[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,739)

Search Parameters:
Keywords = cyclic voltammetry

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 7613 KiB  
Article
Electrochemical Analysis of Amyloid Plaques and ApoE4 with Chitosan-Coated Gold Nanostars for Alzheimer’s Detection
by Min-Kyung Shin, Ariadna Schuck, Minhee Kang and Yong-Sang Kim
Biosensors 2024, 14(10), 510; https://doi.org/10.3390/bios14100510 (registering DOI) - 17 Oct 2024
Abstract
Monitoring the progression of Alzheimer’s disease (AD) is crucial for mitigating dementia symptoms, alleviating pain, and improving mobility. Traditionally, AD biomarkers like amyloid plaques are predominantly identified in cerebrospinal fluid (CSF) due to their concentrated presence. However, detecting these markers in blood is [...] Read more.
Monitoring the progression of Alzheimer’s disease (AD) is crucial for mitigating dementia symptoms, alleviating pain, and improving mobility. Traditionally, AD biomarkers like amyloid plaques are predominantly identified in cerebrospinal fluid (CSF) due to their concentrated presence. However, detecting these markers in blood is hindered by the blood–brain barrier (BBB), resulting in lower concentrations. To address this challenge and identify pertinent AD biomarkers—specifically amyloid plaques and apolipoprotein E4 (ApoE4)—in blood plasma, we propose an innovative approach. This involves enhancing a screen-printed carbon electrode (SPCE) with an immobilization matrix comprising gold nanostars (AuNSs) coated with chitosan. Morphological and electrical analyses confirmed superior dispersion and conductivity with 0.5% chitosan, supported by UV–Vis spectroscopy, cyclic voltammetry, and Nyquist plots. Subsequent clinical assays measured electrical responses to quantify amyloid-β 42 (Aβ42) (15.63–1000 pg/mL) and APoE4 levels (0.41 to 40 ng/mL) in human blood plasma samples. Differential pulse voltammetry (DPV) responses exhibited peak currents proportional to biomarker concentrations, demonstrating high linear correlations (0.985 for Aβ42 and 0.919 for APoE4) with minimal error bars. Cross-reactivity tests with mixed solutions of amyloid-β 40 (Aβ40), Aβ42, and ApoE4 indicated minimal interference between biomarkers (<3% variation), further confirming the high specificity of the developed sensor. Validation studies demonstrated a strong concurrence with the gold-standard enzyme-linked immunosorbent assay (ELISA), while interference tests indicated a minimal variation in peak currents. This improved device presents promising potential as a point-of-care system, offering a less invasive, cost-effective, and simplified approach to detecting and tracking the progression of AD. The substantial surface binding area further supports the efficacy of our method, offering a promising avenue for advancing AD diagnostics. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the proposed methodology for detecting Alzheimer’s disease biomarkers. This study’s overarching concept involves quantifying amyloid plaques and ApoE4 in human blood samples, considering the disruption of the blood–brain barrier (BBB). Plasma is separated from whole blood and analyzed using SPCE devices to quantify low concentrations of Aβ42 and ApoE4. Specific antibodies are conjugated to the CNT-PEG-AuNS-modified working electrodes prior to the assays. A small volume of plasma is applied to the active area of the SPCE, and electrochemical analysis is conducted using the DPV technique, where peak currents correlate with target concentrations.</p>
Full article ">Figure 2
<p>Morphological characterization: Analysis of enhanced carbon nanotubes decorated with PEG-coated gold nanostars. SEM images of CNT-AuNS-PEG at various magnifications: (<b>A</b>) ×100, (<b>B</b>) ×80 k, and (<b>C</b>) ×330 k. (<b>D</b>) FTIR spectra and (<b>E</b>) Raman spectra comparing bare CNTs with the enhanced structures (CNT-AuNS-PEG). (<b>F</b>) UV–Vis absorption spectra of CNT-AuNS-PEG nanocomposites before and after the immobilization of Alzheimer’s disease-specific antibodies.</p>
Full article ">Figure 3
<p>Electrochemical characterization of the SPCE device: Analysis conducted using a buffer solution. (<b>A</b>) Cyclic voltammograms (CVs) of bare CNT, CNT-COOH, CNT-Cl, and CNT-AuNS-PEG at a scan rate of 100 mV s<sup>–1</sup>. (<b>B</b>) Relationship between cathodic and anodic peak currents and the square root of scan rates (inset: CVs of CNT-AuNS-PEG at various scan rates ranging from 10 to 200 mV s<sup>−1</sup>). (<b>C</b>) Nyquist plots for electrodes: bare CNT, CNT-COOH, CNT-Cl, and CNT-AuNS-PEG in 0.1 mol L<sup>−1</sup> KCl solution containing 5.0 mmol L<sup>−1</sup> Fe(CN)<sub>6</sub><sup>3−/4−</sup> (inset: equivalent circuits with corresponding elements).</p>
Full article ">Figure 4
<p>Evaluation of the CNT-AuNS-PEG active layer: Investigation of its effectiveness in detecting Alzheimer’s disease-related biomarkers in a 1× PBS buffer solution by analyzing the variation in peak currents from differential pulse voltammetry (DPV) measurements for (<b>A</b>) amyloid-β and (<b>B</b>) ApoE4. The inset graphs display results at different concentrations on a logarithmic scale. Error bars indicate the standard deviation from measurements obtained across three devices.</p>
Full article ">Figure 5
<p>Electrochemical evaluation: Analysis of Alzheimer’s disease biomarkers in clinical samples, illustrating the variation in peak currents with DPV responses. The inset graphs display the responses at different concentrations of (<b>A</b>) Aβ42 (15.63–1000 pg/mL) and (<b>B</b>) ApoE4 (0.41 to 40 ng/mL) in human plasma, shown on a logarithmic scale. Error bars indicate the standard deviation from measurements taken across three devices.</p>
Full article ">Figure 6
<p>Interference assessment: Evaluation of the impact of various interfering compounds on the CNT-AuNS-PEG device. Differential pulse voltammetry (DPV) peak currents were measured for (<b>A</b>) amyloid-β 42 (circle) and (<b>B</b>) ApoE4 (rhombus), following the addition of interferents (triangle): 60 µM uric acid (UA), 10 µM ascorbic acid (AA), 50 mM lactic acid (LA), and 0.2 mM glucose (GLU) in plasma samples.</p>
Full article ">Figure 7
<p>Cross-reactivity assessment: Analysis of the interaction among Alzheimer’s disease biomarkers. (<b>A</b>) Differential pulse voltammetry (DPV) peak currents of the CNT-AuNS-PEG when measuring amyloid-β 40 as a potential Alzheimer’s biomarker. The inset graph displays the DPV curves of amyloid-β 40 across different concentrations. (<b>B</b>) Evaluation of the enhanced screen-printed carbon electrode (SPCE) for cross-reactivity with samples containing Aβ40, Aβ42, and ApoE4.</p>
Full article ">Figure 8
<p>Validation analysis: Comparison between gold-standard ELISA laboratory tests and CNT-AuNS-PEG-based devices. Bland–Altman plots depict the level of agreement for measurements using diluted human plasma samples containing (<b>A</b>) Aβ42 and (<b>B</b>) ApoE4.</p>
Full article ">
19 pages, 3472 KiB  
Article
Electrochemical DNA Sensor Based on Poly(proflavine) Deposited from Natural Deep Eutectic Solvents for DNA Damage Detection and Antioxidant Influence Assessment
by Anna Porfireva, Anastasia Goida, Vladimir Evtugyn, Milena Mozgovaya, Tatiana Krasnova and Gennady Evtugyn
Chemosensors 2024, 12(10), 215; https://doi.org/10.3390/chemosensors12100215 - 16 Oct 2024
Viewed by 370
Abstract
Electrochemical DNA sensors for DNA damage detection based on electroactive polymer poly(proflavine) (PPFL) that was synthesized at screen-printed carbon electrodes (SPCEs) from phosphate buffer (PB) and two natural deep eutectic solvents (NADESs) consisting of citric or malonic acids, D-glucose, and a certain amount [...] Read more.
