[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,322)

Search Parameters:
Keywords = capacitive sensors

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 8226 KiB  
Article
Design of a Capacitive Tactile Sensor Array System for Human–Computer Interaction
by Fei Fei, Zhenkun Jia, Changcheng Wu, Xiong Lu and Zhi Li
Sensors 2024, 24(20), 6629; https://doi.org/10.3390/s24206629 - 14 Oct 2024
Viewed by 298
Abstract
This paper introduces a novel capacitive sensor array designed for tactile perception applications. Utilizing an all-in-one inkjet deposition printing process, the sensor array exhibited exceptional flexibility and accuracy. With a resolution of up to 32.7 dpi, the sensor array was capable of capturing [...] Read more.
This paper introduces a novel capacitive sensor array designed for tactile perception applications. Utilizing an all-in-one inkjet deposition printing process, the sensor array exhibited exceptional flexibility and accuracy. With a resolution of up to 32.7 dpi, the sensor array was capable of capturing the fine details of touch inputs, making it suitable for applications requiring high spatial resolution. The design incorporates two multiplexers to achieve a scanning rate of 100 Hz, ensuring the rapid and responsive data acquisition that is essential for real-time feedback in interactive applications, such as gesture recognition and haptic interfaces. To evaluate the performance of the capacitive sensor array, an experiment that involved handwritten number recognition was conducted. The results demonstrated that the sensor accurately captured fingertip inputs with a high precision. When combined with an Auxiliary Classifier Generative Adversarial Network (ACGAN) algorithm, the sensor system achieved a recognition accuracy of 98% for various handwritten numbers from “0” to “9”. These results show the potential of the capacitive sensor array for advanced human–computer interaction applications. Full article
(This article belongs to the Section Sensors Development)
Show Figures

Figure 1

Figure 1
<p>Sensor fabrication process: (<b>a</b>) a polyester film as a printing substrate; (<b>b</b>) PVA coating onto the polyester film; (<b>c</b>) printing process of the row electrode; (<b>d</b>) printing process of the column electrode; (<b>e</b>) printing process of the interconnects; (<b>f</b>) soldering of electronic components, and (<b>g</b>) the final fabricated sensor device.</p>
Full article ">Figure 2
<p>Design of the sensor system. (<b>a</b>) A 16 × 16 capacitive sensor with a pixel resolution of 32.7 dpi. (<b>b</b>) A 0.4 mm × 0.4 mm diamond sensing element and interconnects of 0.1 mm.</p>
Full article ">Figure 3
<p>Demonstration of the capacitive tactical sensor system. (<b>a</b>) The hardware system includes the capacitive sensor, the Arduino controller, the capacitance measurement module, and two multiplexers. (<b>b</b>) High-resolution micro-capacitive array. (<b>c</b>) Connection of the hardware system.</p>
Full article ">Figure 4
<p>Comparison of abnormal and normal capacitance values.</p>
Full article ">Figure 5
<p>(<b>a</b>) The dielectric between the emitter and the receiver is the photopolymer and air without the finger touching. (<b>b</b>) The dielectric between the emitter and the receiver is the photopolymer, air, and the finger. (<b>c</b>) Change in the sensor capacitance before and after finger touching.</p>
Full article ">Figure 6
<p>Sequence of capacitance values corresponding to the number “0” trajectory.</p>
Full article ">Figure 7
<p>Sliding motion with the index finger and the visualized trajectory of the number “0”.</p>
Full article ">Figure 8
<p>Visualized trajectories from numbers “1” to “9”.</p>
Full article ">Figure 9
<p>The training processes of the (<b>a</b>) GAN, (<b>b</b>) CGAN, and (<b>c</b>) ACGAN.</p>
Full article ">Figure 10
<p>The generator model of the ACGAN.</p>
Full article ">Figure 11
<p>The discriminator model of the ACGAN.</p>
Full article ">Figure 12
<p>The trajectory heatmaps of the numbers “0–9”: (<b>a</b>) trajectory heatmaps obtained by drawing numbers on the capacitive sensor using a finger; (<b>b</b>) trajectory heatmaps generated using a GAN model’s generator.</p>
Full article ">Figure 13
<p>The (<b>a</b>) loss and (<b>b</b>) auxiliary loss of the model, as well as the confusion matrix for the discriminator model on the (<b>c</b>) validation set and (<b>d</b>) fake image set.</p>
Full article ">
11 pages, 5162 KiB  
Article
Thiol-SAM Concentration Effect on the Performance of Interdigitated Electrode-Based Redox-Free Biosensors
by Abdulaziz K. Assaifan
Micromachines 2024, 15(10), 1254; https://doi.org/10.3390/mi15101254 - 12 Oct 2024
Viewed by 379
Abstract
Despite the direct, redox-free and simple detection non-faradaic impedimetric biosensors offer, considerable optimizations are required to enhance their performance for the detection of various biomarkers. Non-faradaic EIS sensors’ performance depends on the interfacial capacitance between a polarized biosensor surface and the tested sample [...] Read more.
Despite the direct, redox-free and simple detection non-faradaic impedimetric biosensors offer, considerable optimizations are required to enhance their performance for the detection of various biomarkers. Non-faradaic EIS sensors’ performance depends on the interfacial capacitance between a polarized biosensor surface and the tested sample solution. Careful engineering and design of the interfacial capacitance is encouraged to magnify the redout signal upon bioreceptor–antigen interactions. One of the methods to achieve this goal is by optimizing the self-assembled monolayer concentration, which has not been reported for non-faradaic impedimetric sensors. Here, the impact of alkanethiolate (cysteamine) concentration on the performance of gold (Au) interdigitated electrode (Au-IDE) biosensors is reported. Six sets of biosensors were prepared, each with a different cysteamine concentration: 100 nM, 1 μM, 10 μM, 100 μM, 1 mM, and 10 mM. The biosensors were prepared for the direct detection of LDL cholesterol by attaching LDL antibodies on top of the cysteamine via a glutaraldehyde cross-linker. As the concentration of cysteamine increased from 100 nM to 100 μM, the sensitivity of the biosensor increased from 6.7 to 16.2 nF/ln (ng/mL). As the cysteamine concentration increased from 100 μM to 10 mM, the sensitivity deteriorated. The limit of detection (LoD) of the biosensor improved as the cysteamine increased from 100 nM to 100 μM (i.e., 400 ng/mL to 59 pg/mL). However, the LoD started to increase to 67 pg/mL and 16 ng/mL for 1 mM and 10 mM cysteamine concentrations, respectively. This shows that the cysteamine concentration has a detrimental effect on redox-free biosensors. The cysteamine layer has to be as thin as possible and uniformly cover the electrode surfaces to maximize positive readout signals and reduce negative signals, significantly improving both sensitivity and LoD. Full article
Show Figures

