iron oxide superparticles of comparable sizes (<span class="html-italic">D</span> = 160 nm) flowed through the magnetic coil: (<b>a</b>) S1, (<b>b</b>) S2, (<b>c</b>) S3, and (<b>d</b>) S4.</p>
Full article ">Figure 7
<p>The correlation between the change in total impedance of the biosensor (∆<span class="html-italic">Z<sub>total</sub></span>) and the saturation magnetization (<span class="html-italic">M<sub>S</sub></span>) for the SPM (S1 and S2) and FM (S3 and S4) superparticles highlights a significant difference in ∆<span class="html-italic">Z<sub>total</sub></span> between S2 and S3 despite their identical particle size and saturation magnetization. This demonstrates the biosensor’s ability to differentiate the signals of these two types of nanoparticles.</p>
Full article ">Figure 8
<p>Change in total impedance (∆<span class="html-italic">Z<sub>total</sub></span> or detection sensitivity) for the magnet, SPM (S1), and FM (S4) superparticle samples. The coercivity of the material significantly influences the biosensor’s detection sensitivity.</p>
Full article ">Figure 9
<p>Change in total impedance (∆<span class="html-italic">Z<sub>total</sub></span> or detection sensitivity) versus particle size for the SPM superparticles, which have nearly identical crystallite sizes (~10 nm). The SEM images display the particle sizes and morphologies of the superparticles.</p>
Full article ">
23 pages, 8687 KiB  
Article
A Low-Voltage Low-Power Voltage-to-Current Converter with Low Temperature Coefficient Design Awareness
by Haoze Chen and Pak Kwong Chan
Sensors 2025, 25(4), 1204; https://doi.org/10.3390/s25041204 - 16 Feb 2025
Abstract
This paper presents a low-voltage, low-power voltage-to-current converter (V-I Converter) implemented in TSMC 40 nm CMOS technology. Operating at a supply voltage of 0.45 V with an input range of 0.1 V to 0.3 V, the proposed circuit achieves a temperature coefficient of [...] Read more.
This paper presents a low-voltage, low-power voltage-to-current converter (V-I Converter) implemented in TSMC 40 nm CMOS technology. Operating at a supply voltage of 0.45 V with an input range of 0.1 V to 0.3 V, the proposed circuit achieves a temperature coefficient of 54.68 ppm/°C, which is at least 2× better than prior works, ensuring stable performance across a wide temperature range (−20 °C to 80 °C). The design employs a three-stage operational transconductance amplifier (OTA) with a Q-reduction frequency compensation technique to produce programmable output currents while maintaining a power dissipation of less than 2.76 μW. With a bandwidth of 34.45 kHz and a total harmonic distortion (THD) of −56.66 dB at 1 kHz and 0.1 VPP input signal, the circuit demonstrates high linearity and low power consumption under ultra-low voltage design scenarios. These features make the proposed V-I Converter highly suitable for energy-constrained applications such as biomedical sensors, energy harvesting systems, and IoT nodes, where low power consumption and temperature stability are critical parameters. Full article
(This article belongs to the Special Issue Advanced Interface Circuits for Sensor Systems (Volume II))
Show Figures

Figure 1

Figure 1
<p>Conventional V-I structures: (<b>a</b>) Using NMOS as the driving transistor; (<b>b</b>) using PMOS as the driving transistor.</p>
Full article ">Figure 2
<p>FBVA [<a href="#B7-sensors-25-01204" class="html-bibr">7</a>].</p>
Full article ">Figure 3
<p>V-I Converter with nested feedback loops [<a href="#B14-sensors-25-01204" class="html-bibr">14</a>].</p>
Full article ">Figure 4
<p>Bulk-driven transconductor.</p>
Full article ">Figure 5
<p>Low-voltage linear V-I conversion unit [<a href="#B17-sensors-25-01204" class="html-bibr">17</a>].</p>
Full article ">Figure 6
<p>Topology of proposed circuit architecture.</p>
Full article ">Figure 7
<p>PR.</p>
Full article ">Figure 8
<p>Cross-section of the PR.</p>
Full article ">Figure 9
<p>The level shifter.</p>
Full article ">Figure 10
<p>(<b>a</b>) The design of the OTA<sub>o</sub> with resistor array as the passive load in the output stage; (<b>b</b>) the bias circuit.</p>
Full article ">Figure 11
<p>System block of OTA<sub>o</sub>.</p>
Full article ">Figure 12
<p>Resistor array implementation.</p>
Full article ">Figure 13
<p>(<b>a</b>) <span class="html-italic">V<sub>O</sub></span><sub>1</sub> versus <span class="html-italic">T</span>; (<b>b</b>) <span class="html-italic">V<sub>O</sub></span><sub>2</sub> versus <span class="html-italic">T</span>.</p>
Full article ">Figure 14
<p>The plot of the variation of <span class="html-italic">|V<sub>TH</sub>|</span> with temperature with different bulk connections.</p>
Full article ">Figure 15
<p>Bode plot of OTA<sub>o</sub> at TT process corner.</p>
Full article ">Figure 16
<p>Closed-loop gain against frequency.</p>
Full article ">Figure 17
<p>Output current versus input voltage.</p>
Full article ">Figure 18
<p>Monte-Carlo simulation result of (<b>a</b>) output current 1, (<b>b</b>) output current 2, (<b>c</b>) output current 3, (<b>d</b>) output current 4, (<b>e</b>) output current 5, (<b>f</b>) output current 6, and (<b>g</b>) output current 7.</p>
Full article ">Figure 18 Cont.
<p>Monte-Carlo simulation result of (<b>a</b>) output current 1, (<b>b</b>) output current 2, (<b>c</b>) output current 3, (<b>d</b>) output current 4, (<b>e</b>) output current 5, (<b>f</b>) output current 6, and (<b>g</b>) output current 7.</p>
Full article ">Figure 19
<p>T.C. in PVT simulation of (<b>a</b>) output current 1, (<b>b</b>) output current 2, (<b>c</b>) output current 3, (<b>d</b>) output current 4, (<b>e</b>) output current 5, (<b>f</b>) output current 6, and (<b>g</b>) output current 7.</p>
Full article ">Figure 20
<p>T.C. in Monte-Carlo simulation of (<b>a</b>) output current 1, (<b>b</b>) output current 2, (<b>c</b>) output current 3, (<b>d</b>) output current 4, (<b>e</b>) output current 5, (<b>f</b>) output current 6 and (<b>g</b>) output current 7.</p>
Full article ">Figure 20 Cont.
<p>T.C. in Monte-Carlo simulation of (<b>a</b>) output current 1, (<b>b</b>) output current 2, (<b>c</b>) output current 3, (<b>d</b>) output current 4, (<b>e</b>) output current 5, (<b>f</b>) output current 6 and (<b>g</b>) output current 7.</p>
Full article ">Figure 21
<p>Comparison of T.C. of two different bulk connections.</p>
Full article ">Figure 22
<p>The plot of PSRR against frequency.</p>
Full article ">Figure 23
<p>The plot of CMRR against frequency.</p>
Full article ">Figure 24
<p>The input-referred noise spectrum.</p>
Full article ">Figure 25
<p>THD.</p>
Full article ">
12 pages, 4717 KiB  
Article
A Ratiometric Fluorescence Nano pH Biosensor for Live-Cell Imaging Using Cerasome
by Zhongqiao Zhang, Xiaoshan Luo, Xuanbo Wang, Meng Liu, Xiuli Yue and Zhaozhu Zheng
Biosensors 2025, 15(2), 114; https://doi.org/10.3390/bios15020114 - 16 Feb 2025
Abstract
The development of a robust and biocompatible pH-sensing platform is critical for monitoring intracellular processes and diagnosing diseases. Here, we present a smart ultrastable ratiometric fluorescence nano pH sensor based on silica-coated liposome nanoparticles (cerasome, 138.4 nm). The sensor integrates pH-sensitive dye, pyranine, [...] Read more.