Electrochemical DNA sensors for DNA damage detection based on electroactive polymer poly(proflavine) (PPFL) that was synthesized at screen-printed carbon electrodes (SPCEs) from phosphate buffer (PB) and two natural deep eutectic solvents (NADESs) consisting of citric or malonic acids, D-glucose, and a certain amount of water (NADES1 and NADES2) were developed. Poly(proflavine) coatings obtained from the presented media (PPFLPB, PPFLNADES1, and PPFLNADES2) were electrochemically polymerized via the multiple cycling of the potential or potentiostatic accumulation and used for the discrimination of thermal and oxidative DNA damage. The electrochemical characteristics of the poly(proflavine) coatings and their morphology were assessed using cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), and scanning electron microscopy (SEM). The working conditions for calf thymus DNA implementation and DNA damage detection were estimated for all types of poly(proflavine) coatings. The voltammetric approach made it possible to distinguish native and chemically oxidized DNA while the impedimetric approach allowed for the successful recognition of native, thermally denatured, and chemically oxidized DNA through changes in the charge transfer resistance. The influence of different concentrations of conventional antioxidants and pharmaceutical preparations on oxidative DNA damage was characterized. Full article
(This article belongs to the Special Issue Electrochemical Biosensors: Advances and Prospects)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of proflavine hydrochloride.</p>
Full article ">Figure 2
<p>Multiple cyclic voltammograms recorded on the SPCE in (<b>a</b>) 0.025 M PB, pH 7.0, containing 0.5 mM proflavine or 0.085 M proflavine in (<b>b</b>) NADES1 and (<b>c</b>) NADES2; 0.1 V/s. Arrows indicate changes with increased number of cycles.</p>
Full article ">Figure 3
<p>Oxidation (black) and reduction (red) peak currents recorded in 0.025 M PB on the SPCE covered with (<b>a</b>) PPFL<sub>PB</sub>, (<b>b</b>) PPFL<sub>NADES1</sub>, and (<b>c</b>) PPFL<sub>NADES2</sub> at pH values of 2.0–9.0 (average ± S.D. for eight individual sensors).</p>
Full article ">Figure 4
<p>SEM images of (<b>a</b>) bare SPCE and SPCE covered with (<b>b</b>) PPFL<sub>PB</sub>, (<b>c</b>) PPFL<sub>NADES1</sub>, and (<b>d</b>) PPFL<sub>NADES2</sub>.</p>
Full article ">Figure 5
<p>Particle size distributions (Gaussian fitting) for (<b>a</b>) carbon ink nanoparticles of bare SPCE and microspheres of (<b>b</b>) PPFL<sub>PB</sub>, (<b>c</b>) PPFL<sub>NADES1</sub>, and (<b>d</b>) PPFL<sub>NADES2</sub> coatings.</p>
Full article ">Figure 6
<p>Relative changes in peak oxidation (black) and reduction (gray) currents after DNA immobilization on (<b>a</b>) PPFL<sub>PB</sub>, (<b>b</b>) PPFL<sub>NADES1</sub>, and (<b>c</b>) PPFL<sub>NADES2</sub>. DNA immobilization protocol: 1—drying; 2–5—incubation in DNA solution for 10, 20, 30, and 40 min, respectively. Specifications: cyclic voltammetry, 0.025 M PB; pH 7.0; from −0.6 to 0.6 V, 0.1 V/s.</p>
Full article ">Figure 7
<p>Relative changes in peak oxidation (black) and reduction (gray) currents for (<b>a</b>) PPFL<sub>PB</sub>, (<b>b</b>) PPFL<sub>NADES1</sub>, and (<b>c</b>) PPFL<sub>NADES2</sub>. Layer contents: 1—with no DNA, 2—native DNA, 3—thermally denatured DNA, and 4—chemically oxidized (Cu<sup>2+</sup>/H<sub>2</sub>O<sub>2</sub>) DNA.</p>
Full article ">Figure 8
<p>Equivalent circuit <span class="html-italic">Rs</span>(<span class="html-italic">R</span><sub>1</sub><span class="html-italic">C</span><sub>1</sub>)(<span class="html-italic">R</span><sub>2</sub><span class="html-italic">C</span><sub>2</sub>) for the evaluation of EIS parameters.</p>
Full article ">Figure 9
<p>Dependence of charge transfer resistance <span class="html-italic">R</span><sub>1</sub> on modified layer content: 1—bare SPCE, 2—SPCE/PPFL, 3—SPCE/PPFL/native DNA, 4—SPCE/PPFL/denatured DNA, 5—SPCE/PPFL/oxidized DNA, 6—SPCE/PPFL/PSS, and 7—SPCE/PPFL/H<sub>2</sub>O; (<b>a</b>) PPFL<sub>PB</sub>, (<b>b</b>) PPFL<sub>NADES1</sub>, (<b>c</b>) PPFL<sub>NADES2</sub>; 0.025 M PB, pH 7.0, in the presence of 0.01 M [Fe(CN)<sub>6</sub>]<sup>3−/4−</sup>. Average ± S.D. for ten individual sensors.</p>
Full article ">Figure 10
<p>Dependence of EIS parameters on antioxidant concentration. (<b>a</b>) Charge transfer resistance <span class="html-italic">R</span><sub>1</sub>, (<b>b</b>) constant phase element <span class="html-italic">C</span><sub>1</sub>; 0.025 M PB, pH 7.0, in the presence of 0.01 M [Fe(CN)<sub>6</sub>]<sup>3−/4−</sup>. Average ± S.D. values are shown for ten individual sensors.</p>
Full article ">
18 pages, 9604 KiB  
Article
Green Silver Nanoparticles: Plant-Extract-Mediated Synthesis, Optical and Electrochemical Properties
by Natalia Stozhko, Aleksey Tarasov, Viktoria Tamoshenko, Maria Bukharinova, Ekaterina Khamzina and Veronika Kolotygina
Physchem 2024, 4(4), 402-419; https://doi.org/10.3390/physchem4040028 - 16 Oct 2024
Viewed by 195
Abstract
Antioxidants of plant extract play an important role in the phytosynthesis of silver nanoparticles (phyto-AgNPs), providing the reduction of silver ions and capping and stabilization of nanoparticles. Despite the current progress in the studies of phytosynthesis, there is no approach to the selection [...] Read more.
Antioxidants of plant extract play an important role in the phytosynthesis of silver nanoparticles (phyto-AgNPs), providing the reduction of silver ions and capping and stabilization of nanoparticles. Despite the current progress in the studies of phytosynthesis, there is no approach to the selection of plant extract for obtaining phyto-AgNPs with desired properties. This work shows that antioxidant activity (AOA) of plant extracts is a key parameter for targeted phytosynthesis. In support of this fact, the synthesis of phyto-AgNPs was carried out using extracts of four plants with different AOA, increasing in the order Ribes uva-crispa < Lonicera caerulea < Fragaria vesca < Hippophae rhamnoides. Phyto-AgNPs have been characterized using Fourier-transform infrared spectroscopy, transmission electron microscopy, energy-dispersive X-ray spectroscopy, selected area electron diffraction technique, ultraviolet–visible spectroscopy, electrochemical impedance spectroscopy and cyclic voltammetry. It was established that the change in the AOA of the plant extract is accompanied by a size-dependent change in the optical and electrochemical properties of phyto-AgNPs. In particular, an increase in the extract AOA leads to the formation of smaller phyto-AgNPs with higher electrochemical activity and low charge transfer resistance. A “blue shift” and an increase in the plasmon resonance band of silver sols are observed with an increase in the extract AOA. The obtained regularities prove the existence of the “AOA–size–properties” triad, which can be used for controlled phytosynthesis and prediction of phyto-AgNPs’ properties. Full article
(This article belongs to the Section Electrochemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR spectra for gooseberry (<b>a</b>), blue honeysuckle (<b>b</b>), strawberry (<b>c</b>) and sea buckthorn (<b>d</b>). Spectra 1 and 2 correspond to the plant extract and phyto-AgNP sol, respectively.</p>
Full article ">Figure 2
<p>TEM images, histograms of particle distribution by size, EDS spectra and SAED patterns of phyto-AgNPs synthesized using extracts of gooseberry (<b>a</b>), blue honeysuckle (<b>b</b>), strawberry (<b>c</b>) and sea buckthorn (<b>d</b>).</p>
Full article ">Figure 3
<p>Photographs (<b>a</b>) and absorption spectra (<b>b</b>) of phyto-AgNP sols synthesized using extracts of various plants. A<sub>max</sub> (<b>c</b>) and λ<sub>SPR</sub> (<b>d</b>) of phyto-AgNP sols depending on the plant extract AOA.</p>
Full article ">Figure 3 Cont.
<p>Photographs (<b>a</b>) and absorption spectra (<b>b</b>) of phyto-AgNP sols synthesized using extracts of various plants. A<sub>max</sub> (<b>c</b>) and λ<sub>SPR</sub> (<b>d</b>) of phyto-AgNP sols depending on the plant extract AOA.</p>
Full article ">Figure 4
<p>Absorption spectra of phyto-AgNPs(b) sols obtained as a result of synthesis using sea buckthorn extracts with different AOA (<b>a</b>). Dependence of A<sub>max</sub> (<b>b</b>) and λ<sub>SPR</sub> (<b>c</b>) of phyto-AgNPs(b) sols on AOA of sea buckthorn extracts.</p>
Full article ">Figure 5
<p>Dependence of A<sub>max</sub> (<b>a</b>) and λ<sub>SPR</sub> (<b>b</b>) of phyto-AgNP sols on the plant extract AOA.</p>
Full article ">Figure 6
<p>Comparison of A<sub>max</sub> (<b>a</b>) and λ<sub>SPR</sub> (<b>b</b>) of phyto-AgNP sols obtained immediately after synthesis at pH 5 and 16 h after it.</p>
Full article ">Figure 7
<p>Nyquist plot (<b>a</b>) and Bode plot (<b>b</b>) for bare and phyto-AgNP-modified SPCE in 0.1 mol L<sup>−1</sup> KCl solution containing 5 mmol L<sup>−1</sup> K<sub>3</sub>[Fe(CN)<sub>6</sub>].</p>
Full article ">Figure 8
<p>Cyclic voltammograms of plant extracts (1), native (2) and “washed” (3) phytoAgNPs synthesized using gooseberry (<b>a</b>), blue honeysuckle (<b>b</b>), strawberry (<b>c</b>) and sea buckthorn (<b>d</b>). Background 0.1 mol L<sup>–1</sup> H<sub>2</sub>SO<sub>4</sub>. Potential scan rate 50 mV s<sup>−1</sup>.</p>
Full article ">Figure 9
<p>Cyclic voltammograms of phyto-AgNPs synthesized using extracts of different plants at pH 5 (<b>a</b>). Dependences of I<sub>p</sub> (<b>b</b>) and E<sub>p</sub> (<b>c</b>) of phyto-AgNP oxidation on plant extract AOA.</p>
Full article ">
18 pages, 4283 KiB  
Article
A Machine Learning Assisted Non-Enzymatic Electrochemical Biosensor to Detect Urea Based on Multi-Walled Carbon Nanotube Functionalized with Copper Oxide Micro-Flowers
by Jitendra B. Zalke, Manish L. Bhaiyya, Pooja A. Jain, Devashree N. Sakharkar, Jayu Kalambe, Nitin P. Narkhede, Mangesh B. Thakre, Dinesh R. Rotake, Madhusudan B. Kulkarni and Shiv Govind Singh
Biosensors 2024, 14(10), 504; https://doi.org/10.3390/bios14100504 (registering DOI) - 15 Oct 2024
Viewed by 326
Abstract
Detecting urea is crucial for diagnosing related health conditions and ensuring timely medical intervention. The addition of machine learning (ML) technologies has completely changed the field of biochemical sensing, providing enhanced accuracy and reliability. In the present work, an ML-assisted screen-printed, flexible, electrochemical, [...] Read more.