Figure 1

Figure 1
<p>Non-faradaic biosensor structure.</p>
Full article ">Figure 2
<p>(<b>a</b>) Au-IDEs on white plastic substrate. (<b>b</b>) Biofunctionalization steps.</p>
Full article ">Figure 3
<p>(<b>a</b>) Nyquist plot for (I) bare Au-IDEs and Au-IDEs functionalized with (II) 100 nM, (III) 1 µM, (IV) 10 µM, (V) 100 µM, (VI) 1 mM and (VII) 10 mM of cysteamine. (<b>b</b>) Randle circuit for bare Au-IDEs. (<b>c</b>) Randle circuit for cysteamine-functionalized Au-IDEs.</p>
Full article ">Figure 4
<p>Typical Bode plots for Au-IDE biosensors functionalized with (<b>a</b>) 100 nM, (<b>b</b>) 1 µM, (<b>c</b>) 10 µM, (<b>d</b>) 100 µM, (<b>e</b>) 1 mM and (<b>f</b>) 10 mM of cysteamine.</p>
Full article ">Figure 5
<p>Calibration curves for Au-IDE biosensors functionalized with (<b>a</b>) 100 nM, (<b>b</b>) 1 µM, (<b>c</b>) 10 µM, (<b>d</b>) 100 µM, (<b>e</b>) 1 mM and (<b>f</b>) 10 mM of cysteamine.</p>
Full article ">Figure 6
<p>Effect of cysteamine concentration on (<b>a</b>) sensitivity and (<b>b</b>) LoD of Au-IDE biosensors.</p>
Full article ">Figure 7
<p>Selectivity behavior of the Au-IDE biosensor.</p>
Full article ">
17 pages, 4461 KiB  
Article
A Novel Wearable Sensor for Measuring Respiration Continuously and in Real Time
by Amjad Ali, Yang Wei, Yomna Elsaboni, Jack Tyson, Harry Akerman, Alexander I. R. Jackson, Rod Lane, Daniel Spencer and Neil M. White
Sensors 2024, 24(20), 6513; https://doi.org/10.3390/s24206513 - 10 Oct 2024
Viewed by 409
Abstract
In this work, a flexible textile-based capacitive respiratory sensor, based on a capacitive sensor structure, that does not require direct skin contact is designed, optimised, and evaluated using both computational modelling and empirical measurements. In the computational study, the geometry of the sensor [...] Read more.
In this work, a flexible textile-based capacitive respiratory sensor, based on a capacitive sensor structure, that does not require direct skin contact is designed, optimised, and evaluated using both computational modelling and empirical measurements. In the computational study, the geometry of the sensor was examined. This analysis involved observing the capacitance and frequency variations using a cylindrical model that mimicked the human body. Four designs were selected which were then manufactured by screen printing multiple functional layers on top of a polyester/cotton fabric. The printed sensors were characterised to detect the performance against phantoms and impacts from artefacts, normally present whilst wearing the device. A sensor that has an electrode ratio of 1:3:1 (sensor, reflector, and ground) was shown to be the most sensitive design, as it exhibits the highest sensitivity of 6.2% frequency change when exposed to phantoms. To ensure the replicability of the sensors, several batches of identical sensors were developed and tested using the same physical parameters, which resulted in the same percentage frequency change. The sensor was further tested on volunteers, showing that the sensor measures respiration with 98.68% accuracy compared to manual breath counting. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The layout of the designed sensor. (<b>b</b>) The structure of the four sensor designs and ratio combinations of the sensor, reflector, and ground electrode.</p>
Full article ">Figure 2
<p>Diagram of simulation methodology conducted to obtain electric field distribution around each sensor design.</p>
Full article ">Figure 3
<p>(<b>a</b>) The computed vertical electric field is (1) below the ground electrode, (2) between the ground and reflector electrode (has a high electric field which reaches up to 75,000 v/m due to a thin dielectric layer), (3) between the sensor and phantom (design 1: vertical electrical field peak is 50 v/m higher than the rest of the design), and (4) within the phantom (the phantom is a glass cylinder of 80 mm diameter set to a dielectric constant of 5 with a glass wall thickness of 2.5 mm and filled with water having dielectric constant of 80). (<b>b</b>) The computed horizontal electric field distribution across the four sensor designs.</p>
Full article ">Figure 4
<p>The sensor capacitance was recorded when the phantom was located at z = 1, 5, 10, 15, 20, and 25 mm distances. The electric field distribution of the final sensor designs across the perpendicular plane and the parallel plane. The fringing field effect is present at the corners between the sensor and ground electrodes.</p>
Full article ">Figure 5
<p>Simulation capacitance results were obtained with 3 phantoms: muscle, acetone, and water. (<b>a</b>) shows the capacitance of each design between the sensor and ground electrodes, which is referred to as Csg and given a blue colour along with its scale on the left side of each graph. The capacitance between the sensor and the object is referred to as Cso, which is given an orange colour along with its orange scales on the right side of each graph. (<b>b</b>) shows each design’s total capacitance toward the phantom.</p>
Full article ">Figure 6
<p>(<b>a</b>) The final printed sensor in four designs. (<b>b</b>) SEM images show the different layers and their corresponding average thickness (showing the thickness of each layer).</p>
Full article ">Figure 7
<p>(<b>a</b>) Empirical setup to evaluate the sensor response toward phantoms. (<b>b</b>) The equivalent circuit model of the respiratory rate sensor and interfacing circuit.</p>
Full article ">Figure 8
<p>Frequency variation measurements obtained empirically when testing 3 different phantoms.</p>
Full article ">Figure 9
<p>(<b>a</b>) Design 2’s identical replicas (Sample 1, 2, and 3) from three different batches of screen printing, and (<b>b</b>) their consistently similar %f-c toward mowing away water phantom.</p>
Full article ">Figure 10
<p>(<b>a</b>) Humidity variations ranging from 40% to 80% RH at 24 °C cause an average standard deviation of 0.19 in the design 2 response. (<b>b</b>) Temperature variations ranging from 18 °C to 35 °C at 60% relative humidity (RH) impact the response of design 2, resulting in an average standard deviation of 0.34.</p>
Full article ">Figure 11
<p>(<b>a</b>) The results of the flexing durability test conducted empirically on the four designs wrapped on cylinders of multiple diameters. (<b>b</b>) The graph shows the four designs’ responses to increasing pressure. (<b>c</b>) The graph shows the impact of rubbing on sensor response.</p>
Full article ">Figure 12
<p>(<b>a</b>) The left image shows the sensor that is attached to the lower part of the chest of the test subject in a sedentary position. The middle image shows how the test subject’s torso was divided into nine positions. (<b>b</b>) The sensor was attached to each position and measured the corresponding breathing rate for one minute. Precise frequency peaks corresponding to the breathing rate can be seen when the sensor is attached at positions 4, 7, 8, and 9. (<b>c</b>) shows the sensor’s response for a random breathing rate of 11 and 22 in one minute. (<b>d</b>) The sensor is attached at position 8, while the test subject is in standing posture and took 11 breaths in one minute.</p>
Full article ">
14 pages, 6789 KiB  
Article
Real-Time Nonlinear Image Reconstruction in Electrical Capacitance Tomography Using the Generative Adversarial Network
by Damian Wanta, Mikhail Ivanenko, Waldemar T. Smolik, Przemysław Wróblewski and Mateusz Midura
Information 2024, 15(10), 617; https://doi.org/10.3390/info15100617 - 9 Oct 2024
Viewed by 281
Abstract
This study investigated the potential of the generative adversarial neural network (cGAN) image reconstruction in industrial electrical capacitance tomography. The image reconstruction quality was examined using image patterns typical for a two-phase flow. The training dataset was prepared by generating images of random [...] Read more.
This study investigated the potential of the generative adversarial neural network (cGAN) image reconstruction in industrial electrical capacitance tomography. The image reconstruction quality was examined using image patterns typical for a two-phase flow. The training dataset was prepared by generating images of random test objects and simulating the corresponding capacitance measurements. Numerical simulations were performed using the ECTsim toolkit for MATLAB. A cylindrical sixteen-electrode ECT sensor was used in the experiments. Real measurements were obtained using the EVT4 data acquisition system. The reconstructed images were evaluated using selected image quality metrics. The results obtained using cGAN are better than those obtained using the Landweber iteration and simplified Levenberg–Marquardt algorithm. The suggested method offers a promising solution for a fast reconstruction algorithm suitable for real-time monitoring and the control of a two-phase flow using ECT. Full article
(This article belongs to the Special Issue Deep Learning for Image, Video and Signal Processing)
Show Figures