The development of a robust and biocompatible pH-sensing platform is critical for monitoring intracellular processes and diagnosing diseases. Here, we present a smart ultrastable ratiometric fluorescence nano pH sensor based on silica-coated liposome nanoparticles (cerasome, 138.4 nm). The sensor integrates pH-sensitive dye, pyranine, within cerasome, achieving enhanced photostability, sensitivity, and biocompatibility. Its unique ratiometric design enables precise pH monitoring with minimal photobleaching and quenching, covering a linear detection range of pH 6.25–8.5. The hybrid nanoparticles exhibit high morphological stability, making them suitable for real-time intracellular pH measurement. This novel platform shows great promise for applications in cellular biology, disease diagnosis, and therapeutic monitoring, offering a versatile tool for biomedical research. Full article
(This article belongs to the Special Issue Nanotechnology-Based Optical Sensors for Biomedical Applications)
Show Figures

Figure 1

Figure 1
<p>Particle size and morphology characterization of nano pH sensor. (<b>a</b>) SEM image of the nano pH sensor, illustrating its structural stability and uniform morphology. The inset in the upper left corner shows the DLS measurement, confirming a consistent average particle size of 138.4 ± 4.5 nm. (<b>b</b>) TEM image of the nano pH sensor, revealing the encapsulation of the dye molecules within the liposome core and the silica shell structure surrounding the nanoparticles.</p>
Full article ">Figure 2
<p>Spectral characterization of nano pH sensor. Ultraviolet–visible absorption spectra of pyranine unloaded nanoparticles (dark), nano pH sensor (red), and free pyranine dye (blue) (<b>a</b>). Fluorescence spectra of nano pH sensor before and after dialysis (<b>b</b>).</p>
Full article ">Figure 3
<p>Excitation (<b>left</b>, Ex 350–500 nm, Em 513 nm) and emission (<b>right</b>, Ex 460 nm, Em 470–650 nm) spectra of nano pH sensor in PBS at different PH values; pH value: 5.5–8.5.</p>
Full article ">Figure 4
<p>Response curve between fluorescence ratio of nano pH sensor and pH. The fluorescence signal was monitored at 513 nm (F1 was Ex 460 nm, F2 was Ex 413 nm).</p>
Full article ">Figure 5
<p>The impact of ionic strength on the fluorescence intensity of the sensor.</p>
Full article ">Figure 6
<p>Fluorescence ratio of different pH buffers.</p>
Full article ">Figure 7
<p>Photostability of nano pH sensor (significance test: ANOVA or <span class="html-italic">t</span>-test).</p>
Full article ">Figure 8
<p>The pH-induced fluorescence ratio vs. time to evaluate the reproducibility and reversibility of the sensor.</p>
Full article ">Figure 9
<p>Cellular uptake of nano pH sensors by HeLa cells. (<b>a</b>) Optical image of HeLa cells without nano pH sensors. (<b>b</b>) Fluorescence image of HeLa cells after 3 h of incubation with nano pH sensors, showing the distribution of fluorescence within the cells. (<b>c</b>) Fluorescence image of the same cells after adding 10 µL of 0.1 M HCl for 10 s, illustrating the response of the sensors to a change in pH. (<b>d</b>) Fluorescence image after 30 s of HCl addition, highlighting further changes in fluorescence intensity. All fluorescence images were captured with the same exposure time..</p>
Full article ">Scheme 1
<p>A schematic of the preparation method of the nano pH sensor.</p>
Full article ">
32 pages, 6349 KiB  
Review
Liquid Metal–Polymer Hydrogel Composites for Sustainable Electronics: A Review
by Abdollah Hajalilou
Molecules 2025, 30(4), 905; https://doi.org/10.3390/molecules30040905 - 15 Feb 2025
Abstract
Hydrogels, renowned for their hydrophilic and viscoelastic properties, have emerged as key materials for flexible electronics, including electronic skins, wearable devices, and soft sensors. However, the application of pure double network hydrogel-based composites is limited by their poor chemical stability, low mechanical stretchability, [...] Read more.
Hydrogels, renowned for their hydrophilic and viscoelastic properties, have emerged as key materials for flexible electronics, including electronic skins, wearable devices, and soft sensors. However, the application of pure double network hydrogel-based composites is limited by their poor chemical stability, low mechanical stretchability, and low sensitivity. Recent research has focused on overcoming these limitations by incorporating conductive fillers, such as liquid metals (LMs), into hydrogel matrices or creating continuous conductive paths through LMs within the polymer matrix. LMs, including eutectic gallium and indium (EGaIn) alloys, offer exceptional electromechanical, electrochemical, thermal conductivity, and self-repairing properties, making them ideal candidates for diverse soft electronic applications. The integration of LMs into hydrogels improves conductivity and mechanical performance while addressing the challenges posed by rigid fillers, such as mismatched compliance with the hydrogel matrix. This review explores the incorporation of LMs into hydrogel composites, the challenges faced in achieving optimal dispersion, and the unique functionalities introduced by these composites. We also discuss recent advances in the use of LM droplets for polymerization processes and their applications in various fields, including tissue engineering, wearable devices, biomedical applications, electromagnetic shielding, energy harvesting, and storage. Additionally, 3D-printable hydrogels are highlighted. Despite the promise of LM-based hydrogels, challenges such as macrophase separation, weak interfacial interactions between LMs and polymer networks, and the difficulty of printing LM inks onto hydrogel substrates limit their broader application. However, this review proposes solutions to these challenges. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the LM–hydrogel fabrication process, forming conductive paths, along with their applications.</p>
Full article ">Figure 2
<p>(<b>A</b>) Schematic illustration of the fabrication process of hydrogel electronics. (<b>B</b>) The LM hydrogel composite was subjected to 150%. (<b>C</b>) Fabricated electronic circuit with LM hydrogel composite. Reproduced with permission from [<a href="#B19-molecules-30-00905" class="html-bibr">19</a>], Copyright 2018, Wiley.</p>
Full article ">Figure 3
<p>(<b>A</b>) Liquid metal particles initiate free radical polymerization without conventional initiators. (<b>a</b>) Sonication breaks the oxide-coated metal into smaller particles, exposing the metal to trigger polymerization. (<b>b</b>) This forms a physically crosslinked polyacrylamide (PAAm) hydrogel. (<b>c</b>–<b>e</b>) Schematics illustrate the process. Reproduced with permission from [<a href="#B34-molecules-30-00905" class="html-bibr">34</a>], Copyright 2019, ACS publisher. (<b>B</b>) Schematic showing the fabrication process of P-LMGO hydrogel with LMGO nanocomposite fillers. (<b>C</b>) Interactions between GO, LM, and LMGO nanocomposite fillers with polymers. Reproduced with permission from [<a href="#B68-molecules-30-00905" class="html-bibr">68</a>], Copyright 2021, Wiley-VCH.</p>
Full article ">Figure 4
<p>Schematic illustration of the fabrication process of LM hybrid hydrogel scaffolds via 3D printing and their application in infected wound treatment. Reproduced with permission from [<a href="#B77-molecules-30-00905" class="html-bibr">77</a>], Copyright 2024, Elsevier.</p>
Full article ">Figure 5
<p>(<b>A</b>) 3D-printed hydrogel milli-conical needle arrays for solar steam generation: hydrogel with a 3D conical needle architecture for omnidirectional solar harvesting and enhanced performance. Reproduced with permission from [<a href="#B79-molecules-30-00905" class="html-bibr">79</a>], Copyright 2024, ACS Publications. (<b>B</b>) Hydrogel actuator with a PNIPAM rod and liquid metal spring. (<b>a</b>) Resistive heating from the spring controls rod shrinking or bending. (<b>b</b>) The spring is fabricated using bevel-tip nozzle techniques, and the rod is molded with a 3D-printed mold. Reproduced with permission from [<a href="#B80-molecules-30-00905" class="html-bibr">80</a>], Copyright 2020, Wiley-VCH.</p>
Full article ">Figure 6
<p>(<b>A</b>) Alginate–PAAm hydrogel structure with ONB-PEG600 covalent crosslinking and Ca<sup>2+</sup> ionic alginate crosslinking. (<b>B</b>) Enhanced drying tolerance via in situ water-to-glycerol solvent displacement. (<b>C</b>) Digital and stencil printing of the LM-based ink over hydrogel substrate, where (<b>i</b>), and (<b>ii</b>) show the digital printing, and stencil printing of the composite ink over hydrogel. Reproduced with permission from [<a href="#B8-molecules-30-00905" class="html-bibr">8</a>], Copyright 2023, Wiley.</p>
Full article ">Figure 7
<p>(<b>A</b>) Schematic illustration of the fabrication process for the LMCPG hydrogel, (<b>B</b>) EMI shielding mechanism, and (<b>C</b>) EMI shielding performance. Reproduced with permission from [<a href="#B90-molecules-30-00905" class="html-bibr">90</a>], Copyright 2023, Elsevier. (<b>D</b>) Schematic preparation of the PVA–EGaInSn–Ni composite hydrogel with primary crosslinked networks. (<b>E</b>) EMI shielding mechanism of the LM-based hydrogel, attributed to conductive loss, interfacial polarization, and dipole polarization. (<b>F</b>) EMI shielding performance. Reproduced with permission from [<a href="#B95-molecules-30-00905" class="html-bibr">95</a>], Copyright 2023, Springer.</p>
Full article ">Figure 7 Cont.