Detecting urea is crucial for diagnosing related health conditions and ensuring timely medical intervention. The addition of machine learning (ML) technologies has completely changed the field of biochemical sensing, providing enhanced accuracy and reliability. In the present work, an ML-assisted screen-printed, flexible, electrochemical, non-enzymatic biosensor was proposed to quantify urea concentrations. For the detection of urea, the biosensor was modified with a multi-walled carbon nanotube-zinc oxide (MWCNT-ZnO) nanocomposite functionalized with copper oxide (CuO) micro-flowers (MFs). Further, the CuO-MFs were synthesized using a standard sol-gel approach, and the obtained particles were subjected to various characterization techniques, including X-ray diffraction (XRD), field emission scanning electron microscopy (FESEM), and Fourier transform infrared (FTIR) spectroscopy. The sensor’s performance for urea detection was evaluated by assessing the dependence of peak currents on analyte concentration using cyclic voltammetry (CV) at different scan rates of 50, 75, and 100 mV/s. The designed non-enzymatic biosensor showed an acceptable linear range of operation of 0.5–8 mM, and the limit of detection (LoD) observed was 78.479 nM, which is well aligned with the urea concentration found in human blood and exhibits a good sensitivity of 117.98 mA mM−1 cm−2. Additionally, different regression-based ML models were applied to determine CV parameters to predict urea concentrations experimentally. ML significantly improves the accuracy and reliability of screen-printed biosensors, enabling accurate predictions of urea levels. Finally, the combination of ML and biosensor design emphasizes not only the high sensitivity and accuracy of the sensor but also its potential for complex non-enzymatic urea detection applications. Future advancements in accurate biochemical sensing technologies are made possible by this strong and dependable methodology. Full article
(This article belongs to the Special Issue Advances in Biosensing and Bioanalysis Based on Nanozymes)
Show Figures

Figure 1

Figure 1
<p>Process steps to detect urea, based on MWCNT-ZnO functionalized with novel CuO-MF and ML-approach.</p>
Full article ">Figure 2
<p>(<b>A</b>) Process steps for synthesis and preparation of CuO-MFs. (<b>B</b>) Functionalization of Gii-Sens Integrated Graphene SPE.</p>
Full article ">Figure 3
<p>(<b>A</b>) SEM image of MWCNT−ZnO showing the exact morphology of nanofibers, (<b>B</b>) EDX spectrum of MWCNT−ZnO nanofibers showing the elemental content material composition, (<b>C</b>) XRD analysis provides the crystalline structure of the synthesized MWCNT−ZnO nanofibers, (<b>D</b>) FTIR analysis showing the functional groups in the MWCNT−ZnO composite.</p>
Full article ">Figure 4
<p>(<b>A</b>) 3D pictures of the CuO-MFs, derived from a morphological study performed using scanning electron microscopy (SEM). (<b>B</b>) EDX spectra of CuO-MFs. (<b>C</b>) XRD spectra of CuO-MFs. (<b>D</b>) FTIR spectrum of CuO-MFs.</p>
Full article ">Figure 5
<p>SEM image of MWCNT-ZnO/CuO micro-flower deposition on working electrode of Integrated Graphene IG-GII-SENS-01 SPE at magnification levels of (<b>A</b>) 1 mm, (<b>B</b>) 10 µm, (<b>C</b>) 1 µm, and (<b>D</b>) 100 nm.</p>
Full article ">Figure 6
<p>Cyclic voltammetry (CV) responses of the MWCNT−ZnO/CuO−MFs modified SPEs were measured at various concentrations of urea ranging from 0.5 to 10 mM in the presence of a 5 mM solution of Ferroferricyanide [Fe(CN)<sub>6</sub>]<sup>3−4−</sup> as the standard redox probe at scan rates of (<b>A</b>) 50 mV/s, (<b>B</b>) 75 mV/s, and (<b>C</b>) 100 mV/s. Corresponding calibration plot of urea concentration (mM) versus current (A/cm<sup>2</sup>) for a scan rate of (<b>D</b>) 50 mV/s, (<b>E</b>) 75 mV/s, (<b>F</b>) 100 mV/s (n = 5).</p>
Full article ">Figure 7
<p>(<b>A</b>) Selectivity study with commonly identified interfering elements in human blood such as Galactose, Dextrose, Maltose, Lactose, Ascorbic Acid, and Uric Acid with a concentration of 0.1 mM at the scan rate of 100 mV/se, (<b>B</b>) Stability study of MWCNT−ZnO/CuO−MFs modified sensor for urea detection.</p>
Full article ">Figure 8
<p>Prediction of urea concentration using machine learning algorithms: The available machine learning models are (<b>A</b>) LR, (<b>B</b>) DT, (<b>C</b>) RF, (<b>D</b>) KNN, (<b>E</b>) AdaBoost, and (<b>F</b>) GB.</p>
Full article ">
10 pages, 1905 KiB  
Article
Unraveling Asymmetric Electrochemical Kinetics in Low-Mass-Loading LiNi1/3Mn1/3Co1/3O2 (NMC111) Li-Metal All-Solid-State Batteries
by Byoung-Nam Park
Materials 2024, 17(20), 5014; https://doi.org/10.3390/ma17205014 - 14 Oct 2024
Viewed by 366
Abstract
In this study, we fabricated a Li-metal all-solid-state battery (ASSB) with a low mass loading of NMC111 cathode electrode, enabling a sensitive evaluation of interfacial electrochemical reactions and their impact on battery performance, using Li1.3Al0.3Ti1.7(PO4) [...] Read more.
In this study, we fabricated a Li-metal all-solid-state battery (ASSB) with a low mass loading of NMC111 cathode electrode, enabling a sensitive evaluation of interfacial electrochemical reactions and their impact on battery performance, using Li1.3Al0.3Ti1.7(PO4)3 (LATP) as the solid electrolyte. The electrochemical behavior of the battery was analyzed to understand how the solid electrolyte influences charge storage mechanisms and Li-ion transport at the electrolyte/electrode interface. Cyclic voltammetry (CV) measurements revealed the b-values of 0.76 and 0.58, indicating asymmetry in the charge storage process. A diffusion coefficient of 1.5 × 10−9 cm2⋅s−1 (oxidation) was significantly lower compared to Li-NMC111 batteries with liquid electrolytes, 1.6 × 10−8cm2⋅s−1 (oxidation), suggesting that the asymmetric charge storage mechanisms are closely linked to reduced ionic transport and increased interfacial resistance in the solid electrolyte. This reduced Li-ion diffusivity, along with the formation of space charge layers at the electrode/electrolyte interface, contributes to the observed asymmetry in charge and discharge processes and limits the rate capability of the solid-state battery, particularly at high charging rates, compared to its liquid electrolyte counterpart. Full article
(This article belongs to the Special Issue Advanced Materials for Battery Applications and Photoelectric Devices)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic representation of a Li-metal NMC111 ASSB CR2032 cell. (<b>b</b>) The X-ray diffraction (XRD) pattern of the NMC111 electrode coated on Al foil. (<b>c</b>) The Nyquist plot of a Li/LATP/Li cell configuration. (<b>d</b>) The cyclic voltammetry (CV) curves recorded at a scan rate of 0.1 mV·s<sup>−1</sup>.</p>
Full article ">Figure 2
<p>(<b>a</b>) Rate capability of the Li-metal NMC111 ASSB at varying C-rates. (<b>b</b>) Charge/discharge potential profiles. (<b>c</b>) Differential capacity (dQ/dV) versus voltage plots over multiple cycles.</p>
Full article ">Figure 3
<p>(<b>a</b>) CV curves of the Li-metal NMC111 ASSB at varying scan rates. (<b>b</b>) The logarithmic plot of current versus scan rate used for <span class="html-italic">b</span>-value determination. (<b>c</b>) Peak current plotted as a function of the square root of the scan rate to calculate the Li-ion diffusion coefficient. The calculated <span class="html-italic">b</span>-values were 0.76 (anodic) and 0.58 (cathodic), and the Li-ion diffusion coefficient was determined to be 1.5 × 10<sup>−9</sup> cm<sup>2</sup>⋅s<sup>−1</sup> (anodic) and 7.6 × 10<sup>−10</sup> cm<sup>2</sup>⋅s<sup>−1</sup> (cathodic).</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) Comparative analysis of diffusion-controlled and surface capacitive contributions at varying scan rates. (<b>d</b>) Contribution of charge storage mechanisms at different scan rates, with surface capacitive effects increasing at higher scan rates.</p>
Full article ">
13 pages, 1654 KiB  
Article
When a Side Reaction Is a Benefit: A Catalyst-Free Route to Obtain High-Molecular Cobaltocenium-Functionalized Polysiloxanes by Hydroamination
by Anastasia N. Kocheva, Konstantin V. Deriabin, Igor Perevyazko, Nadezhda A. Bokach, Vadim P. Boyarskiy and Regina M. Islamova
Polymers 2024, 16(20), 2887; https://doi.org/10.3390/polym16202887 - 14 Oct 2024
Viewed by 336
Abstract
Cobaltocenium-containing (co)polysiloxanes (Cc-PDMSs) with terminal and side groups were synthesized by the reaction of catalyst-free hydroamination between ethynylcobaltocenium hexafluorophosphate and polysiloxanes comprising amino moieties as terminal and side groups. The conversion of NH2 groups in the polymers reaches 85%. The obtained (co)polysiloxanes [...] Read more.