Figure 1

Figure 1
<p>The classes of training datasets. (<b>a</b>) Typical flow patterns: annular flow, stratified flow, and slug flow (one and two rods). (<b>b</b>) Images containing random circular objects. Two relative permittivity values, 1 and 3, were used. For each map of permittivity distribution, the corresponding capacitance measurements are shown in the form of a 2D 16 × 15 pixels color map.</p>
Full article ">Figure 2
<p>(<b>a</b>) ECT sensor filled with PP pellet, and (<b>b</b>) sketch of the sensors.</p>
Full article ">Figure 3
<p>The signal-to-noise ratio of capacitance measurements obtained using the prepared sensor and EVT4 data acquisition system: ‘min’—measurement with empty sensor; ‘max’—measurement conducted with sensor filled with PP pellet. The graph shows data for the first electrode in the pair.</p>
Full article ">Figure 4
<p>The architecture of the generator network in the cGAN model featuring a U-Net-style structure.</p>
Full article ">Figure 5
<p>Neural network training procedure.</p>
Full article ">Figure 6
<p>Neural network training on 50 epochs: (<b>a</b>) discriminator loss, (<b>b</b>) generator loss, and (<b>c</b>) relative image error.</p>
Full article ">Figure 7
<p>The modeled true distribution of relative permittivity and images were reconstructed using the Landweber method, sLM algorithm (λ = 5 × 10<sup>−3</sup>), and cGAN neural network. Based on examples from the test set, with added noise (set to achieve an SNR of 30 dB for the opposing electrodes in the measurement).</p>
Full article ">Figure 7 Cont.
<p>The modeled true distribution of relative permittivity and images were reconstructed using the Landweber method, sLM algorithm (λ = 5 × 10<sup>−3</sup>), and cGAN neural network. Based on examples from the test set, with added noise (set to achieve an SNR of 30 dB for the opposing electrodes in the measurement).</p>
Full article ">Figure 8
<p>Histograms of mean square error (<b>a</b>,<b>b</b>) and correlation distribution (<b>c</b>,<b>d</b>) for the elements of the testing dataset with flow patterns (<b>a</b>,<b>c</b>) and with random circles (<b>b</b>,<b>d</b>).</p>
Full article ">Figure 9
<p>Histogram of the computation time required for each reconstruction algorithm.</p>
Full article ">Figure 10
<p>Four test objects mimicking the two-phase flow patterns (annular and stratified flows, slug, and circles) and the results of the image reconstructions. From left to right: view of test objects inside the sensor, numerical representations of the test objects, normalized sinograms of measured capacitances, and images reconstructed using the Landweber algorithm, the sLM algorithm, and cGAN.</p>
Full article ">Figure 10 Cont.
<p>Four test objects mimicking the two-phase flow patterns (annular and stratified flows, slug, and circles) and the results of the image reconstructions. From left to right: view of test objects inside the sensor, numerical representations of the test objects, normalized sinograms of measured capacitances, and images reconstructed using the Landweber algorithm, the sLM algorithm, and cGAN.</p>
Full article ">
12 pages, 9025 KiB  
Article
Plantar Load System Analysis Using FSR Sensors and Interpolation Methods
by Gabriel Trujillo-Hernández, Dayanna Ortiz-Villaseñor, Julio C. Rodríguez-Quiñonez, Luis Roberto Ramírez-Hernández, Fabian N. Murrieta-Rico, Abelardo Mercado-Herrera, María E. Raygoza-Limón and Jesús Heriberto Orduño-Osuna
Metrology 2024, 4(4), 566-577; https://doi.org/10.3390/metrology4040035 - 9 Oct 2024
Viewed by 369
Abstract
The foot is considered a wonder of biological engineering due to its structure, formed by bones, ligaments, and tendons that collaborate to ensure stability and mobility. A key area often examined by medical professionals in patients with diabetic feet is the plantar surface, [...] Read more.
The foot is considered a wonder of biological engineering due to its structure, formed by bones, ligaments, and tendons that collaborate to ensure stability and mobility. A key area often examined by medical professionals in patients with diabetic feet is the plantar surface, due to the risk of ulcer development. If left untreated, these ulcers can lead to severe complications, including amputation of the toe, foot, or even the limb. Interpolation methods are used to find areas with overloads in a system of sensor maps that are based on capacitive, load cells, or force-sensitive resistors (FSRs). This manuscript presents the assessment of linear, nearest neighbors, and bicubic methods in comparison with ground truth to calculate the root mean square error (RMSE) in two assessments using a dataset of eight healthy subjects, four men and four women, with an average age of 25 years, height of 1.63 m, and weight of 72 kg with shoe sizes from 7.3 USA using FSR map with 48 sensors. Additionally, this paper describes the conditioning circuit development to implement a plantar surface system that enables interpolating loads on the plantar surface. The proposed system’s results show that the first assessment indicates an RMSE of 0.089, 0.126, and 0.089 for linear, nearest neighbor, and bicubic methods, while the second assessment shows a mean RMSE for linear, nearest neighbor, and bicubic methods of 0.114, 0.159, and 0.112. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Insoles with 48 FSR sensors. (<b>b</b>) Foam and plastic cover for the insole.</p>
Full article ">Figure 2
<p>FSR sensor coordinates in the plantar load insole.</p>
Full article ">Figure 3
<p>This conditioning circuit is composed of a microcontroller, four 40 k<math display="inline"><semantics> <mi>Ω</mi> </semantics></math> resistances, one LM324 Quad OP-AMP.</p>
Full article ">Figure 4
<p>Characteristic curvature kilograms versus conductance. The FSR sensors’ load capacity leads to the selection of <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mn>2</mn> </msub> </mrow> </semantics></math> which define the linear equation.</p>
Full article ">Figure 5
<p>(<b>a</b>) First evaluation with internal ground truth. (<b>b</b>) Second evaluation with external ground truth. (<b>c</b>) External FSR sensor with NAN values with interpolation methods.</p>
Full article ">Figure 6
<p>Experimentation was performed on eight healthy subjects.</p>
Full article ">Figure 7
<p>(<b>a</b>) Obtained RMSE values for the first assessment. (<b>b</b>) Obtained RMSE values for the second assessment.</p>
Full article ">Figure 8
<p>(<b>a</b>) Internal ground truth: (<b>a</b>) linear, (<b>b</b>) nearest neighbor, (<b>c</b>) bicubic.</p>
Full article ">Figure 9
<p>(<b>a</b>) External ground truth: (<b>a</b>) linear, (<b>b</b>) nearest neighbor, (<b>c</b>) bicubic.</p>
Full article ">
23 pages, 10725 KiB  
Review
Hardware Testing Methodologies for Wide Bandgap High-Power Converters
by Zibo Chen, Zhicheng Guo, Chen Chen and Alex Q. Huang
Electronics 2024, 13(19), 3918; https://doi.org/10.3390/electronics13193918 - 3 Oct 2024
Viewed by 523
Abstract
Wide bandgap (WBG) power semiconductor devices are increasingly replacing silicon IGBTs in high-power and high-voltage power electronics applications. However, there is a significant gap in the literature regarding efficient testing methodologies for high-power and high-voltage converters under constrained laboratory resources. This paper addresses [...] Read more.
Wide bandgap (WBG) power semiconductor devices are increasingly replacing silicon IGBTs in high-power and high-voltage power electronics applications. However, there is a significant gap in the literature regarding efficient testing methodologies for high-power and high-voltage converters under constrained laboratory resources. This paper addresses this gap by presenting comprehensive, hardware-focused testing methodologies for high-power and high-voltage WBG power semiconductor-based converter bring-up before the control validation phase steps in. The proposed methods enable thorough evaluation and validation of converter hardware, including device switching characteristics, driving circuit functionality, thermal management performance, insulation integrity, and sustained operation at full power. We utilized the double pulse test (DPT) to characterize switching performance in a two-level phase leg configuration, extract circuit parasitics, and validate magnetic components. The DPT was further applied to optimize gate driving circuits, validate overcurrent protection mechanisms, and measure device on-resistance. Additionally, a multicycle test was introduced to rapidly assess steady-state converter performance and estimate efficiency. Recognizing the critical role of thermal management in high-power converters, our methodologies extend to the experimental extraction of key thermal parameters—such as junction-to-ambient thermal resistance and thermal capacitance—via a heat loss injection method. A correlation method between temperature sensor measurements and junction temperature is presented to enhance the accuracy of device temperature monitoring during tests. To ensure reliability and safety, dielectric withstand tests and partial discharge measurements were conducted at both component and converter levels under conventional 60 Hz sinusoidal and high-frequency PWM waveforms. Finally, we highlight the importance of testing converters under full voltage, current, and thermal conditions through power circulating tests with minimal power consumption, applicable to both non-isolated and isolated high-power converters. Practical examples are provided to demonstrate the effectiveness and applicability of these hardware testing methodologies. Full article
(This article belongs to the Special Issue Advances in Power Converter Design, Control and Applications)
Show Figures

Figure 1

Figure 1
<p>Examples of Power Electronics Power Stages (PEPSs). (<b>a</b>) 1200 V/4.3 mΩ SiC six-phase power stage. (<b>b</b>) 2400 V/7 mΩ SiC full-bridge power stage. (<b>c</b>) 1200 V/7.3 mΩ SiC full-bridge plus IGBT full-bridge power stage. (<b>d</b>) 900 V/2.5 mΩ three level power stage.</p>
Full article ">Figure 2
<p>Summary of the design evaluation test.</p>
Full article ">Figure 3
<p>Test setup for double pulse test (DPT). (<b>a</b>) Bottom side DPT. (<b>b</b>) Top side DPT.</p>
Full article ">Figure 4
<p>Typical waveform for a double pulse test.</p>
Full article ">Figure 5
<p>Commutation loop inductance.</p>
Full article ">Figure 6
<p>Typical switching waveform for double pulse test. (<b>a</b>) Double pulse test with 800 V bus voltage. (<b>b</b>) Zoomed-in waveform for the first turn-on at 0 A. (<b>c</b>) Zoomed-in waveform for the second turn-on at 138 A. (<b>d</b>) Zoomed-in waveform for the second turn-off at 175 A.</p>
Full article ">Figure 7
<p>A gate driver integrating online voltage measurement and overcurrent protection circuit.</p>
Full article ">Figure 8
<p>Overcurrent protection validation in double pulse test.</p>
Full article ">Figure 9
<p>On-resistance extraction in double pulse test.</p>
Full article ">Figure 10
<p>Thermal network equivalent circuit models. (<b>a</b>) Cauer model. (<b>b</b>) Foster model.</p>
Full article ">Figure 11
<p>Loss injection for thermal management evaluation of a six-phase power stage. (<b>a</b>) Reverse conduction for loss injection. (<b>b</b>) Forward conduction for loss injection.</p>
Full article ">Figure 12
<p>Thermal camera image in a forward conduction loss heat injection test.</p>
Full article ">Figure 13
<p>Example for temperature calibration.</p>
Full article ">Figure 14
<p>V<sub>sd</sub> measurement at varies T<sub>j</sub>.</p>
Full article ">Figure 15
<p>A typical waveform for thermal impedance extraction.</p>
Full article ">Figure 16
<p>Correlation curve between thermocouple and junction temperature.</p>
Full article ">Figure 17
<p>Test setup for DC dielectric withstand test.</p>
Full article ">Figure 18
<p>Dielectric withstand test result. (<b>a</b>) Voltage increasing from 2450 V to 2520 V. (<b>b</b>) Voltage increasing from 2627 V to 2722 V.</p>
Full article ">Figure 19
<p>Circuit diagram of a 60 Hz sinusoidal waveform PD test platform. (<b>a</b>) Circuit diagram. (<b>b</b>) Hardware setup.</p>
Full article ">Figure 20
<p>PD test at 60 Hz sinusoidal waveform (orange: applied voltage; blue: PD charge).</p>
Full article ">Figure 21
<p>PD pattern diagram at AC 60 Hz peak-to-peak 7.6 kV.</p>
Full article ">Figure 22
<p>A high-frequency high-voltage PWM waveform PD test platform. (<b>a</b>) Circuit diagram. (<b>b</b>) Hardware setup.</p>
Full article ">Figure 23
<p>PD pattern diagram under 20 kHz 7 kV peak-to-peak PWM waveform.</p>
Full article ">Figure 24
<p>Inductive load three-phase circulating test.</p>
Full article ">Figure 25
<p>Configuration of the six-phase leg pump-back circulating test.</p>
Full article ">Figure 26
<p>Phasor diagram for pump-back circulating test.</p>
Full article ">Figure 27
<p>Switching states in the pump-back circulating test. (<b>a</b>) inductor current increases. (<b>b</b>) freewheeling through top devices. (<b>c</b>) inductor current decreases. (<b>d</b>) freewheeling through bottom devices. (<b>e</b>) inductor current increases in the reverse direction. (<b>f</b>) freewheeling through top devices. (<b>g</b>) inductor current decreases. (<b>h</b>) freewheeling through bottom devices.</p>
Full article ">Figure 27 Cont.
<p>Switching states in the pump-back circulating test. (<b>a</b>) inductor current increases. (<b>b</b>) freewheeling through top devices. (<b>c</b>) inductor current decreases. (<b>d</b>) freewheeling through bottom devices. (<b>e</b>) inductor current increases in the reverse direction. (<b>f</b>) freewheeling through top devices. (<b>g</b>) inductor current decreases. (<b>h</b>) freewheeling through bottom devices.</p>
Full article ">Figure 28
<p>Typical experimental waveform for a circulating test (three inductor current on the top; device drain-source voltage on the bottom).</p>
Full article ">Figure 29
<p>Pump-back circulating test for isolated converters. (<b>a</b>) Pump-back test for isolated DC-DC converter. (<b>b</b>) Pump-back test for isolated AC-AC converter.</p>
Full article ">Figure 29 Cont.
<p>Pump-back circulating test for isolated converters. (<b>a</b>) Pump-back test for isolated DC-DC converter. (<b>b</b>) Pump-back test for isolated AC-AC converter.</p>
Full article ">
25 pages, 7883 KiB  
Article
Estimation of Dry Matter Yield in Mediterranean Pastures: Comparative Study between Rising Plate Meter and Grassmaster II Probe
by João Serrano, Júlio Franco, Shakib Shahidian and Francisco J. Moral
Agriculture 2024, 14(10), 1737; https://doi.org/10.3390/agriculture14101737 - 2 Oct 2024
Viewed by 396
Abstract
This study evaluates two expedient electronic sensors, a rising plate meter (RPM) and a “Grassmaster II” capacitance probe (GMII), to estimate pasture dry matter (DM, in kg ha−1). The sampling process consisted of sensor measurements, followed by pasture collection and a [...] Read more.
This study evaluates two expedient electronic sensors, a rising plate meter (RPM) and a “Grassmaster II” capacitance probe (GMII), to estimate pasture dry matter (DM, in kg ha−1). The sampling process consisted of sensor measurements, followed by pasture collection and a laboratory reference analysis. In this comparative study, carried out throughout the 2023/2024 pasture growing season, a total of 288 pasture samples were collected in two phases (calibration and validation). The calibration phase (n = 144) consisted of measurements on three dates (6 December 2023, 29 February and 10 May 2024) in 48 georeferenced sampling areas of the experimental field “Eco-SPAA” (“MG” field), located at Mitra farm (Évora, Portugal). This pasture is a permanent mixture of various botanical species (grasses, legumes, and others) grazed by sheep, and is representative of biodiverse dryland pastures. The validation phase (n = 144) was carried out between December 2023 and April 2024 in 18 field tests (each with eight pasture samples), in three types of representative pastures: the same mixture for grazing (“MG” field), a commercial and annual mixture for cutting (mowing) and conservation (“MM” field), and legumes for grazing (“LG” field). The best estimation model for DM was obtained based on measurements carried out in February in the case of the GMII probe (R2 = 0.61) and December 2023 and February 2024 in the case of RPM (R2 = 0.76). The estimation decreased very significantly for both sensors based on measurements carried out in May (spring). The validation phase showed greater accuracy (less RMSE) in “MG” field tests (RMSE of 735.4 kg ha−1 with GMII and 512.3 kg ha−1 with the RPM). The results open perspectives for other works that would allow the testing, calibration, and validation of these electronic sensors in a wider range of pasture production conditions, in order to improve their accuracy as decision-making support tools in pasture management. Full article
(This article belongs to the Special Issue Innovations in Precision Farming for Sustainable Agriculture)
Show Figures