<p>(<b>A</b>) Schematic illustration of the fabrication process for the LMCPG hydrogel, (<b>B</b>) EMI shielding mechanism, and (<b>C</b>) EMI shielding performance. Reproduced with permission from [<a href="#B90-molecules-30-00905" class="html-bibr">90</a>], Copyright 2023, Elsevier. (<b>D</b>) Schematic preparation of the PVA–EGaInSn–Ni composite hydrogel with primary crosslinked networks. (<b>E</b>) EMI shielding mechanism of the LM-based hydrogel, attributed to conductive loss, interfacial polarization, and dipole polarization. (<b>F</b>) EMI shielding performance. Reproduced with permission from [<a href="#B95-molecules-30-00905" class="html-bibr">95</a>], Copyright 2023, Springer.</p>
Full article ">Figure 8
<p>(<b>A</b>) Schematic of the rapid assembly process of LMGAs-FeIII, highlighting interface and bulk species formation (<b>a</b>). Digital images showing the progression of the assembly in the LMGAs dispersion (<b>b</b>–<b>d</b>). IR images of the stable temperature of LMGAs-FeIII solution and DI water under one sun irradiation (<b>e</b>). Reproduced with permission from [<a href="#B100-molecules-30-00905" class="html-bibr">100</a>], Copyright 2023, Wiley. (<b>B</b>) Thermoelectric performance of the EGaIn@Ag–PVAG hydrogel: (<b>a</b>) Photograph of the SDTC device; (<b>b</b>) Infrared thermal images and temperature increase profile of the hydrogel under solar irradiation. Reproduced with permission from [<a href="#B98-molecules-30-00905" class="html-bibr">98</a>], Copyright 2024, Elsevier. (<b>C</b>) Patternable electroluminescent devices. Reproduced with permission from [<a href="#B101-molecules-30-00905" class="html-bibr">101</a>], Copyright 2023, Elsevier.</p>
Full article ">Figure 9
<p>(<b>A</b>) Application of gelatin–alginate–glycerol hydrogel with patterned LM and sensing units for electronic skin, along with the resulting data. Reproduced with permission from [<a href="#B105-molecules-30-00905" class="html-bibr">105</a>], Copyright 2022, Wiley. (<b>B</b>) Electrode placement and EMG signal processing. Reproduced with permission from [<a href="#B78-molecules-30-00905" class="html-bibr">78</a>], Copyright 2020, ACS Publications.</p>
Full article ">Figure 10
<p>(<b>A</b>,<b>B</b>) The endovascular embolization process of the LM–CA hydrogel. Micro-CT and 3D CT images show LM–CA hydrogels in a rat femoral vein ((<b>C</b>,<b>E</b>) red circles) compared to the control (<b>D</b>,<b>F</b>). LM–SA mixture solution (LM:SA = 1:2) is shown in a photograph and X-ray ((<b>G</b>,<b>H</b>) scale bar: 1 cm) and in pig heart blood vessels before and after injection ((<b>I</b>–<b>K</b>) scale bar: 5 cm). Reproduced with permission from [<a href="#B108-molecules-30-00905" class="html-bibr">108</a>], Copyright 2020, Wiley.</p>
Full article ">Figure 11
<p>Schematic of the hydrogel composition with lipid-coated LM particles and an anticancer drug, including drug-conjugated gelatin, PEGDA, and LM particles for photothermal therapy. Reproduced with permission from [<a href="#B110-molecules-30-00905" class="html-bibr">110</a>], Copyright 2024, Elsevier.</p>
Full article ">Figure 12
<p>(<b>A</b>) Schematic illustration of LM–PVA hydrogel, (<b>B</b>) a schematic of LP-TENG structure, and (<b>C</b>) repeatable voltage outputs of handwriting recognition based on LP-TENG utilized for handwriting authentication. Reproduced with permission from [<a href="#B96-molecules-30-00905" class="html-bibr">96</a>], Copyright 2024, Springer.</p>
Full article ">
19 pages, 19542 KiB  
Article
A Programmable Gain Amplifier Featuring a High Power Supply Rejection Ratio for a 20-Bit Sigma-Delta ADC
by Wenhui Li, Daishi Tian, Hao Zhu and Qingqing Sun
Electronics 2025, 14(4), 720; https://doi.org/10.3390/electronics14040720 - 12 Feb 2025
Abstract
A programmable gain amplifier (PGA) is commonly used to optimize the input dynamic range of high-performance systems such as headphones and biomedical sensors. But PGA is rather sensitive to electromagnetic interference (EMI), which limits the precision of these systems. Many capacitor-less low-dropout regulator [...] Read more.
A programmable gain amplifier (PGA) is commonly used to optimize the input dynamic range of high-performance systems such as headphones and biomedical sensors. But PGA is rather sensitive to electromagnetic interference (EMI), which limits the precision of these systems. Many capacitor-less low-dropout regulator (LDO) schemes with high power supply rejection have been proposed to act as the independent power supply for PGA, which consumes additional power and area. This paper proposed a PGA with a high power supply rejection ratio (PSRR) and low power consumption, which serves as the analog front-end amplifier in the 20-bit sigma-delta ADC. The PGA is a two-stage amplifier with hybrid compensation. The first stage is the recycling folded cascode amplifier with the gain-boost technique, while the second stage is the class-AB output stage. The PGA was implemented in the 0.18 μm CMOS technology and achieved a 9.44 MHz unity-gain bandwidth (UGBW) and a 57.8° phase margin when driving the capacitor of 5.9 pF. An optimum figure-of-merit (FoM) value of 905.67 has been achieved with the proposed PGA. As the front-end amplifier of a high-precision ADC, it delivers a DC gain of 162.1 dB, the equivalent input noise voltage of 301.6 nV and an offset voltage of 1.61 μV. Within the frequency range below 60 MHz, the measured PSRR of ADC is below −70 dB with an effective number of bits (ENOB), namely 20 bits. Full article
Show Figures

Figure 1

Figure 1
<p>Block diagram of a 20-bit sigma-delta ADC with the front-end PGA.</p>
Full article ">Figure 2
<p>Proposed two-stage amplifier (RFC1) using hybrid compensation.</p>
Full article ">Figure 3
<p>An AC small-signal model for RFC1.</p>
Full article ">Figure 4
<p>The final proposed two-stage amplifier (RFC2) applied in the 20-bit sigma-delta ADC: (<b>a</b>) the detailed amplifier circuit with the gain-boost and chopper circuits; (<b>b</b>) the CMOS switch used in the input end; (<b>c</b>) the CH<sub>p</sub> and CH<sub>n</sub> used in the amplifier and the non-overlapping clock at the gate.</p>
Full article ">Figure 4 Cont.
<p>The final proposed two-stage amplifier (RFC2) applied in the 20-bit sigma-delta ADC: (<b>a</b>) the detailed amplifier circuit with the gain-boost and chopper circuits; (<b>b</b>) the CMOS switch used in the input end; (<b>c</b>) the CH<sub>p</sub> and CH<sub>n</sub> used in the amplifier and the non-overlapping clock at the gate.</p>
Full article ">Figure 5
<p>Die micrographs of RFC1 and RFC2: (<b>a</b>) RFC1; (<b>b</b>) PGA consisting of RFC2 in the 20-bit sigma-delta ADC.</p>
Full article ">Figure 6
<p>Frequency characteristics of amplifiers (RFC1 and RFC2): (<b>a</b>) AC frequency response of RFC1 and RFC2; (<b>b</b>) the variation in UGBW of RFC1 and RFC2 with temperature.</p>
Full article ">Figure 7
<p>The large-signal step response of RFC1.</p>
Full article ">Figure 8
<p>PSRR simulation: (<b>a</b>) PSRR simulation schematic diagram; (<b>b</b>) PSRR simulation results of RFC1 and RFC2.</p>
Full article ">Figure 9
<p>The noise spectral power density of RFC1 and RFC2: (<b>a</b>) the noise spectral power density of RFC1 without the chopper circuit; (<b>b</b>) the noise spectral power density of RFC2 with the chopper circuit on and down.</p>
Full article ">Figure 10
<p>The Monte Carol simulation results of RFC1 and RFC2 under different conditions: (<b>a</b>) the process simulation result of RFC1; (<b>b</b>)the mismatch simulation result of RFC1; (<b>c</b>) the process simulation result of RFC2 with the chopper circuit on; (<b>d</b>) the mismatch simulation result of RFC2 with the chopper circuit on; (<b>e</b>) the process simulation result of RFC2 with the chopper circuit off; (<b>f</b>) the mismatch simulation result of RFC2 with the chopper circuit off.</p>
Full article ">Figure 11
<p>The measurement circuit of GBW.</p>
Full article ">Figure 12
<p>The UGBW testing result of RFC1.</p>
Full article ">Figure 13
<p>The measurement circuit of slew rate.</p>
Full article ">Figure 14
<p>The measured result of slew rate.</p>
Full article ">Figure 15
<p>PSRR measurement: (<b>a</b>) the measurement system diagram of PSRR; (<b>b</b>) the measured result of PSRR. The peak-to-peak values of the sine waves applied to AVDD are 33 mV and 66 mV, respectively.</p>
Full article ">Figure 16
<p>The comprehensive comparison results of FoM<sub>s</sub> and DC gain with prior works [<a href="#B6-electronics-14-00720" class="html-bibr">6</a>,<a href="#B16-electronics-14-00720" class="html-bibr">16</a>,<a href="#B23-electronics-14-00720" class="html-bibr">23</a>,<a href="#B24-electronics-14-00720" class="html-bibr">24</a>,<a href="#B25-electronics-14-00720" class="html-bibr">25</a>].</p>
Full article ">
44 pages, 9682 KiB  
Review
Mid-Infrared Photonic Sensors: Exploring Fundamentals, Advanced Materials, and Cutting-Edge Applications
by Muhammad A. Butt, Marcin Juchniewicz, Mateusz Słowikowski, Łukasz Kozłowski and Ryszard Piramidowicz
Sensors 2025, 25(4), 1102; https://doi.org/10.3390/s25041102 - 12 Feb 2025
Abstract
Mid-infrared (MIR) photonic sensors are revolutionizing optical sensing by enabling precise chemical and biological detection through the interrogation of molecules’ unique vibrational modes. This review explores the core principles of MIR photonics, emphasizing the light–matter interactions within the 2–20 µm wavelength range. Additionally, [...] Read more.