Cobaltocenium-containing (co)polysiloxanes (Cc-PDMSs) with terminal and side groups were synthesized by the reaction of catalyst-free hydroamination between ethynylcobaltocenium hexafluorophosphate and polysiloxanes comprising amino moieties as terminal and side groups. The conversion of NH2 groups in the polymers reaches 85%. The obtained (co)polysiloxanes “gelate” due to an increase in their molecular weight by approx. 30 times, when stored at room temperature over one week. “Gelated” Cc-PDMSs remain soluble in most polar solvents. The structure of Cc-PDMSs and the mechanism of “gelation” were established by 1H, 13C{1H}, 29Si{1H}, 19F{1H}, 31P{1H} nuclear magnetic resonance, infrared, ultraviolet–visible, and X-ray photoelectron spectroscopies. As determined by cyclic voltammetry, Cc-PDMSs possess redox properties (CoII/CoIII transitions at E1/2 = −1.8 and −1.3 V before and after “gelation”, respectively). This synthetic approach allows to increase the molecular weights of the synthesized polysiloxanes functionalized with cobaltocenium groups easily, leading to their higher film-forming ability, which is desirable for some electronic applications. Cc-PDMSs can be utilized as redox-active polymer films in modified electrodes, electrochromic devices, redox-active coatings, and components for batteries. Full article
Show Figures

Figure 1

Figure 1
<p>Synthetic schemes for Cc-APDMSs (<b>a</b>) and P(Cc-AMS-<span class="html-italic">co</span>-DMS) (<b>b</b>).</p>
Full article ">Figure 2
<p>Changes in <sup>1</sup>H NMR of P(Cc-AMS-<span class="html-italic">co</span>-DMS) when being stored in acetone-<span class="html-italic">d</span><sub>6</sub> in air (<b>a</b>), and changes in <sup>29</sup>Si{<sup>1</sup>H} NMR of Cc-APDMS850 when being stored in acetone-<span class="html-italic">d</span><sub>6</sub> in air (<b>b</b>), plausible mechanism of condensation reaction between enamine groups in cobaltocenium-terminated (co)polymers (<b>c</b>).</p>
Full article ">Figure 3
<p>Cyclic voltammograms of Cc-APDMS850: freshly synthesized (<b>a</b>) and after being stored in air for 3 months (<b>b</b>), recorded in 0.1 M Et<sub>4</sub>NBF<sub>4</sub> CH<sub>3</sub>CN solution at potential scan rates of 0.05–50 V·s<sup>−1</sup>.</p>
Full article ">
15 pages, 6318 KiB  
Article
Snowflake Iron Oxide Architectures: Synthesis and Electrochemical Applications
by Anna Kusior, Olga Waś, Zuzanna Liczberska, Julia Łacic and Piotr Jeleń
Molecules 2024, 29(20), 4859; https://doi.org/10.3390/molecules29204859 - 14 Oct 2024
Viewed by 468
Abstract
The synthesis and characterization of iron oxide nanostructures, specifically snowflake architecture, are investigated for their potential applications in electrochemical sensing systems. A Raman spectroscopy analysis reveals phase diversity in the synthesized powders. The pH of the synthesis affects the formation of the hematite [...] Read more.
The synthesis and characterization of iron oxide nanostructures, specifically snowflake architecture, are investigated for their potential applications in electrochemical sensing systems. A Raman spectroscopy analysis reveals phase diversity in the synthesized powders. The pH of the synthesis affects the formation of the hematite (α-Fe2O3) and goethite (α-FeOOH). Scanning electron microscopy (SEM) images confirm the distinct morphologies of the particles, which are selectively obtained through recrystallization during the elongated reaction time. An electrochemical analysis demonstrates the differing behaviors of the particles, with synthesis pH affecting the electrochemical activity and surface area differently for each shape. Cyclic voltammetry measurements reveal reversible dopamine detection processes, with snowflake iron oxide showing lower detection limits than a mixture of snowflakes and cube-like particles. This research contributes to understanding the relationship between iron oxide nanomaterials’ structural, morphological, and electrochemical properties. It offers practical insights into their potential applications in sensor technology, particularly dopamine detection, with implications for biomedical and environmental monitoring. Full article
(This article belongs to the Special Issue Nanomaterials for Electrocatalytic Applications)
Show Figures

Figure 1

Figure 1
<p>The SEM images of the obtained iron oxide nanostructures (S series) after 24, 48, and 72 h from the solution with 8.5 pH.</p>
Full article ">Figure 2
<p>The SEM images of the obtained iron oxide nanostructures (P series) after 24, 48, and 72 h from the solution with 12 pH.</p>
Full article ">Figure 3
<p>The XRD analysis of the iron oxide-based material obtained (<b>a</b>) at pH = 8.5 and (<b>b</b>) at pH = 12. Caption 1 is assigned to the tetrairon(III) hexacyanoferrate(II) 9.3-hydrate 4.7-(dideuriohydrate), and 2 to tetrairon(III) tris(hexacyanoferrate(II)) tetradecahydrate.</p>
Full article ">Figure 4
<p>The Raman spectra of the obtained iron oxide nanostructures (S series) after 24, 48, and 72 h from the solution with 8.5 pH. The scheme of the obtained structures at various synthesis times.</p>
Full article ">Figure 5
<p>The Raman spectra of the obtained iron oxide nanostructures (P series) after 24, 48, and 72 h from the solution with 12 pH. The scheme of the obtained structures at various synthesis times.</p>
Full article ">Figure 6
<p>Current–voltage dependence for the modified carbon electrodes (<b>a</b>) in the 0.1 M KCl + 3 mM [Fe(CN)<sub>6</sub>]<sup>3−/4−</sup>, and comparison with the SPE (<b>b</b>). Data were recorded at various scan rates from 10 to 2000 mVs<sup>−1</sup>.</p>
Full article ">Figure 7
<p>Current–voltage dependence for the modified carbon electrodes in the 0.1 M KCl + 3 mM [Fe(CN)<sub>6</sub>]<sup>3−/4−</sup> solution. Data were recorded at various scan rates from 10 to 2000 mVs<sup>−1</sup>.</p>
Full article ">Figure 8
<p>(<b>a</b>) The plot of the anodic peak current as a function of the square root of the scan rate ν<sup>0.5</sup> for modified electrodes with (<b>b</b>) electrochemically active surface area, EASA, compared to the SPE electrode.</p>
Full article ">Figure 9
<p>(<b>a</b>) Comparison of cyclic voltammograms in various dopamine concentrations (from 1 to 10 mM solution in PBS environment) for modified screen-printed electrodes by the obtained iron oxide powders. (<b>b</b>) The recorded data at 5 mM DA for the pure SPE and S2-modified electrode. Data were recorded at 50 mVs<sup>−1</sup>.</p>
Full article ">Figure 10
<p>(<b>a</b>) Calibration curves for series S modified screen-printed electrodes in the presence of various concentrations of dopamine (from 1 to 10 mM solution) with (<b>b</b>) LOD and LOQ parameters of the SPE-modified electrodes. The square, circle, and triangle symbols correspond to the dopamine changes for S1, S2, and S3 samples.</p>
Full article ">Figure 11
<p>Comparison of cyclic voltammograms in various dopamine concentrations (from 1 to 10 mM solution in PBS environment) for modified screen-printed electrodes by the obtained iron oxide powders. Data were recorded at 50 mVs<sup>−1</sup>.</p>
Full article ">
13 pages, 3126 KiB  
Article
Graphite–Phosphate Composites: Structure and Voltammetric Investigations
by Simona Rada, Alexandra Barbu Gorea and Eugen Culea
Materials 2024, 17(20), 5000; https://doi.org/10.3390/ma17205000 - 12 Oct 2024
Viewed by 549
Abstract
The utilization of lithium-ion batteries (LIBs) is increasing sharply with the increasing use of mobile phones, laptops, tablets, and electric vehicles worldwide. Technologies are required for the recycling and recovery of spent LIBs. In the context of the circular economy, it is urgent [...] Read more.
The utilization of lithium-ion batteries (LIBs) is increasing sharply with the increasing use of mobile phones, laptops, tablets, and electric vehicles worldwide. Technologies are required for the recycling and recovery of spent LIBs. In the context of the circular economy, it is urgent to search for new methods to recycle waste graphite that comes from the retired electrode of LIBs. The conversion of waste graphite into other products, such as new electrodes, in the field of energy devices is attractive because it reduces resource waste and processing costs, as well as preventing environmental pollution. In this paper, new electrode materials were prepared using waste anode graphite originating from a spent mobile phone battery with an xBT·0.1C12H22O11·(0.9-x)(NH4)2HPO4 composition, where x = 0–50 weight% BT from the anodic active mass of the spent phone battery (labeled as BT), using the melt quenching method. Analysis of the diffractograms shows the graphite crystalline phase with a hexagonal structure in all prepared samples. The particle sizes decrease by adding a higher BT amount in the composites. The average band gap is 1.32 eV (±0.3 eV). A higher disorder degree in the host network is the main factor responsible for lower band gap values. The prepared composites were tested as electrodes in an LIB or a fuel cell, achieving an excellent electrochemical performance. The voltammetric studies indicate that doping with 50% BT is the most suitable for applications as electrodes in LIBs and fuel cells. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Spent phone battery; (<b>b</b>) punctured phone battery; (<b>c</b>,<b>d</b>) the disassembled components of the spent phone battery; (<b>e</b>) images of the obtained samples in the xBT·0.1C<sub>12</sub>H<sub>22</sub>O<sub>11</sub>·(0.9-x) (NH<sub>4</sub>)<sub>2</sub>HPO<sub>4</sub> composition, where x = 0–50 weight% BT.</p>
Full article ">Figure 2
<p>X-ray diffractograms of the (<b>a</b>) spent anodic powder (noted with BT) and (NH<sub>4</sub>)<sub>2</sub>HPO<sub>4</sub> glassy; (<b>b</b>–<b>d</b>) prepared samples in the xBT·0.1C<sub>12</sub>H<sub>22</sub>O<sub>11</sub>·(0.9-x)(NH<sub>4</sub>)<sub>2</sub>HPO<sub>4</sub> composition, where x = 0–50 weight% BT.</p>
Full article ">Figure 3
<p>FTIR spectra of the prepared composites in the region between (<b>a</b>,<b>b</b>) 450 and 4000 cm<sup>−1</sup>; (<b>c</b>) 400 and 1500 cm<sup>−1</sup>.</p>
Full article ">Figure 4
<p>EPR spectra of the spent BT and of the graphite–phosphate composites. (<b>a</b>) Spent BT, x = 15 and 20% BT; (<b>b</b>) x = 30 and 50% BT; (<b>c</b>) x = 15–50% BT; (<b>d</b>) x = 15% BT.</p>
Full article ">Figure 5
<p>(<b>a</b>) The UV-Vis data of the spent BT powder and prepared samples; (<b>b</b>) the compositional evolution of optical gap energy, E<sub>g</sub>, for direct transitions with <span class="html-italic">n</span> = 1/2 of the prepared samples.</p>
Full article ">Figure 6
<p>Cyclic voltammograms of the electrode materials with xBT·0.1C<sub>12</sub>H<sub>22</sub>O<sub>11</sub>·(0.9-x) (NH<sub>4</sub>)<sub>2</sub>HPO<sub>4</sub> chemical formula, where (<b>a</b>) x = 20, 30, and 50 weight% BT (BT—cathodic active mass of the spent phone battery) in 0.15 M Li<sub>2</sub>CO<sub>3</sub> and 0.1 M Na<sub>2</sub>CO<sub>3</sub> mix solution and (<b>b</b>) x = 30 and 50 weight% BT in 1 M KOH solution.</p>
Full article ">Figure 7
<p>Cyclic voltammograms of the electrode materials with x = 50 weight% BT in (<b>a</b>) a 1 M KOH electrolyte solution at varied scan rates and (<b>b</b>) in KOH (1 M concentration) and Li<sub>2</sub>CO<sub>3</sub>/Na<sub>2</sub>CO<sub>3</sub> electrolyte solutions.</p>
Full article ">Figure 8
<p>(<b>a</b>) Cyclic voltammograms scanned after two cycles for the prepared electrode materials in KOH and Li<sub>2</sub>CO<sub>3</sub>/Na<sub>2</sub>CO<sub>3</sub> electrolyte solutions; (<b>b</b>) Nyquist plot of the complex impedance of the electrode material with x = 50% BT obtained in the KOH electrolyte solution.</p>
Full article ">
25 pages, 11394 KiB  
Article
Electroanalytical Studies on Codeposition of Cobalt with Ruthenium from Acid Chloride Baths
by Iwona Dobosz and Ewa Rudnik
Coatings 2024, 14(10), 1301; https://doi.org/10.3390/coatings14101301 - 11 Oct 2024
Viewed by 506
Abstract
The aim of this study was to systematically analyze the influence of potential and the Co(II)–Ru(III) molar ratio on the electrochemical behavior of the Co–Ru system during codeposition from acidic chloride electrolytes. The equilibrium speciation of the baths was investigated spectrophotometrically and compared [...] Read more.