Figure 1

Figure 1
<p>A schematic diagram illustrating the chronological sequence of field tests carried out in this study.</p>
Full article ">Figure 2
<p>Location of experimental fields in Mitra farm and indication of number of field test codes.</p>
Full article ">Figure 3
<p>Detail of “Eco-SPAA” field in Mitra farm with location of 48 sampling areas of calibration phase and 6 sampling areas of validation phase and indication of predominant botanical species.</p>
Full article ">Figure 4
<p>Predominant pasture species of the experimental fields: “M<sub>G</sub>” (mixture of species for sheep grazing); “M<sub>M</sub>” (mixture of species for mowing and conservation); and “L<sub>G</sub>” (legumes for cattle grazing).</p>
Full article ">Figure 5
<p>Monthly mean temperature and rainfall of meteorologic station of Évora, between July 2023 and June 2024.</p>
Full article ">Figure 6
<p>Equipment used in this experimental work: (<b>a</b>) electronic graduated ruler; (<b>b</b>) Grassmaster II probe; and (<b>c</b>) rising plate meter.</p>
Full article ">Figure 7
<p>Pasture cutting in the field: (<b>a</b>) quadrats; (<b>b</b>) grass shears.</p>
Full article ">Figure 8
<p>Relationship between pasture height (H) and pasture compressed height (H<sub>RPM</sub>) in calibration field tests: in December (<b>a</b>), in February (<b>b</b>), in May (<b>c</b>), and on all dates (<b>d</b>).</p>
Full article ">Figure 9
<p>Relationship between pasture height (H) and dry matter (DM) in calibration field tests: in December (<b>a</b>), in February (<b>b</b>), in May (<b>c</b>), and on all dates (<b>d</b>).</p>
Full article ">Figure 10
<p>Relationship between pasture compressed height (H<sub>RPM</sub>) and dry matter (DM) in calibration field tests: in December (<b>a</b>), in February (<b>b</b>), in May (<b>c</b>), and on all dates (<b>d</b>).</p>
Full article ">Figure 11
<p>Relationship between corrected meter reading (CMR) and pasture dry matter (DM) in calibration field tests: in December (<b>a</b>), in February (<b>b</b>), in May (<b>c</b>), and on all dates (<b>d</b>).</p>
Full article ">Figure 11 Cont.
<p>Relationship between corrected meter reading (CMR) and pasture dry matter (DM) in calibration field tests: in December (<b>a</b>), in February (<b>b</b>), in May (<b>c</b>), and on all dates (<b>d</b>).</p>
Full article ">Figure 12
<p>Better model to estimate pasture dry matter (DM) based on pasture compressed height (H<sub>RPM</sub>) measured in calibration field tests of December and February (n = 96).</p>
Full article ">Figure 13
<p>Validation field tests: (<b>a</b>) relationship between pasture height (H) and pasture compressed height (H<sub>RPM</sub>); relationship between pasture dry matter (DM) and height (H), and pasture compressed height (H<sub>RPM</sub>) and corrected meter reading (CMR), (<b>b</b>), (<b>c</b>), and (<b>d</b>), respectively.</p>
Full article ">Figure 14
<p>Validation field tests for Grassmaster II probe (<b>a</b>) and for rising plate meter (RPM; <b>b</b>); indication of respective RMSE.</p>
Full article ">Figure 15
<p>Measurements with the rising plate meter (RPM) in the field of the grassland mixture for cutting in the validation phase.</p>
Full article ">Figure 16
<p>Sources of sensor error. Practical recommendations for future studies.</p>
Full article ">Figure 17
<p>Maps of pasture productivity (DM) measured (obtained from reference laboratory determinations) and estimated, based on Grassmaster II probe and RPM readings in February 2024.</p>
Full article ">
31 pages, 3793 KiB  
Article
A Circular Touch Mode Capacitive Rainfall Sensor: Analytical Solution and Numerical Design and Calibration
by Xiao-Ting He, Jun-Song Ran, Ji Wu, Fei-Yan Li and Jun-Yi Sun
Sensors 2024, 24(19), 6291; https://doi.org/10.3390/s24196291 - 28 Sep 2024
Viewed by 430
Abstract
A circular capacitive rainfall sensor can operate from non-touch mode to touch mode; that is, under the action of enough rainwater, its movable electrode plate can form a circular contact area with its fixed electrode plate. Therefore, the weight of rainwater is borne [...] Read more.
A circular capacitive rainfall sensor can operate from non-touch mode to touch mode; that is, under the action of enough rainwater, its movable electrode plate can form a circular contact area with its fixed electrode plate. Therefore, the weight of rainwater is borne by only its movable electrode plate in non-touch mode operation but by both its movable and fixed electrode plates in touch mode operation, and the total capacitance of its touch mode operation is much larger than that of its non-touch mode operation. Essential to its numerical design and calibration is the ability to predict the deflection shape of its moveable electrode plate to determine its total capacitance. This requires the analytical solution to the fluid–structure interaction problem of its movable electrode plate under rainwater. In our previous work, only the analytical solution for the fluid–structure interaction problem before its movable electrode plate touches its fixed electrode plate was obtained, and how to numerically design and calibrate a circular non-touch mode capacitive rainfall sensor was illustrated. In this paper, the analytical solution for the fluid–structure interaction problem after its movable electrode plate touches its fixed electrode plate is obtained, and how to numerically design and calibrate a circular touch mode capacitive rainfall sensor is illustrated for the first time. The numerical results show that the total capacitance and rainwater volume when the circular capacitive rainfall sensor operates in touch mode is indeed much larger than that when the same circular capacitive rainfall sensor operates in non-touch mode, and that the average increase in the maximum membrane stress per unit rainwater volume when the circular capacitive rainfall sensor operates in touch mode can be about 20 times smaller than that when the same circular capacitive rainfall sensor operates in non-touch mode. This is where the circular touch mode capacitive rainfall sensor excels. Full article
(This article belongs to the Special Issue Recent Advances in Low Cost Capacitive Sensors)
Show Figures