Mid-infrared (MIR) photonic sensors are revolutionizing optical sensing by enabling precise chemical and biological detection through the interrogation of molecules’ unique vibrational modes. This review explores the core principles of MIR photonics, emphasizing the light–matter interactions within the 2–20 µm wavelength range. Additionally, it examines innovative sensor architectures, such as integrated photonic platforms and optical fibers, that enhance sensitivity, specificity, and device miniaturization. The discussion extends to groundbreaking applications in environmental monitoring, medical diagnostics, industrial processes, and security, highlighting the transformative impact of these technologies. This comprehensive overview aims to illuminate the current state-of-the-art while inspiring future developments in MIR photonic sensing. Full article
(This article belongs to the Special Issue New Trends and Progress in Plasmonic Sensors and Sensing Technology)
Show Figures

Figure 1

Figure 1
<p>A structured outline illustrates the paper’s content’s logical flow and progression.</p>
Full article ">Figure 2
<p>An example of an MIR waveguide on an SOI platform demonstrates propagation losses ranging from 2 to 3 dB/cm across the broad MIR spectrum of 3.68–3.88 μm. The bending losses were as low as 0.02 dB per 90° turn for radii larger than 10 μm. The figure includes (<b>a</b>) a schematic of the single-mode SOI channel waveguide, (<b>b</b>) SEM image of the straight waveguide with an inset of the inverse taper tip, and (<b>c</b>) optical images of waveguide bends with a radius of 5 μm [<a href="#B65-sensors-25-01102" class="html-bibr">65</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) A contour plot of the optical mode of the SOS waveguide, (<b>b</b>) a false-color scanning electron micrograph of the cleaved end facet of a waveguide. Silicon is shown in green and sapphire in blue [<a href="#B79-sensors-25-01102" class="html-bibr">79</a>].</p>
Full article ">Figure 4
<p>SEM image of the GOS waveguide with the 3 μm thickness of Ge slot waveguide. Experimental setup for measuring the propagation loss of waveguides [<a href="#B87-sensors-25-01102" class="html-bibr">87</a>].</p>
Full article ">Figure 5
<p>Schematics of a suspended germanium waveguide with subwavelength-grating metamaterial lateral cladding. (<b>a</b>) 3D view. (<b>b</b>) Front view (<b>c</b>) Top view of the guiding layer [<a href="#B88-sensors-25-01102" class="html-bibr">88</a>].</p>
Full article ">Figure 6
<p>The process for fabricating suspended germanium waveguides with subwavelength grating lateral cladding involves multiple steps. (<b>a</b>) Dry etching is used to transfer the cladding holes from the photoresist to the germanium layer and silicon film. (<b>b</b>) The silicon dioxide and remaining silicon film are etched away using a wet etching technique. (<b>c</b>) The process concludes with the complete removal of the silicon film and buried oxide [<a href="#B88-sensors-25-01102" class="html-bibr">88</a>].</p>
Full article ">Figure 7
<p>Waveguide propagation losses in various material platforms [<a href="#B107-sensors-25-01102" class="html-bibr">107</a>].</p>
Full article ">Figure 8
<p>Cross-sectional schematics of the active regions for four different DFB QCL designs integrated on the SONOI waveguide [<a href="#B116-sensors-25-01102" class="html-bibr">116</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) Precipitate of the contaminant (Sample A) retained after filtering the white sugar solution [<a href="#B156-sensors-25-01102" class="html-bibr">156</a>], (<b>b</b>) statistical analysis of the FT-MIR spectra for contaminants in white sugar (samples analyzed: 3; total scans: 96). For each data point, the average represents the arithmetic mean of Y values, variance denotes the standard deviation of Y values, and range indicates the margin of Y values [<a href="#B156-sensors-25-01102" class="html-bibr">156</a>].</p>
Full article ">Figure 10
<p>Key components of the sensor design: (<b>a</b>) the cross-sectional view of the slot waveguide [<a href="#B161-sensors-25-01102" class="html-bibr">161</a>], (<b>b</b>) the sensor chip layout featuring double-tip couplers, (<b>c</b>) simulation results showing the air confinement factor (solid lines) and substrate leakage loss (dotted lines) as functions of slot width for various strip widths, and (<b>d</b>) a schematic of the double-tip coupler, including simulated mode profiles at the coupler facet and slot waveguide cross-section [<a href="#B161-sensors-25-01102" class="html-bibr">161</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Diagram of the on-chip PTS setup with a modulated pump, where the phase of the probe beam is altered due to pump absorption. The inset displays the C<sub>2</sub>H<sub>2</sub> absorption cross-section at the pump and probe wavelengths [<a href="#B168-sensors-25-01102" class="html-bibr">168</a>]. (<b>b</b>) Illustration of the processes responsible for PT-induced phase modulation in the waveguide [<a href="#B168-sensors-25-01102" class="html-bibr">168</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) Diagram of the proposed multivariable sensor, showing the chemiresistive-potentiometric sensing region. (<b>b</b>) Chemiresistive and (<b>c</b>) potentiometric (relative to Pt CE) response times for various 2-EH concentrations. (<b>d</b>) Chemiresistive and (<b>e</b>) potentiometric responses to 2-EH, ethanol, acetone, toluene, NH<sub>3</sub>, CO, and NO<sub>2</sub> at different concentrations. The sensor operates at 400 °C [<a href="#B187-sensors-25-01102" class="html-bibr">187</a>].</p>
Full article ">Figure 13
<p>(<b>a</b>) Diagram of the suspended Si waveguide gas sensing platform, which includes grating couplers, tapers, a Y-junction power splitter, and spiral waveguides. Inset: Cross-sectional SEM image of the waveguide [<a href="#B194-sensors-25-01102" class="html-bibr">194</a>]. (<b>b</b>) Optical image of the suspended Si spiral waveguide [<a href="#B194-sensors-25-01102" class="html-bibr">194</a>]. (<b>c</b>) Close-up view of the sensing waveguide surrounded by toluene molecules, highlighted by the yellow box in (<b>a</b>) [<a href="#B194-sensors-25-01102" class="html-bibr">194</a>]. (<b>d</b>) Schematic of suspended slot membrane waveguide based on Ge-on-SOI platform for CO<sub>2</sub> gas detection. The inset shows the E-field distribution at an operational wavelength of 4.23 µm [<a href="#B195-sensors-25-01102" class="html-bibr">195</a>].</p>
Full article ">Figure 14
<p>(<b>a</b>) Schematic of a DLSPP waveguide [<a href="#B196-sensors-25-01102" class="html-bibr">196</a>]. (<b>b</b>) Cross-sectional view and (<b>c</b>) top view of the E-field distribution in the waveguide [<a href="#B196-sensors-25-01102" class="html-bibr">196</a>]. (<b>d</b>) 3D representation and (<b>e</b>) cross-sectional view of MLGMT waveguide [<a href="#B200-sensors-25-01102" class="html-bibr">200</a>].</p>
Full article ">Figure 15
<p>Various configurations of the probe’s sensing zone: (<b>a</b>) U-shaped; (<b>b</b>) single loop; (<b>c</b>) double loops. (<b>d</b>) Configuration for fiber-optic evanescent wave spectroscopy (FEWS) analysis: 1—IR Fourier spectrometer; 2—ZnSe focusing lenses; 3—chalcogenide fiber with polymer coating; 4—uncoated fiber sensing segment; 5—container with the liquid sample; 6—HgCdTe detector; 7—amplifier. (<b>e</b>) Image of the experimental setup used for in situ sensor testing. Absorption spectra of isopropyl alcohol in water–alcohol mixtures: 1—pure water; 2—0.1% vol. isopropyl alcohol; 3—0.5% vol.; 4—1.0% vol.; 5—2.0% vol.; 6—3.0% vol.; 7—4.0% vol.; 8—5.0% vol [<a href="#B212-sensors-25-01102" class="html-bibr">212</a>].</p>
Full article ">Figure 16
<p>MIR spectrum with marked specific molecules absorption regions [<a href="#B213-sensors-25-01102" class="html-bibr">213</a>].</p>
Full article ">Figure 17
<p>A demonstration of IRFlex’s Laser-Based IRCM technology.</p>
Full article ">
75 pages, 13093 KiB  
Review
Review on Advancements in Carbon Nanotubes: Synthesis, Purification, and Multifaceted Applications
by Anil Kumar Madikere Raghunatha Reddy, Ali Darwiche, Mogalahalli Venkatashamy Reddy and Karim Zaghib
Batteries 2025, 11(2), 71; https://doi.org/10.3390/batteries11020071 - 8 Feb 2025
Abstract
Since their discovery over two decades ago, carbon nanotubes (CNTs) have captivated researchers due to their exceptional electrical, optical, mechanical, and thermal properties, making them versatile candidates for various advanced applications. CNTs have transformed numerous scientific domains, including nanotechnology, electronics, materials science, and [...] Read more.