The aim of this study was to systematically analyze the influence of potential and the Co(II)–Ru(III) molar ratio on the electrochemical behavior of the Co–Ru system during codeposition from acidic chloride electrolytes. The equilibrium speciation of the baths was investigated spectrophotometrically and compared with theoretical calculations based on the stability constants of Co(II) and Ru(III) complexes. The codeposition of the metals was characterized using electroanalytical methods, including cyclic voltammetry, chronoamperometry, and anodic stripping linear voltammetry. The alloys obtained at different potentials were analyzed for their elemental composition (EDS, mapping), phase composition (XRD), and surface morphology (SEM). The morphology and composition of the alloys were mainly dependent on the deposition potential, which controlled the cobalt incorporation. Ruthenium–rich alloys were produced at potentials of −0.6 V and −0.7 V (vs. SCE). In these conditions, cobalt anomalously codeposited due to the formation of the CoOH+ intermediate, triggered by the intense hydrogen evolution on the ruthenium sublayer. Bulk cobalt electrodeposition began at a potential of around −0.8 V, resulting in the formation of cobalt-rich alloys. The early stages of the electrodeposition were investigated using different nucleation models. A transition from 2D progressive nucleation to 3D instantaneous nucleation at around −0.8 V was identified as being caused by cobalt incorporation. This was well correlated with electroanalytical data, partial polarization curves of alloy deposition, elemental mapping analysis, and the structure of the deposits. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of some Co(II) (<b>a</b>) and Ru(III) (<b>b</b>) complexes. Actual structures based on [<a href="#B27-coatings-14-01301" class="html-bibr">27</a>,<a href="#B32-coatings-14-01301" class="html-bibr">32</a>].</p>
Full article ">Figure 2
<p>UV–Vis absorption spectra of Co(II)–Ru(III) (<b>a</b>), Ru(III) (<b>b</b>), and Co(II) (<b>c</b>) acid chloride solutions. Co(II)–Ru(III) molar concentration ratio 10:1.</p>
Full article ">Figure 3
<p>Fractions of Co(II) and Ru(III) species in acid chloride solutions (3 M Cl<sup>−</sup>, pH 1): (<b>a</b>) 1 M Co(II) + 0.01 M Ru(III) and (<b>b</b>) 0.1 M Co(II) + 0.01 M Ru(III). For simplicity, water molecules are omitted in the complex formulas.</p>
Full article ">Figure 4
<p>Cyclic voltammetry curves registered in (<b>a</b>) Ru(III), (<b>b</b>) Co(II), and (<b>c</b>) Co(II)–Ru(III) acid chloride solutions: 1 M Co(II) and/or 0.01 M Ru(III).</p>
Full article ">Figure 5
<p>Cyclic voltammetry curves registered in (<b>a</b>) Ru(III), (<b>b</b>) Co(II), and (<b>c</b>) Co(II)–Ru(III) acid chloride solutions: 0.1 M Co(II) and/or 0.01 M Ru(III).</p>
Full article ">Figure 6
<p>Cyclic voltammetry curves for different switch potentials (from −0.8 V to −1.3 V) registered in Co(II)–Ru(III) acid chloride solutions with (<b>a</b>) 1 M Co(II) and (<b>b</b>) 0.1 M Co(II).</p>
Full article ">Figure 7
<p>Chronoamperometric curves registered in Co(II)–Ru(III) acid chloride solutions with (<b>a</b>) 1 M Co(II) and (<b>b</b>) 0.1 M Co(II). Lower plots show details for deposition potentials between −0.60 V and −0.85 V.</p>
Full article ">Figure 8
<p>Anodic sweep linear voltammetry curves registered in Co(II)–Ru(III) acid chloride solutions with (<b>a</b>) 1 M Co(II) and (<b>b</b>) 0.1 M Co(II).</p>
Full article ">Figure 9
<p>Anodic to cathodic charge ratios in Co(II)–Ru(III) solutions registered during cathodic accumulation step (40 s) and anodic sweep linear voltammetry.</p>
Full article ">Figure 10
<p>Schemes of nuclei formation and growth: (<b>a</b>) two-dimensional nuclei and (<b>b</b>) three-dimensional nuclei.</p>
Full article ">Figure 11
<p>Normalized i–t curves for 2D (upper) and 3D (lower) instantaneous IN and progressive PN nucleation and growth for deposition from Co(II)–Ru(III) acid chloride solutions with (<b>a</b>) 1 M Co(II) and (<b>b</b>) 0.1 M Co(II). Dashed and dotted lines represent model reference lines.</p>
Full article ">Figure 12
<p>Influence of deposition potential on metal content in deposits: (<b>a</b>) ruthenium and (<b>b</b>) cobalt.</p>
Full article ">Figure 13
<p>Influence of deposition potential on (<b>a</b>) deposit mass and (<b>b</b>) cathodic current efficiency.</p>
Full article ">Figure 14
<p>Morphology of deposits produced from 1 M Co(II)–Ru(III) bath at potentials (<b>a</b>) −0.8 V, (<b>b</b>) −1.1 V, and (<b>c</b>) −1.3 V.</p>
Full article ">Figure 15
<p>Morphology of deposits produced from 0.1 M Co(II)–Ru(III) bath at potentials (<b>a</b>) −0.6 V, (<b>b</b>) −0.7 V, (<b>c</b>) −0.8 V, (<b>d</b>) −0.9 V, (<b>e</b>) −1.0 V, (<b>f</b>) −1.1 V, and (<b>g</b>) −1.3 V.</p>
Full article ">Figure 16
<p>Distribution of elements (mapping) on surface of deposit produced from 0.1 M Co(II)–Ru(III) bath at potentials (<b>a</b>) −0.7 V, (<b>b</b>) −0.9 V, and (<b>c</b>) −1.1 V.</p>
Full article ">Figure 17
<p>X-ray diffraction patterns of deposits produced at different potentials from Co(II)–Ru(III) baths: (<b>a</b>) 0.1 M Co(II), (<b>b</b>) 1 M Co(II), and (<b>c</b>) substrates Au—glass covered with gold layer; H—aluminum holder. XRD standard cards: Co—03–065–9722, Ru—00–001–1256, and Co–Ru—01–071–7425.</p>
Full article ">Figure 18
<p>Partial polarization curves calculated for alloy deposition from the 0.1 M Co(II)–Ru(III) bath shown in two coordination systems: (<b>a</b>) E–I and (<b>b</b>) E–log I.</p>
Full article ">
17 pages, 4252 KiB  
Article
Novel Biochar-Modified ZIF-8 Metal–Organic Frameworks as a Potential Material for Optoelectronic and Electrochemical Energy Storage Applications
by Sarah Al-atawi, Meshari M. Aljohani, Taymour A. Hamdalla, S. A. Al-Ghamdi, Abdulrhman M. Alsharari and Syed Khasim
Catalysts 2024, 14(10), 705; https://doi.org/10.3390/catal14100705 - 10 Oct 2024
Viewed by 546
Abstract
Herein, we report the preparation of nanocomposites using activated biochar derived from rice husk (RHBC) by doping with a metal–organic framework, namely the zeolitic imidazolate framework (ZIF-8). The morphological and structural characterization of the prepared nanocomposite was performed using SEM, BET, XRD, FTIR, [...] Read more.