Figure 1

Figure 1
<p>Schematic of a rainfall measurement system using a circular capacitive sensor operating in touch mode: (<b>a</b>) the occasion when the movable electrode plate is just in contact with the insulator layer; and (<b>b</b>) the case after contact between the movable electrode plate and the insulator layer.</p>
Full article ">Figure 2
<p>Structural parameters of a circular touch mode capacitive rainfall sensor.</p>
Full article ">Figure 3
<p>Half cross-sectional view of a circular touch mode capacitive rainfall sensor.</p>
Full article ">Figure 4
<p>The geometric relationship between the micro radial straight line element <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>NM</mi> </mrow> <mo stretchy="true">¯</mo> </mover> </mrow> </semantics></math> and the micro meridional curve element <math display="inline"><semantics> <mrow> <mover> <mrow> <mrow> <msup> <mi mathvariant="normal">N</mi> <mo>′</mo> </msup> <msup> <mi mathvariant="normal">M</mi> <mo>′</mo> </msup> </mrow> </mrow> <mo>⏜</mo> </mover> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>The geometric relationship between the circumferential curve micro elements <math display="inline"><semantics> <mrow> <mover> <mrow> <mi>NM</mi> </mrow> <mo>⏜</mo> </mover> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mover> <mrow> <mrow> <mi mathvariant="normal">N</mi> <mo>′</mo> <mi mathvariant="normal">M</mi> <mo>′</mo> </mrow> </mrow> <mo>⏜</mo> </mover> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>The variation in the deflection <span class="html-italic">w</span> with the coordinate variable <span class="html-italic">r</span>, where the dashed line (“Solution 1”) refers to the calculated result by the analytical solution for the plate/membrane non-contact problem in Section 2 in [<a href="#B51-sensors-24-06291" class="html-bibr">51</a>], where the rainwater height <span class="html-italic">H</span> takes only 8 mm; and the solid line (“Solution 2”) refers to the calculated results by the analytical solution for the plate/membrane contact problem in <a href="#sec2-sensors-24-06291" class="html-sec">Section 2</a> in this paper, where the rainwater height <span class="html-italic">H</span> takes 8 mm, 10 mm, 20 mm, 40 mm, 80 mm, 200 mm, and 1000 mm, respectively.</p>
Full article ">Figure 7
<p>The relationship between the capacitance <span class="html-italic">C</span><sub>1</sub> of the plate/membrane contact region (0 ≤ <span class="html-italic">r</span> ≤ <span class="html-italic">b</span>), the capacitance <span class="html-italic">C</span><sub>2-3</sub> of the plate/membrane non-contact region (<span class="html-italic">b</span> ≤ <span class="html-italic">r</span> ≤ <span class="html-italic">a</span> and <span class="html-italic">C</span><sub>2-3</sub> = <span class="html-italic">C</span><sub>2</sub><span class="html-italic">C</span><sub>3</sub>/(<span class="html-italic">C</span><sub>2</sub> + <span class="html-italic">C</span><sub>3</sub>)), and the total capacitance <span class="html-italic">C</span> when <span class="html-italic">a</span> = 70 mm, <span class="html-italic">h</span> = 0.3 mm, <span class="html-italic">t</span> = 0.1 mm, <span class="html-italic">E</span> = 3.05 MPa, <span class="html-italic">ν</span> = 0.45, and <span class="html-italic">g</span> = 10 mm.</p>
Full article ">Figure 8
<p>The three <span class="html-italic">C–V</span> analytical relationships fitted to the numerical calculation values of the total capacitance <span class="html-italic">C</span> and rainwater volume <span class="html-italic">V</span> in <a href="#sensors-24-06291-t001" class="html-table">Table 1</a> when <span class="html-italic">a</span> = 70 mm, <span class="html-italic">h</span> = 0.3 mm, <span class="html-italic">t</span> = 0.1 mm, <span class="html-italic">E</span> = 3.05 MPa, <span class="html-italic">ν</span> = 0.45, and <span class="html-italic">D</span> = 10 mm, where Function 1 refers to the <span class="html-italic">C–V</span> analytical relationship fitted by a curve, Functions 2 and 3 refer to the <span class="html-italic">C–V</span> analytical relationships fitted by two straight lines, and the fitted analytical expressions of Functions 1, 2, and 3 are shown in <a href="#sensors-24-06291-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 9
<p>The effect of changing the initially parallel gap <span class="html-italic">D</span> on <span class="html-italic">C</span>–<span class="html-italic">V</span> relationship when <span class="html-italic">a</span> = 70 mm, <span class="html-italic">h</span> = 0.3 mm, <span class="html-italic">t</span> = 0.1 mm, <span class="html-italic">E</span> = 3.05 MPa, <span class="html-italic">ν</span> = 0.45, and <span class="html-italic">D</span> takes 5 mm, 10 mm, and 15 mm, respectively.</p>
Full article ">Figure 10
<p>The effect of changing the insulator layer thickness <span class="html-italic">t</span> on the <span class="html-italic">C</span>–<span class="html-italic">V</span> relationship when <span class="html-italic">a</span> = 70 mm, <span class="html-italic">D</span> = 10 mm, <span class="html-italic">h</span> = 0.3 mm, <span class="html-italic">E</span> = 3.05 MPa, <span class="html-italic">ν</span> = 0.45, and <span class="html-italic">t</span> takes 0.1 mm, 0.15 mm, and 0.3 mm, respectively.</p>
Full article ">Figure 11
<p>The effect of changing the Young’s modulus of elasticity <span class="html-italic">E</span> on the <span class="html-italic">C</span>–<span class="html-italic">V</span> relationship when <span class="html-italic">a</span> = 70 mm, <span class="html-italic">D</span> = 10 mm, <span class="html-italic">h</span> = 0.3 mm, <span class="html-italic">t</span> = 0.1 mm, <span class="html-italic">ν</span> = 0.45, and <span class="html-italic">E</span> takes 7.84 MPa, 5.45 MPa, and 3.05 MPa, respectively.</p>
Full article ">Figure 12
<p>The effect of changing the Poisson’s ratio <span class="html-italic">v</span> on the <span class="html-italic">C</span>–<span class="html-italic">V</span> relationship when <span class="html-italic">a</span> = 70 mm, <span class="html-italic">D</span> = 10 mm, <span class="html-italic">h</span> = 0.3 mm, <span class="html-italic">t</span> = 0.1 mm, <span class="html-italic">E</span> = 3.05 MPa, and <span class="html-italic">v</span> takes 0.45, 0.3, and 0.15, respectively.</p>
Full article ">Figure 13
<p>The effect of changing the membrane thickness <span class="html-italic">h</span> on the <span class="html-italic">C</span>–<span class="html-italic">V</span> relationship when <span class="html-italic">a</span> = 70 mm, <span class="html-italic">D</span> = 10 mm, <span class="html-italic">t</span> = 0.1 mm, <span class="html-italic">E</span> = 3.05 MPa, <span class="html-italic">ν</span> = 0.45, and <span class="html-italic">h</span> takes 0.3 mm, 0.45 mm, and 0.6 mm, respectively.</p>
Full article ">Figure 14
<p>The effect of changing the circular membrane radius <span class="html-italic">a</span> on the <span class="html-italic">C</span>–<span class="html-italic">V</span> relationship when <span class="html-italic">D</span> = 10 mm, <span class="html-italic">h</span> = 0.3 mm, <span class="html-italic">t</span> = 0.1 mm, <span class="html-italic">E</span> = 3.05 MPa, <span class="html-italic">ν</span> = 0.45, and <span class="html-italic">a</span> takes 80 mm, 70 mm, and 60 mm, respectively.</p>
Full article ">
13 pages, 3334 KiB  
Article
Gelatin-Coated High-Sensitivity Microwave Sensor for Humidity-Sensing Applications
by Junho Yeo and Younghwan Kwon
Sensors 2024, 24(19), 6286; https://doi.org/10.3390/s24196286 - 28 Sep 2024
Viewed by 391
Abstract
In this paper, the humidity-sensing characteristics of gelatin were compared with those of poly(vinyl alcohol) (PVA) at L-band (1 ~ 2 GHz) microwave frequencies. A capacitive microwave sensor based on a defected ground structure with a modified interdigital capacitor (DGS-MIDC) in a microstrip [...] Read more.
In this paper, the humidity-sensing characteristics of gelatin were compared with those of poly(vinyl alcohol) (PVA) at L-band (1 ~ 2 GHz) microwave frequencies. A capacitive microwave sensor based on a defected ground structure with a modified interdigital capacitor (DGS-MIDC) in a microstrip transmission line operating at 1.5 GHz without any coating was used. Gelatin is a natural polymer based on protein sourced from animal collagen, whereas PVA is a high-sensitivity hydrophilic polymer that is widely used for humidity sensors and has a good film-forming property. Two DGS-MIDC-based microwave sensors coated with type A gelatin and PVA, respectively, with a thickness of 0.02 mm were fabricated. The percent relative frequency shift (PRFS) and percent relative magnitude shift (PRMS) based on the changes in the resonant frequency and magnitude level of the transmission coefficient for the microwave sensor were used to compare the humidity-sensing characteristics. The relative humidity (RH) was varied from 50% to 80% with a step of 10% at a fixed temperature of around 25 °C using a low-reflective temperature and humidity chamber manufactured with Styrofoam. The experiment’s results show that the capacitive humidity sensitivity of the gelatin-coated microwave sensor in terms of the PRFS and PRMS was higher compared to that of the PVA-coated one. In particular, the sensitivity of the gelatin-coated microwave sensor at a low RH from 50% to 60% was much greater compared to that of the PVA-coated one. In addition, the relative permittivity of the fabricated microwave sensors coated with PVA and gelatin was extracted by using the measured PRFS and the equation was derived by curve-fitting the simulated results. The change in the extracted relative permittivity for the gelatin-coated microwave sensor was larger than that of the PVA-coated one for varying the RH. Full article