Since their discovery over two decades ago, carbon nanotubes (CNTs) have captivated researchers due to their exceptional electrical, optical, mechanical, and thermal properties, making them versatile candidates for various advanced applications. CNTs have transformed numerous scientific domains, including nanotechnology, electronics, materials science, and biomedical engineering. Their applications range from nanoelectronics, robust nanocomposites, and energy storage devices to innovative materials, sensors, conducting polymers, field emission sources, and Li-ion batteries. Furthermore, CNTs have found critical roles in biosensing, water purification, bone scaffolding, and targeted gene and drug delivery. The chemical reactivity and functional versatility of CNTs are profoundly influenced by their structural and physicochemical properties, such as surface area, surface charge, size distribution, surface chemistry, and purity. This review comprehensively explores the current state of CNT research, focusing on widely used synthesis, purification, and characterization techniques alongside emerging applications. By highlighting recent advancements and addressing unresolved challenges, it aims to present a novel perspective on the transformative potential of CNTs, fostering innovation across diverse scientific and technological fields. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Wrapping of graphene sheet to form (<b>a</b>) SWCNTs, (<b>b</b>) DWCNTs, and (<b>c</b>) MWCNTs. Schematic illustrations of SWCNT structures, reproduced with permission [<a href="#B5-batteries-11-00071" class="html-bibr">5</a>]. (<b>d</b>) Zigzag arrangement, (<b>e</b>) armchair configuration, and (<b>f</b>) chiral conformation, reproduced with permission [<a href="#B4-batteries-11-00071" class="html-bibr">4</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) An arc discharge setup. Adopted with permission from [<a href="#B36-batteries-11-00071" class="html-bibr">36</a>]. Copyright © 2014 Elsevier. (<b>b</b>) A laser ablation setup. Adopted with permission from [<a href="#B74-batteries-11-00071" class="html-bibr">74</a>]. (<b>c</b>) Chemical vapor deposition method. Adopted with permission from [<a href="#B74-batteries-11-00071" class="html-bibr">74</a>].</p>
Full article ">Figure 3
<p>Strategy to measure and optimize selectivity and sensitivity of neurotransmitter sensors. Candidate sensors are synthesized from single-walled carbon nanotubes (SWCNTs) and DNA oligonucleotides, and their responses to neurotransmitters such as dopamine, epinephrine, and norepinephrine are quantified. Crucial for the success of these sensors is the discrimination between different but chemically very similar neurotransmitters, Adopted with permission from [<a href="#B184-batteries-11-00071" class="html-bibr">184</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic of deposition process: a substrate wire wetted with alcohol/PVA solution is fed to the reactor for deposition of CNT assembly and finally wound up. (<b>b</b>) Vinylon wire is used as a substrate for depositing the first layer of CNTs. (<b>c</b>) Hollow cylindrical CNT assembly shrinking and depositing on the substrate wire wetted with alcohol/PVA solution. (<b>d</b>) CNT/PVA composite fiber wound on a glass tube. Adopted with permission from [<a href="#B326-batteries-11-00071" class="html-bibr">326</a>]. (<b>e</b>) The proposed atom hydrogen spillover mechanism of SWCNTs. Adopted with permission from [<a href="#B339-batteries-11-00071" class="html-bibr">339</a>]. (<b>f</b>) Schematic of DSSC with a pair of CNTs-based electrodes. Adopted with permission from [<a href="#B376-batteries-11-00071" class="html-bibr">376</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) CNTs-based supercapacitors. Adopted with permission from [<a href="#B263-batteries-11-00071" class="html-bibr">263</a>]. (<b>b</b>) The specific capacitance of the supercapacitor textile depends on the strain. C<sub>0</sub> and C correspond to specific capacitances at 0 and the other strains, respectively. (<b>c</b>,<b>d</b>) Photographs of a flexible and stretchable textile before and after stretching by 100%, respectively. Adopted with permission from [<a href="#B392-batteries-11-00071" class="html-bibr">392</a>]. (<b>e</b>) Photograph of balsam pear with the stem tendril coiled in a helix around a thin rod; inset is a close-up image showing the helical structure. (<b>f</b>) Demonstration of SC fabrication process of ultra-flexible wire-shaped supercapacitor. Adopted with permission from [<a href="#B393-batteries-11-00071" class="html-bibr">393</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Galvanostatic charge/discharge curves of the CGM-SC under strains ranging from 0 to 850%. (<b>b</b>) Evolution of the specific capacitance ratio over several stretching–releasing cycles with a 700% strain. The current density for all tests was 0.5 A/cm<sup>3</sup>. The inset is a schematic illustration showing the stretching and the release of the CGM-SC. (<b>c</b>) Photographs of the CGM-SC with tensile strains of 0, 100%, 300%, 500%, 700%, and 850%, respectively. The insets show the light emission diode powered by the CGM-SC. Adopted with permission from [<a href="#B393-batteries-11-00071" class="html-bibr">393</a>]. (<b>d</b>) An illustration of a lithium-ion battery’s working mechanism. Adopted with permission from [<a href="#B432-batteries-11-00071" class="html-bibr">432</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Effect of defects on lithium insertion in a (5, 5) SWCNT. Note: For each nanotube structure, both side and top views are provided to illustrate the impact of defects on lithium-ion insertion. Adopted with permission from [<a href="#B435-batteries-11-00071" class="html-bibr">435</a>]. (<b>b</b>) Purified and 60 min ball-milled MWNT reversible capacity and (<b>c</b>) coulombic efficiency as a function of cycle count. Adopted with permission from [<a href="#B448-batteries-11-00071" class="html-bibr">448</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) The fluctuation of the Li/C ratio with tube diameter is shown. The insets display the equilibrium configurations of SWNTs packed with Li atoms. In the visualization, Li and carbon atoms are represented by white and grey spheres, respectively. Adopted with permission from [<a href="#B460-batteries-11-00071" class="html-bibr">460</a>]. (<b>b</b>) The 1st and (<b>c</b>) the 50th discharge and charge curves for the as-prepared MWCNT electrodes at a current density of 50 mA/g and potential window of 0–2 V. Adopted with permission from [<a href="#B461-batteries-11-00071" class="html-bibr">461</a>].</p>
Full article ">Figure 9
<p>SEM (<b>a</b>) and TEM (<b>b</b>) micrographs of the as-synthesized PPy; SEM (<b>c</b>) and TEM (<b>d</b>) image of AB@PPy composites. Adopted with permission from [<a href="#B467-batteries-11-00071" class="html-bibr">467</a>].</p>
Full article ">Figure 10
<p>Impact of the areal density on the specific capacity at 5 A/g, Adopted with permission from [<a href="#B476-batteries-11-00071" class="html-bibr">476</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) The fabrication procedure for the 3DC-Sn@CNT composite. (<b>b</b>,<b>c</b>) TEM images of the 3DC-SnO<sub>2</sub> composite. (<b>d</b>–<b>h</b>) TEM images and (<b>i</b>) element mapping of the 3DC-Sn@CNT composite. Adopted with permission from [<a href="#B477-batteries-11-00071" class="html-bibr">477</a>].</p>
Full article ">Figure 12
<p>Rate performance of Asnano reduced on MWCNTs (squares, 70:17:13 As MWCNT:PAA <span class="html-italic">w</span>:<span class="html-italic">w</span>, loaded at C/10, unloaded at the specified rate) and Asnano reduced on MWCNTs (circles, 70:17:13 As MWCNT:PAA <span class="html-italic">w</span>:<span class="html-italic">w</span>, Symmetric Load/Unload) cycled in 1:1 FEC: DEC <span class="html-italic">w</span>:<span class="html-italic">w</span> with 1M LiPF6. Lines between data points are to guide the eye. Adopted with permission from [<a href="#B479-batteries-11-00071" class="html-bibr">479</a>].</p>
Full article ">Figure 13
<p>(<b>a</b>) Cyclic voltammograms at a sweep rate of 0.1 mV/s, (<b>b</b>) voltage profiles at 2C, (<b>c</b>) galvanostatic discharging up to 500C, (<b>d</b>) areal power density for discharging, (<b>e</b>) areal energy density for discharging, (<b>f</b>) long-term stability test at 10C. Adopted with permission from [<a href="#B495-batteries-11-00071" class="html-bibr">495</a>].</p>
Full article ">Figure 14
<p>(<b>A</b>) EIS spectra of the LTO (a), LTO-CNT (b), LTO-G (c), and LTO-CNT-G (d) electrodes with corresponding equivalent circuit model (inset) [<a href="#B496-batteries-11-00071" class="html-bibr">496</a>] (<b>B</b>) Schematic diagram of the network structure in the LTO-CNT-G composite. Adopted with permission from [<a href="#B496-batteries-11-00071" class="html-bibr">496</a>]. Copyright © 2016 Elsevier. (<b>C</b>) Schematic of the forming process of CNT-entangled Mn<sub>3</sub>O<sub>4</sub> octahedrons. Adopted with permission from [<a href="#B488-batteries-11-00071" class="html-bibr">488</a>].</p>
Full article ">Figure 15
<p>Electrochemical properties: (<b>a</b>) charge/discharge profiles at 0.2 C, inset: the magnified selected-region comparing the potential plateaus, (<b>b</b>) rate capabilities from 0.2 to 60 C, (<b>c</b>) cycling stabilities combined with Coulombic efficiency at 10 C of C@LFP and C@LFP/CNTs, and (<b>d</b>) charge current and capacity as well as charge efficiency of C@LFP/CNTs as a function of charge time. Adopted with permission from [<a href="#B535-batteries-11-00071" class="html-bibr">535</a>]. Copyright © 2016 WILEY—VCH Verlag GmbH &amp; Co.</p>