Herein, we report the preparation of nanocomposites using activated biochar derived from rice husk (RHBC) by doping with a metal–organic framework, namely the zeolitic imidazolate framework (ZIF-8). The morphological and structural characterization of the prepared nanocomposite was performed using SEM, BET, XRD, FTIR, TGA, and UV–Vis spectroscopy. The average particle sizes as observed from SEM micrographs for ZIF-8 and ZIF-8@RHBC were 67 nm and 78 nm, respectively. The BET surface analysis of the ZIF-8@RHBC composite showed a value of 308 m2/g and a pore diameter of about 42.56 A°. The inclusion of RHBC in ZIF-8 resulted in a 4% increase in the optical band gap and a 5% increase in the optical conductivity. The electrochemical properties of this nanocomposite were investigated through cyclic voltammetry, and it was observed that ZIF-8@RHBC showed improved CV curves in comparison to bare ZIF-8. The specific capacitance of ZIF-8@RHBC was significantly enhanced from 348 F/g to 452 F/g at a 1 A/g current density after incorporating ZIF-8 into the RHBC matrix. The formation of a mesoporous structure in the ZIF-8@RHBC composite contributed to the improved diffusion rate at the electrode surface, resulting in excellent electrochemical features in the composite. Furthermore, the EIS studies confirmed the reduced charge transfer resistance and increased conduction at the electrode surface in the case of the ZIF-8@RHBC composite. Owing to the ease of its green synthesis and its excellent structural and morphological features and optical and electrochemical properties, this ZIF@RHBC nanocomposite could represent a novel multifunctional material to be used in optoelectronics and energy storage applications. Full article
(This article belongs to the Special Issue Two-Dimensional Materials in Photo(electro)catalysis)
Show Figures

Figure 1

Figure 1
<p>Functional groups for ZIF-8 and ZIF-8@RHBC using FTIR spectra.</p>
Full article ">Figure 2
<p>XRD for ZIF-8 and ZIF-8@RHBC.</p>
Full article ">Figure 3
<p>SEM micrographs of ZIF-8 and ZIF-8@RHBC.</p>
Full article ">Figure 4
<p>TGA analysis for ZIF and ZIF-8@RHBC.</p>
Full article ">Figure 5
<p>(<b>A</b>) The optical transmission and reflection of ZIF-8 and ZIF-8@RHBC. (<b>B</b>) The dependence of (<span class="html-italic">αE</span>)<sup>1/2</sup> on the photon energy (<span class="html-italic">E</span>) for ZIF-8 and ZIF-8@RHBC.</p>
Full article ">Figure 6
<p>(<b>A</b>) Refractive index spectra and (<b>B</b>) absorption spectra for ZIF-8 and ZIF-8@RHBC (<b>B</b>).</p>
Full article ">Figure 7
<p>(<b>A</b>) Plot of ln(<span class="html-italic">α</span>) (optical conductivity) as a function of photon energy (<span class="html-italic">E</span>) for ZIF-8 and ZIF-8@RHBC. (<b>B</b>) Dependence of thermal emissivity (<span class="html-italic">ε</span><sub>th</sub>) on wavelength (λ) for ZIF-8 and ZIF-8@RHBC.</p>
Full article ">Figure 8
<p>(<b>A</b>) Cyclic voltammetry (CV) curves for ZIF-8 and ZIF-8@RHBC at a scanning rate of 2 mVs<sup>−1</sup> (in aqueous 1M LiPF<sub>6</sub> electrolyte). (<b>B</b>) Specific capacitance as a function of the scan rate at a constant current density of 0.1 Ag<sup>−1</sup> for ZIF-8 and ZIF-8@RHB.</p>
Full article ">Figure 9
<p>(<b>A</b>) Specific capacitance as a function of the current density for ZIF-8 and ZIF-8@RHBC. (<b>B</b>) Nyquist plot for ZIF-8 and ZIF-8@RHBC.</p>
Full article ">Figure 10
<p>Galvanostatic charge–discharge (GCD) of activated ZIF-8 and ZIF@RHBC.</p>
Full article ">Figure 11
<p>GCD results under 3000 cycles: cyclic stability of (<b>a</b>) ZIF-8 and (<b>b</b>) ZIF@RHBC.</p>
Full article ">Scheme 1
<p>The energy gap of ZIF-8@RHBC.</p>
Full article ">Scheme 2
<p>The experimental setup used in the present study.</p>
Full article ">Scheme 3
<p>The chemical structures of (<b>a</b>) ZIF-8 and (<b>b</b>) activated biochar.</p>
Full article ">
13 pages, 4598 KiB  
Article
The Transformative Role of Nano-SiO2 in Polymer Electrolytes for Enhanced Energy Storage Solutions
by S. Jayanthi, M. Vahini, S. Karthickprabhu, A. Anusuya, N. Karthik, K. Karuppasamy, Tholkappiyan Ramachandran, A. Nichelson, M. Mahendran, B. Sundaresan and Dhanasekaran Vikraman
Processes 2024, 12(10), 2174; https://doi.org/10.3390/pr12102174 - 7 Oct 2024
Viewed by 772
Abstract
In lithium–polymer batteries, the electrolyte is an essential component that plays a crucial role in ion transport and has a substantial impact on the battery’s overall performance, stability, and efficiency. This article presents a detailed study on developing nanostructured composite polymer electrolytes (NCPEs), [...] Read more.
In lithium–polymer batteries, the electrolyte is an essential component that plays a crucial role in ion transport and has a substantial impact on the battery’s overall performance, stability, and efficiency. This article presents a detailed study on developing nanostructured composite polymer electrolytes (NCPEs), prepared using the solvent casting technique. The materials selected for this investigation include poly(vinyl chloride) (PVC) as the host polymer, lithium bromide (LiBr) as the salt, and silica (SiO2) as the nanofiller. The addition of nano-SiO2 dramatically enhanced the ionic conductivity of the electrolytes, with the highest value of 6.2 × 10−5 Scm−1 observed for the sample containing 7.5 wt% nano-SiO2. This improvement is attributed to an increased amorphicity resulting from the interactions between the polymer, salt, and filler components. A structural analysis of the prepared NCPEs using X-ray diffraction revealed the presence of both crystalline and amorphous phases, further validating the enhanced ionic transport. Additionally, the thermal stability of the NCPEs was found to be excellent, withstanding temperatures up to 334 °C, thereby reinforcing their potential application in lithium–polymer batteries. This work explores the electrochemical performance of a fabricated lithium-ion-conducting primary electrochemical cell (Zn + ZnSO4·7H2O|PVC: LiBr: SiO2|PbO2 + V2O5), which demonstrated an open circuit voltage of 2.15 V. The discharge characteristics of the fabricated cell were thoroughly studied, showcasing the promising potential of these NCPEs. With the support of superior morphological and electrical properties, as-prepared electrolytes offer an effective pathway for future advancements in lithium–polymer battery technology, making them a highly viable candidate for enhanced energy storage solutions. Full article
(This article belongs to the Special Issue High-Efficiency Nanomaterials Synthesis and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram for solution-casting method; (<b>b</b>) X-ray diffractograms of the prepared films (i) pure PVC, (ii) PVC:LiBr, (iii) Nano SiO<sub>2</sub> and (iv) PVC:LiBr:SiO<sub>2</sub> (7.5 wt%); (<b>c</b>) SEM image of pure PVC, (<b>d</b>) PVC:LiBr, and (<b>e</b>) PVC:LiBr:SiO<sub>2</sub> (7.5 wt%).</p>
Full article ">Figure 2
<p>(<b>a</b>) Nyquist plot of PVC:LiBr:SiO<sub>2</sub> system at ambient conditions; (<b>b</b>) variation in dielectric permittivity of real and (<b>c</b>) imaginary parts concerning frequency for PVC:LiBr and PVC:LiBr:SiO<sub>2</sub> added to system at ambient temperature; (<b>d</b>) frequency dependence of M′ (storage modulus) and (<b>e</b>) M″ (loss modulus) for the PVC:LiBr:SiO<sub>2</sub> system at room temperature; (<b>f</b>) plot of current vs. time for PVC:LiBr and PVC:LiBr:SiO<sub>2</sub> (7.5 wt%) (maximum ionic conductivity) systems.</p>
Full article ">Figure 3
<p>(<b>a</b>) Conductivity and migration vs. nano-SiO<sub>2</sub> concentration; (<b>b</b>) conductivity and free ion concentration as a function of nano-SiO<sub>2</sub> concentration; (<b>c</b>) conductivity and diffusion coefficient vs. nano-SiO<sub>2</sub> concentration; (<b>d</b>) TG/DTA graph for PVC:LiBr:SiO<sub>2</sub> (7.5 wt%) system.</p>
Full article ">Figure 4
<p>(<b>a</b>) Linear Sweep Voltammetry of higher-ionic-conducting PVC/LiBr/nano-SiO<sub>2</sub> system at room temperature; (<b>b</b>) Cyclic Voltammetry of higher-ionic-conducting PVC/LiBr/nano-SiO<sub>2</sub> system at room temperature.</p>
Full article ">Figure 5
<p>(<b>a</b>) Arrangement of electrochemical cell; (<b>b</b>) photo of the open circuit voltage of the fabricated electrochemical cell at room temperature; (<b>c</b>) open circuit voltage of the fabricated electrochemical cell at room temperature; (<b>d</b>) discharge characterization of electrochemical cell with 1 MΩ.</p>
Full article ">
14 pages, 4146 KiB  
Article
Acridone Derivatives for Near-UV Radical Polymerization: One-Component Type II vs. Multicomponent Behaviors
by Adel Noon, Francesco Calogero, Andrea Gualandi, Hiba Hammoud, Tayssir Hamieh, Joumana Toufaily, Fabrice Morlet-Savary, Michael Schmitt, Pier Giorgio Cozzi and Jacques Lalevée
Molecules 2024, 29(19), 4715; https://doi.org/10.3390/molecules29194715 (registering DOI) - 5 Oct 2024
Viewed by 548
Abstract
In this work, two novel acridone-based photoinitiators were designed and synthesized for the free radical polymerization of acrylates with a light-emitting diode emitting at 405 nm. These acridone derivatives were employed as mono-component Type II photoinitiators and as multicomponent photoinitiating systems in the [...] Read more.