(This article belongs to the Special Issue RF and IoT Sensors: Design, Optimization and Applications)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of (<b>a</b>) PVA and (<b>b</b>) gelatin.</p>
Full article ">Figure 2
<p>DGS-MIDC-based microwave sensor: (<b>a</b>) geometry, (<b>b</b>) electric-field distribution at 1.5 GHz, and (<b>c</b>) S-parameter characteristics and simplified equivalent circuit model.</p>
Full article ">Figure 3
<p>Performance characteristics of the DGS-MIDC-based microwave sensor for varying relative permittivity of the coated polymer with tan <span class="html-italic">δ</span> = 0: (<b>a</b>) S<sub>21</sub>, (<b>b</b>) <span class="html-italic">f</span><sub>r</sub>, and (<b>c</b>) PRFS.</p>
Full article ">Figure 4
<p>Extracted equivalent circuit parameters of the DGS-MIDC-based microwave sensor for varying relative permittivity of the coated polymer with tan <span class="html-italic">δ</span> = 0: (<b>a</b>) <span class="html-italic">C</span><sub>1</sub>; (<b>b</b>) <span class="html-italic">L</span><sub>1</sub>; (<b>c</b>) Δ<span class="html-italic">C</span><sub>1</sub>/<span class="html-italic">C</span><sub>1</sub>(%) and (<b>d</b>) Δ<span class="html-italic">L</span><sub>1</sub>/<span class="html-italic">L</span><sub>1</sub>(%).</p>
Full article ">Figure 5
<p>Photographs of the fabricated microwave sensors coated with (<b>a</b>) PVA and (<b>b</b>) gelatin.</p>
Full article ">Figure 6
<p>Block diagram and photographs of the experiment setup for the humidity-sensing measurements: (<b>a</b>) block diagram, (<b>b</b>) experiment setup with an open-top cover, and (<b>c</b>) experiment setup with closed-top cover.</p>
Full article ">Figure 7
<p>Measured S<sub>21</sub> characteristics of the fabricated microwave sensors coated with the polymers for varying RH. (<b>a</b>) PVA and (<b>b</b>) gelatin.</p>
Full article ">Figure 8
<p>Performance comparison of the fabricated microwave sensors coated with the polymers for varying RH. (<b>a</b>) <span class="html-italic">f</span><sub>r</sub>, (<b>b</b>) PRFS, (<b>c</b>) S<sub>21</sub> magnitude, and (<b>d</b>) PRMS.</p>
Full article ">Figure 9
<p>Comparison of the extracted relative permittivity from measured PRFSs of PVA- and gelatin-coated microwave sensors for varying RH.</p>
Full article ">
9 pages, 1032 KiB  
Article
Dirac Electrons with Molecular Relaxation Time at Electrochemical Interface between Graphene and Water
by Alexey V. Butko, Vladimir Y. Butko and Yurii A. Kumzerov
Int. J. Mol. Sci. 2024, 25(18), 10083; https://doi.org/10.3390/ijms251810083 - 19 Sep 2024
Viewed by 468
Abstract
The time dynamics of charge accumulation at the electrochemical interface between graphene and water is important for supercapacitors, batteries, and chemical and biological sensors. By using impedance spectroscopy, we have found that measured capacitance (Cm) at this interface with the gate [...] Read more.
The time dynamics of charge accumulation at the electrochemical interface between graphene and water is important for supercapacitors, batteries, and chemical and biological sensors. By using impedance spectroscopy, we have found that measured capacitance (Cm) at this interface with the gate voltage Vgate ≈ 0.1 V follows approximate laws Cm~T1.2 and Cm~T0.11 (T is Vgate period) in frequency ranges (1000–50,000) Hz and (0.02–300) Hz, respectively. In the first range, this dependence demonstrates that the interfacial capacitance (Cint) is only partially charged during the charging period. The observed weaker frequency dependence of the measured capacitance (Cm) at frequencies below 300 Hz is primarily determined by the molecular relaxation of the double-layer capacitance (Cdl) and by the graphene quantum capacitance (Cq), and it also implies that Cint is mostly charged. We have also found a voltage dependence of Cm below 10 Hz, which is likely related to the voltage dependence of Cq. The observation of this effect only at low frequencies indicates that Cq relaxation time is much longer than is typical for electron processes, probably due to Dirac cone reconstruction from graphene electrons with increased effective mass as a result of their quasichemical bonding with interfacial molecular charges. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustrations of equivalent circuits of C<sub>int</sub>.</p>
Full article ">Figure 2
<p>Typical frequency dependencies of the capacitance measured in water between graphene and gate electrodes normalized to the graphene surface area (<b>a</b>) at V<sub>gate</sub> = 0.1 V with a single gold wire gate electrode (graphene#1) and a gate electrode of 6 gold wires (graphene#2) on the logarithmic scale, (<b>b</b>) at 8 different gate voltages with a gate electrode of 6 gold wires on the logarithmic scale, (<b>c</b>) at 7 different gate voltages with a single gold wire gate electrode on the logarithmic scale, (<b>d</b>) at 8 different gate voltages with a gate electrode of 6 gold wires on the linear scale, and (<b>e</b>) at 7 different gate voltages with a single gold wire gate electrode on the linear scale.</p>
Full article ">Figure 2 Cont.
<p>Typical frequency dependencies of the capacitance measured in water between graphene and gate electrodes normalized to the graphene surface area (<b>a</b>) at V<sub>gate</sub> = 0.1 V with a single gold wire gate electrode (graphene#1) and a gate electrode of 6 gold wires (graphene#2) on the logarithmic scale, (<b>b</b>) at 8 different gate voltages with a gate electrode of 6 gold wires on the logarithmic scale, (<b>c</b>) at 7 different gate voltages with a single gold wire gate electrode on the logarithmic scale, (<b>d</b>) at 8 different gate voltages with a gate electrode of 6 gold wires on the linear scale, and (<b>e</b>) at 7 different gate voltages with a single gold wire gate electrode on the linear scale.</p>
Full article ">Figure 3
<p>Schematic illustrations of equivalent circuits of C<sub>m</sub>.</p>
Full article ">Figure 4
<p>Schematic illustrations of the bonding between Dirac electrons in graphene and molecular charges at the electrochemical interface between graphene and water and of the Dirac cone formed from the bonded electrons.</p>
Full article ">
14 pages, 975 KiB  
Article
Evaluation of a Multivariate Calibration Model for the WET Sensor That Incorporates Apparent Dielectric Permittivity and Bulk Soil Electrical Conductivity
by Panagiota Antonia Petsetidi and George Kargas
Land 2024, 13(9), 1490; https://doi.org/10.3390/land13091490 - 14 Sep 2024
Viewed by 612
Abstract
The measurement of apparent dielectric permittivity (εs) by low-frequency capacitance sensors and its conversion to the volumetric water content of soil (θ) through a factory calibration is a valuable tool in precision irrigation. Under certain soil conditions, however, εs readings [...] Read more.
The measurement of apparent dielectric permittivity (εs) by low-frequency capacitance sensors and its conversion to the volumetric water content of soil (θ) through a factory calibration is a valuable tool in precision irrigation. Under certain soil conditions, however, εs readings are substantially affected by the bulk soil electrical conductivity (ECb) variability, which is omitted in default calibration, leading to inaccurate θ estimations. This poses a challenge to the reliability of the capacitance sensors that require soil-specific calibrations, considering the ECb impact to ensure the accuracy in θ measurements. In this work, a multivariate calibration equation (multivariate) incorporating both εs and ECb for the determination of θ by the capacitance WET sensor (Delta-T Devices Ltd., Cambridge, UK) is examined. The experiments were conducted in the laboratory using the WET sensor, which measured θ, εs, and ECb simultaneously over a range of soil types with a predetermined actual volumetric water content value (θm) ranging from θ = 0 to saturation, which were obtained by wetting the soils with four water solutions of different electrical conductivities (ECi). The multivariate model’s performance was evaluated against the univariate CAL and the manufacturer’s (Manuf) calibration methods with the Root Mean Square Error (RMSE). According to the results, the multivariate model provided the most accurate θ estimations, (RMSE ≤ 0.022 m3m−3) compared to CAL (RMSE ≤ 0.027 m3m−3) and Manuf (RMSE ≤ 0.042 m3m−3), across all the examined soils. This study validates the effects of ECb on θ for the WET and recommends the multivariate approach for improving the capacitance sensors’ accuracy in soil moisture measurements. Full article
Show Figures