Full article ">Figure 16
<p>Schematic description of three-dimensional conductive network of monodispersed CNTs. (<b>a</b>) PVP-coated CNTs; (<b>b</b>) three-dimensional conductive network with CNTs and acetylene black. Adopted with permission from [<a href="#B500-batteries-11-00071" class="html-bibr">500</a>]. Copyright © 2016 Elsevier.</p>
Full article ">Figure 17
<p>(<b>a</b>) Discharge capability vs. charge (1 C)/discharge (10 C) cycle; (<b>b</b>) capacity-cycle relation of the LiFePO<sub>4</sub>/CNTs(3 wt%)/AB(7 wt%) and commercial LiFePO<sub>4</sub>/AB(10 wt%); (<b>c</b>) coulombic efficiency vs. cycle number of the LiFePO<sub>4</sub>/CNTs(3 wt%)/AB(7 wt%). Adopted with permission from [<a href="#B500-batteries-11-00071" class="html-bibr">500</a>]. (<b>d</b>) Energy densities (based on the total mass of the electrode material layer and the current collector) of the LiCoO<sub>2</sub>-2 wt% Super P-SACNT and the sandwich-structured LiCoO<sub>2</sub>-5 wt% Super P electrodes (of the same thickness). Adopted with permission from [<a href="#B512-batteries-11-00071" class="html-bibr">512</a>].</p>
Full article ">Figure 18
<p>(<b>a</b>) Cross-sectional SEM image of the three-layer LiCoO<sub>2</sub>—Super P—SACNT cathode after 50 cycles. Adopted with permission from [<a href="#B512-batteries-11-00071" class="html-bibr">512</a>]. (<b>b</b>) Cycling performance at 0.25 C rate. (<b>c</b>) Rate performance of pristine NCA and NCA/CNT (10 wt%) composite. Adopted with permission from [<a href="#B510-batteries-11-00071" class="html-bibr">510</a>].</p>
Full article ">Figure 19
<p>(<b>a</b>) A diagram illustrating the creation of a sheet-structured V<sub>2</sub>O<sub>5</sub>–CNT nanocomposite via ice-templating using a “brick-and-mortar” assembly method. The blue tubes symbolize CNTs while the pink ribbons denote V<sub>2</sub>O<sub>5</sub>. Adopted with permission from [<a href="#B506-batteries-11-00071" class="html-bibr">506</a>]. Rate capability of 3D Li<sub>2</sub>FeSiO<sub>4</sub>–CNT at various current rates: (<b>b</b>) Galvanostatic charge and discharge curves at various rates. (<b>c</b>) Rate cycling performance at various rates and a fixed rate of 1 C. Adopted with permission from [<a href="#B517-batteries-11-00071" class="html-bibr">517</a>].</p>
Full article ">Figure 20
<p>Schematic illustration of (<b>a</b>) the organic modification process of carbon nanotubes and (<b>b</b>) the cell configuration of lithium-sulfur (Li-S) batteries. Adopted with permission from [<a href="#B572-batteries-11-00071" class="html-bibr">572</a>].</p>
Full article ">Figure 21
<p>(<b>a</b>) Cycling performance of the blank separator and CNTs-coated separator at 0.1 C; (<b>b</b>) cycling performance of the CNTs-coated separator at the rates of 1 C and 2 C; (<b>c</b>) rate performance of the Li-S batteries with the CNTs-coated separator; (<b>d</b>) EIS data of the cells with the blank separator and CNTs-coated separator. Adopted with permission from [<a href="#B572-batteries-11-00071" class="html-bibr">572</a>].</p>
Full article ">Figure 22
<p>(<b>a</b>) Discharge/charge profiles of the Li-O<sub>2</sub> batteries with pristine-CNTs, Mo<sub>2</sub>C, and Mo<sub>2</sub>C/CNT electrodes at a discharge capacity of 500 mAh/g<sub>total</sub> and a current density of 100 mAh/g<sub>total</sub>. (<b>b</b>) Cycling performance of the LiO<sub>2</sub> battery with the Mo<sub>2</sub>C/carbon nanotube electrode at a discharge capacity of 500 mAh/g<sub>total</sub> and at a current density of 100 mAh/g<sub>total</sub>. (<b>c</b>) The first 10 cycles of Li-O<sub>2</sub> batteries with the Mo<sub>2</sub>C/carbon nanotube electrode at current densities of 100, 200, 500, and 1000 mAh/g<sub>total</sub> with a discharge capacity of 500 mAh/g<sub>total</sub>. (<b>d</b>) Cycling performance of the Li-O<sub>2</sub> battery with the Mo<sub>2</sub>C/carbon nanotube electrode at a discharge capacity of 1000 mAh/g<sub>total</sub> and at a current density of 200 mAh/g<sub>total</sub>. Adopted with permission from [<a href="#B587-batteries-11-00071" class="html-bibr">587</a>].</p>
Full article ">
17 pages, 595 KiB  
Article
A Comparative Study Between Micro and Millimeter Impedance Sensor Designs for Type-2 Diabetes Detection
by Santu Guin, Debjyoti Chowdhury and Madhurima Chattopadhyay
Micro 2025, 5(1), 7; https://doi.org/10.3390/micro5010007 - 1 Feb 2025
Abstract
In recent years, various types of sensors have been developed at both millimeter (mm) and micrometer (µm) scales for numerous biomedical applications. Each design has its own advantages and limitations. This study compares the electrical characteristics and sensitivity of millimeter- and micrometer-scale sensors, [...] Read more.
In recent years, various types of sensors have been developed at both millimeter (mm) and micrometer (µm) scales for numerous biomedical applications. Each design has its own advantages and limitations. This study compares the electrical characteristics and sensitivity of millimeter- and micrometer-scale sensors, emphasizing the superior performance of millimeter-scale designs for detecting type-2 diabetes. Elevated glucose levels in type-2 diabetes alter the complex permittivity of red blood cells (RBCs), affecting their rheological and electrical properties, such as viscosity, volume, relative permittivity, dielectric loss, and AC conductivity. These alterations may manifest as a unique bio-impedance signature, offering a diagnostic topology for diabetes. In view of this, various concentrations (ranging from 10% to 100%) of 400 µL of normal and diabetic RBCs suspended in phosphate-buffered saline (PBS) solution are examined to record the changes in bio-impedance signatures across a spectrum of frequencies, ranging from 1 MHz to 10 MHz. In this study, simulations are performed using the finite element method (FEM) with COMSOL Multiphysics® to analyze the electrical behavior of the sensors at both millimeter (mm) and micrometer (µm) scales. These simulations provide valuable insights into the performance parameters of the sensors, aiding in the selection of the most effective design by using this topology. Full article
Show Figures

Figure 1

Figure 1
<p>Electrode layout with millimeter (mm) dimension with (<b>a</b>) top view; (<b>b</b>) cross-section view.</p>
Full article ">Figure 2
<p>Electrode layout with micrometer (µm) dimension with (<b>a</b>) top view; (<b>b</b>) cross-section view [<a href="#B53-micro-05-00007" class="html-bibr">53</a>,<a href="#B55-micro-05-00007" class="html-bibr">55</a>].</p>
Full article ">Figure 3
<p>Variation in the permittivity of the cell–medium with an increasing number of cells [<a href="#B53-micro-05-00007" class="html-bibr">53</a>].</p>
Full article ">Figure 4
<p>Surface-charge density in diabetic RBC with mm electrodes.</p>
Full article ">Figure 5
<p>Surface-charge density of diabetic RBC with mm electrodes (Y-Z direction).</p>
Full article ">Figure 6
<p>Surface-charge density of diabetic RBC with µm electrodes.</p>
Full article ">Figure 7
<p>Variation of the double-layer impedance of IDE within 1 MHz–10 MHz using electrodes with mm dimension and µm dimension.</p>
Full article ">Figure 8
<p>Variation in the average complex impedance (<math display="inline"><semantics> <msub> <mi>Z</mi> <mrow> <mi>mix</mi> </mrow> </msub> </semantics></math>) over the whole concentration range within 1–100 MHz frequency range for <span class="html-italic">N</span> = 40 (in mm dimension).</p>
Full article ">Figure 9
<p>Variation in the average complex impedance (<math display="inline"><semantics> <msub> <mi>Z</mi> <mrow> <mi>mix</mi> </mrow> </msub> </semantics></math>) over the whole concentration range within 1–100 MHz frequency range for <span class="html-italic">N</span> = 40 (in µm dimension).</p>
Full article ">Figure 10
<p>Standard deviation (SD) of <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mrow> <mi>mix</mi> <mspace width="4.pt"/> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="normal">k</mi> <mo>Ω</mo> <mo>)</mo> </mrow> </mrow> </semantics></math> for normal and diabetic RBCs over the whole range of concentration by varying the frequency range within 1–10 MHz using electrodes with mm dimension.</p>
Full article ">Figure 11
<p>Standard deviation (SD) of <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mrow> <mi>mix</mi> <mspace width="4.pt"/> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="normal">k</mi> <mo>Ω</mo> <mo>)</mo> </mrow> </mrow> </semantics></math> for normal and diabetic RBCs over the whole range of concentration by varying the frequency range 1–10 MHz using electrodes with µm dimension.</p>
Full article ">
10 pages, 3072 KiB  
Communication
Acoustic Sensing Fiber Coupled with Highly Magnetostrictive Ribbon for Small-Scale Magnetic-Field Detection
by Zach Dejneka, Daniel Homa, Logan Theis, Anbo Wang and Gary Pickrell
Sensors 2025, 25(3), 841; https://doi.org/10.3390/s25030841 - 30 Jan 2025
Abstract
Fiber-optic sensing has shown promising development for use in detecting magnetic fields for downhole and biomedical applications. Coupling existing fiber-based strain sensors with highly magnetostrictive materials allows for a new method of magnetic characterization capable of distributed and high-sensitivity field measurements. This study [...] Read more.