In this work, two novel acridone-based photoinitiators were designed and synthesized for the free radical polymerization of acrylates with a light-emitting diode emitting at 405 nm. These acridone derivatives were employed as mono-component Type II photoinitiators and as multicomponent photoinitiating systems in the presence of an iodonium salt or an amine synergist (EDB) in which they achieved excellent polymerization initiating abilities and high final conversions of the acrylate group. Photoinitiation mechanisms through which reactive species are produced were investigated employing different complementary techniques including steady-state photolysis, steady-state fluorescence, cyclic voltammetry, UV–visible absorption spectroscopy, and electron spin resonance spectroscopy. Finally, these molecules were also used in the direct laser writing process for the fabrication of 3D objects. Full article
(This article belongs to the Special Issue Synthesis and Application of Photoactive Compounds)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>UV–visible absorption properties of compounds Bn-Acr and DPM-Acr in acetonitrile.</p>
Full article ">Figure 2
<p>Photopolymerization profiles of TA (acrylate function conversion vs. irradiation time) in laminate (thickness = 25 μm) upon exposure to an LED (λ = 405 nm) in the presence of PIs (1% <span class="html-italic">w</span>/<span class="html-italic">w</span>), PIs/Iod (1%/1% <span class="html-italic">w</span>/<span class="html-italic">w</span>), PIs/EDB (1%/1% <span class="html-italic">w</span>/<span class="html-italic">w</span>), Iod alone (1% <span class="html-italic">w</span>/<span class="html-italic">w</span>), and EDB alone (1% <span class="html-italic">w</span>/<span class="html-italic">w</span>). The irradiation starts at t = 10 s.</p>
Full article ">Figure 3
<p>Photolysis of (<b>A</b>) Bn-Acr with Iod (10<sup>−2</sup> M) and (<b>B</b>) DPM-Acr with Iod (10<sup>−2</sup> M) in acetonitrile using LED at λ = 385 nm.</p>
Full article ">Figure 4
<p>Fluorescence quenching study of Bn-Acr (<b>A</b>) and DPM-Acr (<b>B</b>) by Iod in acetonitrile; fluorescence quenching study of Bn-Acr (<b>C</b>) and DPM-Acr (<b>D</b>) by EDB in acetonitrile.</p>
Full article ">Figure 5
<p>ESR spectra for Bn-Acr recorded in the presence of PBN and tert-butylbenzene with an LED@405nm: (<b>A</b>) before and after irradiation; (<b>B</b>) experimental and simulated spectra observed after irradiation (at t = 180 s).</p>
Full article ">Figure 6
<p>ESR spectra for Bn-Acr/Iod recorded in the presence of PBN and tert-butylbenzene with an LED@405nm: (<b>A</b>) before and after irradiation; (<b>B</b>) experimental and simulated spectra observed after irradiation (at t = 60 s).</p>
Full article ">Figure 7
<p>Three-dimensional patterns for compound DPM-Acr with Iod (0.1%/1% <span class="html-italic">w</span>/<span class="html-italic">w</span>) in TA were produced after exposure to a laser diode at 405 nm and were analyzed by numerical microscopy.</p>
Full article ">Scheme 1
<p>Chemical structures and abbreviations of acridone derivatives: previously studied (A-2DPA, A-2PTz) in [<a href="#B26-molecules-29-04715" class="html-bibr">26</a>], and the newly investigated ones in this work (Bn-Acr and DPM-Acr).</p>
Full article ">Scheme 2
<p>Synthesis of the target PIs.</p>
Full article ">Scheme 3
<p>The proposed photochemical mechanism for acridone-based PIs as one-component Type II PIs.</p>
Full article ">Scheme 4
<p>Chemical structures and abbreviations of the benchmark monomer TA, and the additives (Iod) and (EDB).</p>
Full article ">
19 pages, 9166 KiB  
Article
Development of Fluorine-Free Electrolytes for Aqueous-Processed Olivine-Type Phosphate Cathodes
by Claudia Limachi, Klaudia Rogala, Marek Broszkiewicz, Marta Cabello, Leszek Niedzicki, Michel Armand and Władysław Wieczorek
Molecules 2024, 29(19), 4698; https://doi.org/10.3390/molecules29194698 - 4 Oct 2024
Viewed by 484
Abstract
Environmental impacts and resource availability are significant concerns for the future of lithium-ion batteries. This study focuses on developing novel fluorine-free electrolytes compatible with aqueous-processed cobalt-free cathode materials. The new electrolyte contains lithium 1,1,2,3,3-pentacyanopropenide (LiPCP) salt. After screening various organic carbonates, a mixture [...] Read more.
Environmental impacts and resource availability are significant concerns for the future of lithium-ion batteries. This study focuses on developing novel fluorine-free electrolytes compatible with aqueous-processed cobalt-free cathode materials. The new electrolyte contains lithium 1,1,2,3,3-pentacyanopropenide (LiPCP) salt. After screening various organic carbonates, a mixture of 30:70 wt.% ethylene carbonate and dimethyl carbonate was chosen as the solvent. The optimal salt concentration, yielding the highest conductivity of 9.6 mS·cm−1 at 20 °C, was 0.8 mol·kg−1. Vinylene carbonate was selected as a SEI-stabilizing additive, and the electrolyte demonstrated stability up to 4.4 V vs. Li+/Li. LiFePO4 and LiMn0.6Fe0.4PO4 were identified as suitable cobalt-free cathode materials. They were processed using sodium carboxymethyl cellulose as a binder and water as the solvent. Performance testing of various cathode compositions was conducted using cyclic voltammetry and galvanostatic cycling with the LiPCP-based electrolyte and a standard LiPF6-based one. The optimized cathode compositions, with an 87:10:3 ratio of active material to conductive additive to binder, showed good compatibility and performance with the new electrolyte. Aqueous-processed LiFePO4 and LiMn0.6Fe0.4PO4 achieved capacities of 160 mAh·g−1 and 70 mAh·g−1 at C/10 after 40 cycles, respectively. These findings represent the first stage of investigating LiPCP for the development of greener and more sustainable lithium-ion batteries. Full article
(This article belongs to the Special Issue A Perspective on Novel Electrochemical Capacitors and Batteries)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structure, lithium 1,1,2,3,3-pentacyanopropenide (LiPCP), novel fluorine-free lithium salt for electrolytes.</p>
Full article ">Figure 2
<p>(<b>a</b>) Ionic conductivity of LiPCP and LiPF<sub>6</sub> in various organic carbonate solvent mixtures from 0 to 50 °C. (<b>b</b>) Ionic conductivity of LiPCP in EC:DMC (30:70 wt.%) at concentrations in the range of 0.1–1.2 mol·kg<sup>−1</sup> from 0 to 50 °C.</p>
Full article ">Figure 3
<p>Linear sweep voltammetry (LSV) of 0.8 mol·kg<sup>−1</sup> LiPCP-based electrolytes without additives and with various electrolyte additive concentrations, with the Pt disc as the working electrode and the Li metal disc as the reference electrode in a Swagelok-cell system, at a scan rate of 0.5 mV·s<sup>−1</sup>. For comparison purposes, data on a LiPF<sub>6</sub> electrolyte are included. The inset shows the zoomed area.</p>
Full article ">Figure 4
<p>Passive layer resistances (R<sub>p</sub>) and charge transfer resistances (R<sub>ct</sub>) for Li|electrolyte|Li cells with three different electrolytes.</p>
Full article ">Figure 5
<p>(<b>a</b>) CV voltammograms (1st cycle) of corresponding Li/LFP cells at different LFP compositions, with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) and 5 wt.% of VC at a scan rate of 0.5 mV·s<sup>−1</sup> at room temperature. (<b>b</b>) CV voltammograms (1st cycle) of corresponding Li/LFP2 cells with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) with 5 wt.% VC and with 5 wt.% VC plus 10 wt.% AN. For comparison purposes, data with the reference LiPF<sub>6</sub> electrolyte are shown. Tests were carried out with a scan rate of 0.5 mV·s<sup>−1</sup> at room temperature.</p>
Full article ">Figure 6
<p>(<b>a</b>) CV voltammograms (1st cycle) of corresponding Li/LMFP cells at different LMFP compositions, with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) and 5 wt.% of VC at a scan rate of 0.5 mV·s<sup>−1</sup> at room temperature. (<b>b</b>) CV voltammograms (1st cycle) of Li/LMFP2 cells with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) with 5 wt.% of VC and with 5 wt.% VC plus 10 wt.% AN. For comparison purposes, data with a reference LiPF<sub>6</sub> electrolyte are shown. Tests were carried out with a scan rate of 0.5 mV·s<sup>−1</sup> at room temperature.</p>
Full article ">Figure 7
<p>Rate performance of Li/LiFePO<sub>4</sub> cells with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) and 5 wt.% of VC. Potential range: 2.5–3.9 V vs. Li<sup>+</sup>/Li. (<b>a</b>) Discharge capacities in mAh·g<sup>−1</sup>. (<b>b</b>) Coulombic efficiency in (%).</p>
Full article ">Figure 8
<p>Rate performance and coulombic efficiency of Li/LiFePO<sub>4</sub> (LFPB) cells with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) and 5 wt.% of VC; and with 1.0 mol·kg<sup>−1</sup> LiPF<sub>6</sub> in EC:DMC. Potential range: 2.5–3.9 V vs. Li<sup>+</sup>/Li.</p>
Full article ">Figure 9
<p>Rate performance of Li/LiMn<sub>0.6</sub>Fe<sub>0.4</sub>PO<sub>4</sub> cells with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) and 5 wt.% of VC. Potential range: 2.4–4.2 V vs. Li<sup>+</sup>/Li. (<b>a</b>) Discharge capacities in mAh·g<sup>−1</sup>. (<b>b</b>) Coulombic efficiency in percentage (%).</p>
Full article ">Figure 10
<p>Rate performance of Li/LiMn<sub>0.6</sub>Fe<sub>0.4</sub>PO<sub>4</sub> (LMFP2) cells with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) and 5 wt.% of VC; and with 1.0 mol·kg<sup>−1</sup> LiPF<sub>6</sub> in EC:DMC. Potential range: 2.4–4.2 V vs. Li<sup>+</sup>/Li. (<b>a</b>) Discharge capacities in mAh·g<sup>−1</sup>. (<b>b</b>) Coulombic efficiency in percentage (%).</p>
Full article ">Figure 11
<p>Cycling stability and coulombic efficiency at C/10 of Li/LiFePO<sub>4</sub> (LFPB) cells with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) and 5 wt.% of VC; and with 1.0 mol·kg<sup>−1</sup> LiPF<sub>6</sub> in EC:DMC. Potential range: 2.5–3.9 V vs. Li<sup>+</sup>/Li. The first formation cycle was performed at C/25.</p>
Full article ">Figure 12
<p>Cycling stability and coulombic efficiency at C/10 of Li/LiMn<sub>0.6</sub>Fe<sub>0.4</sub>PO<sub>4</sub> (LMFP2) cells with 0.8 mol·kg<sup>−1</sup> LiPCP in EC:DMC (30:70 wt.%) and 5 wt.% of VC; and with 1.0 mol·kg<sup>−1</sup> LiPF<sub>6</sub> in EC:DMC. Potential range: 2.4–4.2 V vs. Li<sup>+</sup>/Li. The first formation cycle was performed at C/25.</p>
Full article ">
14 pages, 3066 KiB  
Article
Coffee Biomass-Based Carbon Material for the Electrochemical Determination of Antidepressant in Synthetic Urine
by Francisco Contini Barreto, Naelle Kita Mounienguet, Erika Yukie Ito, Quan He and Ivana Cesarino
Chemosensors 2024, 12(10), 205; https://doi.org/10.3390/chemosensors12100205 - 3 Oct 2024
Viewed by 645
Abstract
Escitalopram (ESC) is commonly prescribed as an antidepressant to enhance serotonin levels in the brain, effectively addressing conditions such as depression and anxiety. The COVID-19 pandemic, along with ongoing mental health crises, has exacerbated the prevalence of these disorders, largely due to factors [...] Read more.