Figure 1

Figure 1
<p>Relationship between the apparent dielectric permittivity (ε<sub>s</sub>) and the actual volumetric water content (θ<sub>m</sub>) at various salinity levels (EC<sub>i</sub>) (0.28, 1.2, 3, 6 dSm<sup>−1</sup>) for all the examined soils: (<b>a</b>) SL 1, (<b>b</b>) SL 2, (<b>c</b>) SL 3, (<b>d</b>) CL, (<b>e</b>) S, (<b>f</b>) L, and (<b>g</b>) C.</p>
Full article ">Figure 1 Cont.
<p>Relationship between the apparent dielectric permittivity (ε<sub>s</sub>) and the actual volumetric water content (θ<sub>m</sub>) at various salinity levels (EC<sub>i</sub>) (0.28, 1.2, 3, 6 dSm<sup>−1</sup>) for all the examined soils: (<b>a</b>) SL 1, (<b>b</b>) SL 2, (<b>c</b>) SL 3, (<b>d</b>) CL, (<b>e</b>) S, (<b>f</b>) L, and (<b>g</b>) C.</p>
Full article ">
18 pages, 6428 KiB  
Article
Calibration of Low-Cost Moisture Sensors in a Biochar-Amended Sandy Loam Soil with Different Salinity Levels
by María José Gómez-Astorga, Karolina Villagra-Mendoza, Federico Masís-Meléndez, Aníbal Ruíz-Barquero and Renato Rimolo-Donadio
Sensors 2024, 24(18), 5958; https://doi.org/10.3390/s24185958 - 13 Sep 2024
Viewed by 971
Abstract
With the increasing focus on irrigation management, it is crucial to consider cost-effective alternatives for soil water monitoring, such as multi-point monitoring with low-cost soil moisture sensors. This study assesses the accuracy and functionality of low-cost sensors in a sandy loam (SL) soil [...] Read more.
With the increasing focus on irrigation management, it is crucial to consider cost-effective alternatives for soil water monitoring, such as multi-point monitoring with low-cost soil moisture sensors. This study assesses the accuracy and functionality of low-cost sensors in a sandy loam (SL) soil amended with biochar at rates of 15.6 and 31.2 tons/ha by calibrating the sensors in the presence of two nitrogen (N) and potassium (K) commercial fertilizers at three salinity levels (non/slightly/moderately) and six soil water contents. Sensors were calibrated across nine SL-soil combinations with biochar and N and K fertilizers, counting for 21 treatments. The best fit for soil water content calibration was obtained using polynomial equations, demonstrating reliability with R2 values greater than 0.98 for each case. After a second calibration, low-cost soil moisture sensors provide acceptable results concerning previous calibration, especially for non- and slightly saline treatments and at soil moisture levels lower than 0.17 cm3cm−3. The results showed that at low frequencies, biochar and salinity increase the capacitance detected by the sensors, with calibration curves deviating up to 30% from the control sandy loam soil. Due to changes in the physical and chemical properties of soil resulting from biochar amendments and the conductive properties influenced by fertilization practices, it is required to conduct specific and continuous calibrations of soil water content sensor, leading to better agricultural management decisions. Full article
(This article belongs to the Section Smart Agriculture)
Show Figures

Figure 1

Figure 1
<p>Soil moisture sensor used (<b>a</b>) capacitive soil moisture sensor v1.2 and (<b>b</b>) set of 5 soil moisture sensors installed for calibration.</p>
Full article ">Figure 2
<p>Sensor output performance over time under non-saline (NS) conditions for six soil moisture levels.</p>
Full article ">Figure 3
<p>Sensor output dispersion per treatment under non-saline (NS) conditions for six soil moisture levels.</p>
Full article ">Figure 4
<p>Sensor output performance over time under slightly saline (SS) conditions for different soil moisture levels.</p>
Full article ">Figure 5
<p>Sensor output dispersion per treatment under slightly saline (SS) conditions for different soil moisture levels.</p>
Full article ">Figure 6
<p>Sensor output performance over time under moderately saline (MS) conditions for six soil moisture levels.</p>
Full article ">Figure 7
<p>Sensor output dispersion per treatment under moderately saline (MS) conditions for six soil moisture levels.</p>
Full article ">Figure 8
<p>Calibration curves for all treatments under the NS, SS, and MS scenarios. The blue continuous line indicates the calibration curve for SLB0 without fertilizer and the gray lines represent the error deviation of 10%, 20%, and 30% with respect to SLB0.</p>
Full article ">Figure 9
<p>Performance of the soil moisture sensor through errors in treatment calibrations in a 2-week interval.</p>
Full article ">Figure 10
<p>Heat map of the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>M</mi> <mi>S</mi> <mi>E</mi> </mrow> </semantics></math> pattern for all treatments for (<b>a</b>) first calibration and (<b>b</b>) second calibration after a 2-week interval.</p>
Full article ">
13 pages, 3430 KiB  
Article
Assessment of Low-Cost and Higher-End Soil Moisture Sensors across Various Moisture Ranges and Soil Textures
by Rajesh Nandi and Dev Shrestha
Sensors 2024, 24(18), 5886; https://doi.org/10.3390/s24185886 - 11 Sep 2024
Viewed by 974
Abstract
The accuracy and unit cost of sensors are important factors for a continuous soil moisture monitoring system. This study compares the accuracy of four soil moisture sensors differing in unit costs in coarse-, fine-, and medium-textured soils. The sensor outputs were recorded for [...] Read more.
The accuracy and unit cost of sensors are important factors for a continuous soil moisture monitoring system. This study compares the accuracy of four soil moisture sensors differing in unit costs in coarse-, fine-, and medium-textured soils. The sensor outputs were recorded for the VWC, ranging from 0% to 50%. Low-cost capacitive and resistive sensors were evaluated with and without the external 16-bit analog-to-digital converter ADS1115 to improve their performances without adding much cost. Without ADS1115, using only Arduino’s built-in analog-to-digital converter, the low-cost sensors had a maximum RMSE of 4.79% (v/v) for resistive sensors and 3.78% for capacitive sensors in medium-textured soil. The addition of ADS1115 showed improved performance of the low-cost sensors, with a maximum RMSE of 2.64% for resistive sensors and 1.87% for capacitive sensors. The higher-end sensors had an RMSE of up to 1.8% for VH400 and up to 0.95% for the 5TM sensor. The RMSE differences between higher-end and low-cost sensors with the use of ADS1115 were not statistically significant. Full article
(This article belongs to the Special Issue Sensor-Based Crop and Soil Monitoring in Precise Agriculture)
Show Figures

Figure 1

Figure 1
<p>FC-28 resistive sensor circuit diagram (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> = 5 V and <math display="inline"><semantics> <mrow> <mi>R</mi> </mrow> </semantics></math> = 10 kΩ. <span class="html-italic">R<sub>s</sub></span> represents the actual sensor. Analog output <span class="html-italic">Ao</span> can be either directly read from a microcontroller’s analog-to-digital converter (ADC) or read digitally if an external ADC is used).</p>
Full article ">Figure 2
<p>The capacitive soil probe circuit diagram. <math display="inline"><semantics> <mrow> <mi>R</mi> </mrow> </semantics></math> = 10 kΩ. Adapted from [<a href="#B17-sensors-24-05886" class="html-bibr">17</a>]. Physical pin numbers are indicated on 555 Timer.</p>
Full article ">Figure 3
<p>Sensor connected to ADS1115 breakout board (<b>left</b>) and ADS1115 block diagram (<b>right</b>), adapted from the user manual of ADS1115 [<a href="#B19-sensors-24-05886" class="html-bibr">19</a>]. Only one channel was used with unity gain. Reprinted with permission courtesy of Texas Instruments.</p>
Full article ">Figure 4
<p>The fine (<b>left</b>) and coarse (<b>right</b>) soil used in this study.</p>
Full article ">Figure 5
<p>Circuit connection diagram of (<b>a</b>) low-cost capacitive sensor (v.1.2), (<b>b</b>) FC-28 low-cost resistive sensor, (<b>c</b>) VH400 sensor, and (<b>d</b>) 5TM sensor.</p>
Full article ">Figure 6
<p>Experimental setup for this study with ADS1115.</p>
Full article ">Figure 7
<p>Calibration curve for sensors in coarse, medium, and fine soil: (<b>a</b>) capacitive, (<b>b</b>) resistive, (<b>c</b>) VH400, and (<b>d</b>) 5TM, each color representing a specific sensor. Each experiment was conducted with three replications.</p>
Full article ">Figure 8
<p>Calibration curve using ADS1115 with capacitive and resistive sensors, also differentiated by color in (<b>a</b>) coarse-, (<b>b</b>) fine-, and (<b>c</b>) medium-textured soil. Each experiment was conducted with three replications.</p>
Full article ">Figure 9
<p>ADS1115 frequency response at 8 samples per second data rate [<a href="#B19-sensors-24-05886" class="html-bibr">19</a>]. Reprinted with permission courtesy of Texas Instruments.</p>
Full article ">
8 pages, 411 KiB  
Article
Modeling Electronic Devices with a Casimir Cavity
by G. Jordan Maclay
Physics 2024, 6(3), 1124-1131; https://doi.org/10.3390/physics6030070 - 10 Sep 2024
Viewed by 1138
Abstract
The Casimir effect has been exploited in various MEMS (micro-electro-mechanical system) devices, especially to make sensitive force sensors and accelerometers. It has also been used to provide forces for a variety of purposes, for example, for the assembly of considerably small parts. Repulsive [...] Read more.
The Casimir effect has been exploited in various MEMS (micro-electro-mechanical system) devices, especially to make sensitive force sensors and accelerometers. It has also been used to provide forces for a variety of purposes, for example, for the assembly of considerably small parts. Repulsive forces and torques have been produced using various configurations of media and materials. Just a few electronic devices have been explored that utilize the electrical properties of the Casimir effect. Recently, experimental results were presented that described the operation of an electronic device that employed a Casimir cavity attached to a standard MIM (metal–insulator–metal) structure. The DC (direct current) conductance of the novel MIM device was enhanced by the attached cavity and found to be directly proportional to the capacitance of the attached cavity. The phenomenological model proposed assumed that the cavity reduced the vacuum fluctuations, which resulted in a reduced injection of carriers. The analysis presented here indicates that the optical cavity actually enhances vacuum fluctuations, which would predict a current in the opposite direction from that observed. Further, the vacuum fluctuations near the electrode are shown to be approximately independent of the size of the optical cavity, in disagreement with the experimental data which show a dependence on the size. Thus, the proposed mechanism of operation does not appear correct. A more detailed theoretical analysis of these devices is needed, in particular, one that uses real material parameters and computes the vacuum fluctuations for the entire device. Such an analysis would reveal how these devices operate and might suggest design principles for a new genre of electronic devices that make use of vacuum fluctuations. Full article
(This article belongs to the Section Atomic Physics)
Show Figures