Fiber-optic sensing has shown promising development for use in detecting magnetic fields for downhole and biomedical applications. Coupling existing fiber-based strain sensors with highly magnetostrictive materials allows for a new method of magnetic characterization capable of distributed and high-sensitivity field measurements. This study investigates the strain response of the highly magnetostrictive alloys Metglas® 2605SC and Vitrovac® 7600 T70 using Fiber Bragg Grating (FBG) acoustic sensors and an applied AC magnetic field. Sentek Instrument’s picoDAS interrogated the distributed FBG sensors set atop a ribbon of magnetostrictive material, and the corresponding strain response transferred to the fiber was analyzed. Using the Vitrovac® ribbon, a minimal detectable field amplitude of 60 nT was achieved. Using Metglas®, an even better sensitivity was demonstrated, where detected field amplitudes as low as 3 nT were measured via the strain response imparted to the FBG sensor. Distributed FBG sensors are readily available commercially, easily integrated into existing interrogation systems, and require no bonding to the magnetostrictive material for field detection. The simple sensor configuration with nanotesla-level sensitivity lends itself as a promising means of magnetic characterization and demonstrates the potential of fiber-optic acoustic sensors for distributed measurements. Full article
(This article belongs to the Section Chemical Sensors)
Show Figures

Figure 1

Figure 1
<p>Magnetic domains in a magnetostrictive material lattice at equilibrium (top) and under an external magnetic field at saturation strength (bottom) with a maximum magnetostrictive strain λ.</p>
Full article ">Figure 2
<p>Experimental test setup with air-core solenoid connected to function generator and amplifier. The magnetostrictive ribbon is positioned in the middle of the solenoid with the sensing fiber on top.</p>
Full article ">Figure 3
<p>Single-sided amplitude spectrums with an applied AC magnetic field at 350 Hz with amplitudes of 274 nT (<b>a</b>) and 2740 nT (<b>b</b>) from a Metglas<sup>®</sup> ribbon sample.</p>
Full article ">Figure 4
<p>Vitrovac<sup>®</sup> ribbon fiber sensor: 100 Hz amplitude spectrum intensity vs. 100 Hz AC magnetic-field amplitude.</p>
Full article ">Figure 5
<p>Three air-core solenoids containing Vitrovac<sup>®</sup> ribbon.</p>
Full article ">Figure 6
<p>Software capture of 3-dimensional matrix showing position, strain, and time for a half-second interval along 16 m of the fiber and 8 broadband FBG pairs.</p>
Full article ">Figure 7
<p>Metglas<sup>®</sup> ribbon fiber sensor: 200 Hz amplitude spectrum intensity vs. 100 Hz AC magnetic field amplitude.</p>
Full article ">Figure 8
<p>Metglas<sup>®</sup> ribbon fiber sensor: 350 Hz amplitude spectrum intensity vs. 350 Hz AC magnetic-field amplitude. (<b>a</b>) up to 60 nT and; (<b>b</b>) up to 6000 nT.</p>
Full article ">
37 pages, 5182 KiB  
Article
Hybrid Lanthanide Metal–Organic Compounds with Flavonoids: Magneto-Optical Properties and Biological Activity Profiles
by Sevasti Matsia, Anastasios Papadopoulos, Antonios Hatzidimitriou, Lars Schumacher, Aylin Koldemir, Rainer Pöttgen, Angeliki Panagiotopoulou, Christos T. Chasapis and Athanasios Salifoglou
Int. J. Mol. Sci. 2025, 26(3), 1198; https://doi.org/10.3390/ijms26031198 - 30 Jan 2025
Abstract
Lanthanides have seen rapid growth in the pharmaceutical and biomedical field, thus necessitating the development of hybrid metal–organic materials capable of exerting defined biological activities. Ternary hybrid lanthanide compounds were synthesized through reaction systems of Ln(III) (Ln = La, Nd, Eu) involving the [...] Read more.
Lanthanides have seen rapid growth in the pharmaceutical and biomedical field, thus necessitating the development of hybrid metal–organic materials capable of exerting defined biological activities. Ternary hybrid lanthanide compounds were synthesized through reaction systems of Ln(III) (Ln = La, Nd, Eu) involving the antioxidant flavonoid chrysin (Chr) and 1,10-phenanhtroline (phen) under solvothermal conditions, thus leading to pure crystalline materials. The so-derived compounds were characterized physicochemically in the solid state through analytical (elemental analysis), spectroscopic (FT-IR, UV-visible, luminescence, ESI-MS, circular dichroism, 151Eu Mössbauer), magnetic susceptibility, and X-ray crystallographic techniques. The analytical and spectroscopic data corroborate the 3D structure of the mononuclear complex assemblies and are in line with theoretical calculations (Bond Valence Sum and Hirshfeld analysis), with their luminescence suggesting quenching on the flavonoid-phen electronic signature. Magnetic susceptibility data suggest potential correlations, which could be envisioned, supporting future functional sensors. At the biological level, the title compounds were investigated for their (a) ability to interact with bovine serum albumin and (b) antibacterial efficacy against Gram(−) (E. coli) and Gram(+) (S. aureus) bacteria, collectively revealing distinctly configured biological profiles and suggesting analogous applications in cellular (patho)physiologies. Full article
(This article belongs to the Section Physical Chemistry and Chemical Physics)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Molecular structure of <b>1</b>; aromatic hydrogen atoms as well as solvate methanol molecules are omitted for clarity; atom colors: lanthanum, yellow; nitrogen, blue; oxygen, red; hydrogen, green. (<b>B</b>) Coordination polyhedron (real positions of the coordinated atoms together with the normal polyhedron) of <b>1</b>; atom colors: lanthanum, yellow; nitrogen, blue; oxygen, red.</p>
Full article ">Figure 2
<p>(<b>A</b>) Molecular structure of <b>2</b>; Aromatic hydrogen atoms as well as solvate methanol molecules are omitted for clarity; atom colors: neodymium, yellow; nitrogen, blue; oxygen, red; hydrogen, green. (<b>B</b>) Coordination polyhedron (real positions of the coordinated atoms together with the normal polyhedron) of <b>2</b>; atom colors: neodymium, yellow; nitrogen, blue; oxygen, red. (<b>C</b>) Hydrogen bonding interactions (blue dotted lines) in <b>2</b>.</p>
Full article ">Figure 3
<p>(<b>A</b>) Molecular structure of <b>3</b>; aromatic hydrogen atoms are omitted for clarity; atom colors: europium, yellow; nitrogen, blue; oxygen, red; hydrogen, green. (<b>B</b>) Coordination polyhedron (real positions of the coordinated atoms together with the normal polyhedron) of <b>3</b>; atom colors: europium, yellow; nitrogen, blue; oxygen, red.</p>
Full article ">Figure 4
<p>(<b>A</b>) Crystal Explorer plot of <b>1</b>. (<b>B</b>) d<sub>norm</sub> mapping of <b>1</b> through Hirshfeld surface analysis. (<b>C</b>) Shape index mapping of <b>1</b> through Hirshfeld surface analysis. (<b>D</b>) Curvedness mapping of <b>1</b> through Hirshfeld surface analysis. The different colors shown in the figure are identified and explained in detail in the text.</p>
Full article ">Figure 5
<p>(<b>A</b>) Full fingerprint plot of <b>1</b> and d<sub>norm</sub> mapping. (<b>B</b>) Internal vs. external 2D fingerprint plot distances of H···H contacts of <b>1</b> with the relevant percentage contribution mapped over d<sub>norm</sub>. (<b>C</b>) 2D fingerprint plot of H···O/O···H contacts and their appropriate percentage contribution reflected onto the Hirshfeld surface area mapper over d<sub>norm</sub> of <b>1</b>. (<b>D</b>) 2D Fingerprint plot of H···C/C···H contacts, with the relevant percentage contribution reflected onto the Hirshfeld surface area mapper over d<sub>norm</sub> of <b>1</b>.</p>
Full article ">Figure 6
<p>(<b>A</b>) Comparative UV-visible spectra of <b>1</b> with phen and Chr in methanol at 10<sup>−5</sup> M. (<b>B</b>) Electronic spectrum (red line) and spectral fitting (scatter) of compound <b>1</b> in methanol (10<sup>−5</sup> M).</p>
Full article ">Figure 7
<p>ESI-MS spectra of <b>3</b> and the appropriate species in methanol solution through the positive mode of ionization.</p>
Full article ">Figure 8
<p>Comparative and normalized solid-state luminescence spectra between <b>2</b> and (<b>A</b>) Chr at λ<sub>ex</sub> 445 nm. (<b>B</b>) Phen at λ<sub>ex</sub> 373 nm.</p>