Escitalopram (ESC) is commonly prescribed as an antidepressant to enhance serotonin levels in the brain, effectively addressing conditions such as depression and anxiety. The COVID-19 pandemic, along with ongoing mental health crises, has exacerbated the prevalence of these disorders, largely due to factors such as social isolation, fear of the virus, and financial difficulties. This study presents the enhancement of a glassy carbon electrode (GC) through the incorporation of hydrochar (HDC) derived from spent coffee grounds and copper nanoparticles (CuNPs) for the detection of ESC in synthetic urine. Characterization of the nanocomposite was conducted using scanning electron microscopy (SEM), energy-dispersive spectroscopy (EDS), and cyclic voltammetry (CV). The analytical parameters were systematically optimized, and a sensing platform was utilized for the quantification of ESC via square-wave voltammetry (SWV). The established linear range was found to be between 1.0 µmol L−1 and 50.0 µmol L−1, with a limit of detection (LOD) of 0.23 µmol L−1. Finally, an electrochemical sensor was employed to measure ESC levels in synthetic urine, yielding recovery rates ranging from 91.7% to 94.3%. Consequently, the HDC-CuNPs composite emerged as a promising, sustainable, and cost-effective alternative for electroanalytical applications. Full article
Show Figures

Figure 1

Figure 1
<p>Scanning electron microscopy (SEM) images depict (<b>A</b>) HDC; (<b>B</b>) HDC enhanced with CuNPs (HDC-CuNPs), accompanied by an inset showing the energy-dispersive spectroscopy (EDS) spectrum; and (<b>C</b>) the measurements of the diameters of the copper nanoparticles.</p>
Full article ">Figure 2
<p>The electrochemical characterization was conducted using CV at a scan rate of 50 mV s<sup>−1</sup> for the GC, GC/HDC, and GC/HDC-CuNPs electrodes in a PBS solution with a concentration of 0.2 mol L<sup>−1</sup> and a pH of 7.0.</p>
Full article ">Figure 3
<p>Different working electrodes were evaluated using cyclic voltammetry (CV) at a scan rate of 50 mV s<sup>−1</sup> in a phosphate-buffered saline (PBS) solution with a concentration of 0.2 mol L<sup>−1</sup> and a pH of 7.4, which contained 5.0 × 10<sup>−3</sup> mol L<sup>−1</sup> of potassium ferricyanide/ferrocyanide.</p>
Full article ">Figure 4
<p>Comparison among the SWV voltammograms of GC, GC/HDC and GC/HDC-CuNPs in 0.2 μmol L<sup>−1</sup> PBS pH 7.0 in the oxidation of 50 μmol L<sup>−1</sup> of ESC.</p>
Full article ">Figure 5
<p>The cyclic voltammograms obtained with 100.0 μmol L<sup>−1</sup> of ESC (indicated by the blue line) is compared to the voltammogram recorded without ESC (represented by the dotted line) in a 0.2 mol L<sup>−1</sup> PBS at a pH of 7.0, utilizing a scan rate of 50 mV s<sup>−1</sup> (inset: illustrates the oxidation mechanism of the molecule).</p>
Full article ">Figure 6
<p>Optimization experiments conducted to enhance the analytical signal for the oxidation of ESC. The following parameters were studied: (<b>A</b>) the optimization of the copper proportion relative to the mass of HDC; (<b>B</b>) the optimization of frequency; (<b>C</b>) the optimization of amplitude modulation; (<b>D</b>) the optimization of step potential; and (<b>E</b>) the optimal pH for the study of ESC utilizing the GC/HDC-CuNPs electrode.</p>
Full article ">Figure 7
<p>(<b>A</b>) Depicts the results of square-wave voltammetry (SWV) performed in a 0.2 mol L<sup>−1</sup> PBS solution at a pH of 7.0, with ESC concentrations ranging from 1.0 to 50.0 μmol L<sup>−1</sup>. (<b>B</b>) Demonstrates the linear relationship observed between anodic peak currents and ESC concentration.</p>
Full article ">Figure 8
<p>(<b>A</b>) SWV was executed in PBS at a concentration of 0.2 mol L<sup>−1</sup> and a pH of 7.0, incorporating a sample of synthetic urine with a final concentration of 3.00 μmol L<sup>−1</sup> of ESC (depicted by the red line) alongside three additions of a known concentration of the standard analyte (1.00 μmol L<sup>−1</sup> for each addition). (<b>B</b>) The graph illustrates the linear correlation between the anodic peak currents and the concentrations of ESC.</p>
Full article ">
15 pages, 6864 KiB  
Article
Advanced Electrochemical Monitoring of Carbendazim Fungicide in Foods Using Interfacial Superassembly of NRPC/NiMn Frameworks
by Shakila Parveen Asrafali, Thirukumaran Periyasamy, Seong Cheol Kim and Jaewoong Lee
Biosensors 2024, 14(10), 474; https://doi.org/10.3390/bios14100474 - 2 Oct 2024
Viewed by 601
Abstract
A simple, sensitive and reliable sensing system based on nitrogen-rich porous carbon (NRPC) and transition metals, NRPC/Ni, NRPC/Mn and NRPC/NiMn was developed and successfully applied as electrode materials for the quantitative determination of carbendazim (CBZ). The synergistic effect of NRPC and bimetals with [...] Read more.
A simple, sensitive and reliable sensing system based on nitrogen-rich porous carbon (NRPC) and transition metals, NRPC/Ni, NRPC/Mn and NRPC/NiMn was developed and successfully applied as electrode materials for the quantitative determination of carbendazim (CBZ). The synergistic effect of NRPC and bimetals with acceptable pore structure together with flower-like morphology resulted in producing a highly conductive and interconnected network in NRPC/NiMn@GCE, which significantly enhanced the detection performance of CBZ. The electrochemical behavior investigated by cyclic voltammetry (CV) showed improved CBZ detection for NRPC/NiMn, due to the controlled adsorption/diffusion process of CBZ by the NRPC/NiMn@GCE electrode. The influences of various factors such as pH, NRPC/NiMn concentration, CBZ concentration and scan rate were studied. Under optimal conditions, 0.1 M phosphate-buffered saline (PBS) with a pH of 7.0 containing 30 µg/mL NRPC/NiMn, a favourable linear range detection of CBZ from 5 to 50 µM was obtained. Moreover, a chronoamperometric analysis showed excellent repeatability, reproducibility and anti-interfering ability of the fabricated NRPC/NiMn@GCE sensor. Furthermore, the sensor showed satisfactory results for CBZ detection in real samples with acceptable recoveries of 96.40–104.98% and low RSD values of 0.25–3.45%. Full article
(This article belongs to the Special Issue Electrochemical Biosensing Platforms for Food, Drug and Health Safety)
Show Figures

Figure 1

Figure 1
<p>Synthesis of NRPC, NRPC/Mn, NRPC/Ni and NRPC/NiMn.</p>
Full article ">Figure 2
<p>XPS spectrum of NRPC/NiMn showing the (<b>a</b>) survey spectrum and (<b>b</b>–<b>f</b>) deconvoluted spectrum for each element.</p>
Full article ">Figure 3
<p>SEM images of NRPC, NRPC/Mn, NRPC/Ni and NRPC/NiMn at different magnifications.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) TEM images of NRPC/NiMn, (<b>d</b>) SAED pattern, (<b>e</b>–<b>j</b>) elemental mapping and (<b>k</b>) EDX spectrum.</p>
Full article ">Figure 5
<p>Electrochemical studies showing (<b>a</b>,<b>b</b>) CV, (<b>c</b>) EIS, (<b>d</b>,<b>e</b>) effect of pH, (<b>f</b>,<b>g</b>) effect of material concentration and (<b>h</b>,<b>i</b>) effect of analyte concentration of the prepared materials.</p>
Full article ">Figure 6
<p>Electrochemical studies showing (<b>a</b>,<b>b</b>) effect of scan rate, (<b>c</b>) mechanism of carbendazim detection, (<b>d</b>,<b>e</b>) repeatability, (<b>f</b>) reproducibility, (<b>g</b>) stability and (<b>h</b>,<b>i</b>) anti-interfering ability of NRPC/NiMn.</p>
Full article ">Figure 7
<p>Real-time analysis of NRPC/NiMn sensor: (<b>a</b>) Apple, (<b>b</b>) Carrot, (<b>c</b>) Grapes, (<b>d</b>) Blueberry, (<b>e</b>) Broccoli and (<b>f</b>) Tap water.</p>
Full article ">
Back to TopTop