Figure 1

Figure 1
<p>The MIMOC (metal–insulator–metal optical cavity) device. An optical cavity (OC) of thickness <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>c</mi> </mrow> </semantics></math> made from PMMA (polymethyl methacralate, spin coated photoresist) or SiO<sub>2</sub> is bounded by an aluminum mirror and a MIM interface. The latter consists of a palladium electrode of thickness <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </semantics></math>, a layer of insulator of thickness <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>I</mi> </mrow> </semantics></math>, and a thick nickel electrode. A current is positive if it flows from the Pd electrode to the grounded Ni electrode.</p>
Full article ">Figure 2
<p>The conductance in mS of the device shown in <a href="#physics-06-00070-f001" class="html-fig">Figure 1</a>, as a function of 100/thickness <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>c</mi> </mrow> </semantics></math> of the optical cavity made from PMMA. <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>c</mi> </mrow> </semantics></math> varies from 33 nm to 1100 nm for the data shown. The data are taken from Figure 3b of Ref. [<a href="#B7-physics-06-00070" class="html-bibr">7</a>] and Figure 4a of Ref. [<a href="#B8-physics-06-00070" class="html-bibr">8</a>]. The solid line just connects the data points. A linear fit <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>=</mo> <mn>0.3519</mn> <mi>x</mi> </mrow> </semantics></math> through the origin is also shown (as the dotted line) along with the coefficient of determination (<math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math>).</p>
Full article ">Figure 3
<p>The dimensionless normalized variance in energy, <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mn>12</mn> <mo>/</mo> <msup> <mi>q</mi> <mn>2</mn> </msup> <msup> <mi>v</mi> <mn>2</mn> </msup> <mo>)</mo> </mrow> <mfenced separators="" open="&#x2329;" close="&#x232A;"> <msup> <mrow> <mo>(</mo> <mo>Δ</mo> <mi>U</mi> <mo>)</mo> </mrow> <mn>2</mn> </msup> </mfenced> <mo>=</mo> <mn>1</mn> <mo>+</mo> <mn>3</mn> <msup> <mo form="prefix">csc</mo> <mn>2</mn> </msup> <mrow> <mo>(</mo> <mrow> <mi>π</mi> <msub> <mi>z</mi> <mn>0</mn> </msub> </mrow> <mo>/</mo> <mi>a</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, from Equation (<a href="#FD8-physics-06-00070" class="html-disp-formula">8</a>) as a function of the location within the cavity <span class="html-italic">z</span> nm for a cavity of width <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> nm. The variance increases without bound at the locations of the plates, <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> nm and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> nm.</p>
Full article ">Figure 4
<p>Normalized variance in energy for cavities of width for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>33</mn> </mrow> </semantics></math> nm (black dashed) and <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1100</mn> </mrow> </semantics></math> nm (in red). Near the origin, the variances are almost identical.</p>
Full article ">Figure 5
<p>Fractional difference in variance for optical cavities of width 33 nm and 1100 nm, corresponding to data in <a href="#physics-06-00070-f004" class="html-fig">Figure 4</a>. The fractional difference is calculated as the ratio of the difference of the variances at 33 and 1100 nm to the variance at 33 nm.</p>
Full article ">
20 pages, 3135 KiB  
Article
Parameter Optimization and Capacitance-Based Monitoring of In Situ Cell Detachment in Microcarrier Cultures
by Atefeh Ebrahimian, Mona Schalk, Mark Dürkop, Michael Maurer, Rudolf Bliem and Harald Kühnel
Processes 2024, 12(9), 1887; https://doi.org/10.3390/pr12091887 - 3 Sep 2024
Viewed by 657
Abstract
This study delves into the scale-down optimization of the in situ cell detachment process for MA 104 cells cultivated on Cytodex 1 microcarriers (MCs). Through a systematic exploration, critical operational parameters—the agitation speed, incubation time, Trypsin–EDTA volume and corresponding activity, and washing steps—were [...] Read more.
This study delves into the scale-down optimization of the in situ cell detachment process for MA 104 cells cultivated on Cytodex 1 microcarriers (MCs). Through a systematic exploration, critical operational parameters—the agitation speed, incubation time, Trypsin–EDTA volume and corresponding activity, and washing steps—were identified as key factors influencing the efficiency and scalability of in situ cell detachment in microcarrier-based cell culture. Maintaining an appropriate agitation speed (1.25 × Njs, minimum agitation speed at which no microcarriers remain stationary for the signification period of 5 s), optimizing the Trypsinization incubation time (up to 60 min), and implementing multiple washing steps (two times) post-medium removal were found to be crucial for efficient cell detachment and subsequent growth. Our study demonstrates the feasibility of reducing the final Trypsin volume to 50 mL per gram of microcarrier while maintaining a Trypsin activity above 380 USP/mL. These conditions ensure complete cell dissociation and improve the cost effectiveness in large-scale productions. Additionally, we introduced real-time monitoring using a capacitance sensor during in situ cell detachment. This method has proven to be an effective process analytical technology (PAT) tool for tracking the cell detachment progress and efficiency. It allows for the prediction of cell detachment based on signals recorded between 3 and 7 min of Trypsinization, enabling rapid process decisions without the need for offline sampling, thereby enhancing the overall process control. This systematic approach not only optimizes in situ cell detachment processes but also has significant implications for the scalability and efficiency of microcarrier-based cell culture systems. Full article
(This article belongs to the Section Biological Processes and Systems)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Cell recovery percentage during Trypsinization at two periods, 10 and 20 min, for two different agitation speeds, 100 rpm (1.25 × N<sub>js</sub>) and 200 rpm (2.5 × N<sub>js</sub>), in 1 L DASGIP bioreactor; (<b>b</b>) cell growth on MCs after undergoing in situ cell detachment. For each condition, two separate experiments were performed.</p>
Full article ">Figure 2
<p>Effect of incubation time: (<b>a</b>) cell detachment efficiency over different incubation times with Trypsin solution (0.25%) (ns: not significant), (<b>b</b>) cell growth monitoring on MCs post-detachment using Trypsin solution (0.25%) at different incubation times, and (<b>c</b>,<b>d</b>) single cells subjected with different incubation time with Trypsin solution (0.25%) and subsequently seeded on plates, with the monitoring of apoptotic (RLU) and necrotic (RFU) cells over 24 h.</p>
Full article ">Figure 3
<p>Impact of different volume ratios of Trypsin–EDTA solution (0.25%) at the time of cell detachment: (<b>a</b>) cell detachment efficiency over cell treatment with Trypsin solution (0.25%) with two ratios; (<b>b</b>) subsequent cell growth after cell detachment with two different ratios of Trypsin solution (0.25%).</p>
Full article ">Figure 4
<p>Impact of washing step times on cell detachment and subsequent cell growth: (<b>a</b>) efficiency of cell detachment during treatment with Trypsin–EDTA (0.25%) solution combined with one or two steps of PBS washing at a 40% iv; (<b>b</b>) subsequent cell growth post detachment; (<b>c</b>) monitoring the cell aggregation on MCs during the culture after undergoing one-time and two-time PBS washing during in situ cell detachment: (<b>c.1</b>,<b>c.2</b>) notable presence of cell aggregation on MCs at 2nd day and 4th day of culture for the condition of one-time PBS washing, and (<b>c.3</b>,<b>c.4</b>) no notable cell aggregation on MCs at 2nd day and 4th day of culture for condition of two time-PBS washing; (<b>d</b>) comparison of observed aggregated cells on MCs, and the level of full (confluent) or partial (semi-confluent) coverage of MCs at the end of culture with one or two PBS washing steps (1X PBS or 2X PBS) before enzymatic cell detachment.</p>
Full article ">Figure 5
<p>Cell dissociation monitoring: (<b>a</b>) development of cell dissociation over the course of in situ cell detachment monitored by taking samples and offline measurement of cell density; (<b>b</b>) permittivity signal over cell dissociation with different final Trypsin concentrations (measured with two independent capacitance sensors per each enzyme concentration).</p>
Full article ">Figure 6
<p>The predictive performance of the model using permittivity signals obtained from capacitance sensors across seven experimental runs with different Trypsin concentrations: 1.65 g/L (runs 1 and 2), 1.29 g/L (run 3), 0.86 g/L (run 4), 0.82 g/L (run 5), 0.64 g/L (run 6), and 0.43 g/L (run 7).</p>
Full article ">Figure 7
<p>(<b>a</b>) The effectiveness of cell release during trypsin treatment; (<b>b</b>) measured permittivity signal (red) and predicted permittivity signal (blue) during in situ cell detachment using Trypsin solutions with high enzyme concentration but lower activity (1.29 g/L, 488 USP/mL); (<b>c</b>) permittivity signals from two generations of capacitance sensors for two distinct enzyme concentrations: 0.86 g/L and 0.43 g/L.</p>
Full article ">
Back to TopTop