Full article ">Figure 9
<p>(<b>A</b>) Temperature dependence of the magnetic susceptibility of <b>1</b> measured at 10 kOe. (<b>B</b>) Magnetic properties of <b>2</b>: (<b>top</b>) temperature dependence of the magnetic susceptibility (<span class="html-italic">χ</span> and <span class="html-italic">χ</span><sup>−1</sup> data) measured at 10 kOe; (<b>bottom</b>) magnetization isotherms at 3, 10, and 50 K. (<b>C</b>) Magnetic properties of <b>3</b>: (<b>top</b>) temperature dependence of the magnetic susceptibility measured at 10 kOe. The calculated susceptibilities (red line) were obtained using the Van Vleck expression for the paramagnetic susceptibilities of free Eu(III) ions with λ = 734(1) K; (<b>bottom</b>) magnetization isotherms at 3, 10 and 50 K.</p>
Full article ">Figure 10
<p>Experimental (data points) and simulated (red line) <sup>151</sup>Eu Mössbauer spectrum of <b>3</b> measured at 78 K.</p>
Full article ">Figure 11
<p>UV-visible absorption spectra of solutions containing BSA (3 μΜ, PBS) and increasing molar ratios of <b>2</b> (DMSO).</p>
Full article ">Figure 12
<p>Fluorescence spectra of solutions containing BSA (1.5 μM, PBS) and molar ratios of <b>2</b> (DMSO). <b>Inset</b>: Stern–Volmer plot acquired from steady-state fluorescence at (<b>A</b>) 20 °C, (<b>B</b>) 30 °C, and (<b>C</b>) 37 °C.</p>
Full article ">Figure 13
<p>Van’t Hoff plot of <b>2</b> from measurements at 20 °C, 30 °C, and 37 °C.</p>
Full article ">Figure 14
<p>Circular dichroism spectra of solutions containing BSA (1 μM, PBS) and increasing molar ratios of (<b>A</b>) <b>1</b> (MeOH) and (<b>B</b>) <b>3</b> (MeOH).</p>
Full article ">Figure 15
<p>(<b>A</b>) Compound <b>1</b> was docked against the 3D structure of BSA. (<b>B</b>) 2D interaction diagrams illustrate the interactions between compound <b>1</b> and the BSA binding motif (hydrophobic contacts are indicated in red).</p>
Full article ">
17 pages, 2854 KiB  
Article
High-Accuracy Clock Synchronization in Low-Power Wireless sEMG Sensors
by Giorgio Biagetti, Michele Sulis, Laura Falaschetti and Paolo Crippa
Sensors 2025, 25(3), 756; https://doi.org/10.3390/s25030756 - 26 Jan 2025
Abstract
Wireless surface electromyography (sEMG) sensors are very practical in that they can be worn freely, but the radio link between them and the receiver might cause unpredictable latencies that hinder the accurate synchronization of time between multiple sensors, which is an important aspect [...] Read more.
Wireless surface electromyography (sEMG) sensors are very practical in that they can be worn freely, but the radio link between them and the receiver might cause unpredictable latencies that hinder the accurate synchronization of time between multiple sensors, which is an important aspect to study, e.g., the correlation between signals sampled at different sites. Moreover, to minimize power consumption, it can be useful to design a sensor with multiple clock domains so that each subsystem only runs at the minimum frequency for correct operation, thus saving energy. This paper presents the design, implementation, and test results of an sEMG sensor that uses Bluetooth Low Energy (BLE) communication and operates in three different clock domains to save power. In particular, this work focuses on the synchronization problem that arises from these design choices. It was solved through a detailed study of the timings experimentally observed over the BLE connection, and through the use of a dual-stage filtering mechanism to remove timestamp measurement noise. Time synchronization through three different clock domains (receiver, microcontroller, and ADC) was thus achieved, with a resulting total jitter of just 47 µs RMS for a 1.25 ms sampling period, while the dedicated ADC clock domain saved between 10% to 50% of power, depending on the selected data rate. Full article
Show Figures

Figure 1

Figure 1
<p>Block diagram of the sEMG sensor. It is based on a Nordic Semiconductor nRF52840 SoC that integrates an ARM Cortex-M4 CPU running at 64 MHz and a multi-protocol radio compatible with Bluetooth 5 Low-Energy mode. A Texas Instruments ADS1293 integrated analog front-end (AFE) and ADC are at the core of the sEMG signal acquisition chain, while an STMicroelectronics LSM6DSO inertial measurement unit complements the system.</p>
Full article ">Figure 2
<p>Pictures of the assembled prototype (scale 2:1). (<b>a</b>) Front side: The ADS1293 AFE (U2) is clearly visible in the center, the BLE radio and microcontroller at the top (beneath the RF shield), and pads for the electrodes at the bottom. (<b>b</b>) Back side: The 4.096 MHz crystal was added here as there was no room on the front side. Being a THT component, it was bound to be soldered manually anyway, so this type of double-sided assembly did not significantly increase production costs.</p>
Full article ">Figure 3
<p>Examples of measured timestamps. The microcontroller time scale is corrected (to take into account oscillator frequency error) by multiplying it by a constant <span class="html-italic">k</span> obtained by linear regression with the RX time scale, so that the curves fit within the graph even for large deviations of <span class="html-italic">k</span> from 1. As a consequence, the blue dot “lines” appear almost horizontal, even though their actual slope (frequency error) should be <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mi>k</mi> </mrow> </semantics></math>. (<b>a</b>) Ideal results as obtained from high-level VHDL simulation of the system, (<b>b</b>) measured differences using the 64 MHz HFXO to clock the microcontroller and the PWM fractional divider to clock the ADC, (<b>c</b>) measured differences using the 64 MHz HFXO to clock the microcontroller and the independent 4.096 MHz crystal to clock the ADC, (<b>d</b>) measured differences using the internal RC 64 MHz low-power oscillator to clock the microcontroller and the independent 4.096 MHz crystal to clock the ADC.</p>
Full article ">Figure 4
<p>Jitter measurement. The blue dots represent clock periods measured with the PWM output (fractional divider); 1/16th of them are around 2500 ns (<math display="inline"><semantics> <mrow> <mn>40</mn> <mspace width="0.166667em"/> <msub> <mi>T</mi> <mi>PWM</mi> </msub> </mrow> </semantics></math>), and 15/16th of them around 2437.5 ns (<math display="inline"><semantics> <mrow> <mn>39</mn> <mspace width="0.166667em"/> <msub> <mi>T</mi> <mi>PWM</mi> </msub> </mrow> </semantics></math>). The red dots are measured with the 4.096 MHz crystal oscillator output internally divided by 10 to produce the required 409.6 kHz (2441.4 ns) modulator clock.</p>
Full article ">Figure 5
<p>Synchronization of the received data to an absolute time scale. The time axis represents the absolute time (obtained through NTP) from the rising edge of the applied pulses (dashed blue line). The black lines are 10 randomly selected measures over a span of 2 min.</p>
Full article ">Figure 6
<p>Measurement of the current drawn by the device while streaming data at 800 Hz, 3 channels, 24 bit per channel. Top: using the PWM fractional divider as the clock source for the ADC. Bottom: using the dedicated crystal for the ADC clock. Peaks correspond to packet transmission, while the clock source mainly affect the baseline draw and will thus be much more evident with lower rate configurations.</p>
Full article ">Figure 7
<p>Measurement of the current drawn by the device while streaming data at 200 Hz, 1 channel, 20 bit per channel, VLDE encoding. Top: using the PWM fractional divider as the clock source for the ADC. Bottom: using the dedicated crystal for the ADC clock.</p>
Full article ">
4 pages, 137 KiB  
Editorial
Biomedical Signal Processing and Health Monitoring Based on Sensors
by Sang Ho Choi, Heenam Yoon, Hyun Jae Baek and Xi Long
Sensors 2025, 25(3), 641; https://doi.org/10.3390/s25030641 - 22 Jan 2025
Viewed by 402
Abstract
The healthcare industry is undergoing rapid transformation driven by advancements in Internet of Things (IoT) technologies, particularly in biomedical signal processing and health monitoring [...] Full article
(This article belongs to the Special Issue Biomedical Signal Processing and Health Monitoring Based on Sensors)
Back to TopTop