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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (10)

Search Parameters:
Keywords = bi-platform calibration

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 1531 KiB  
Article
A Prospective Study on Risk Prediction of Preeclampsia Using Bi-Platform Calibration and Machine Learning
by Zhiguo Zhao, Jiaxin Dai, Hongyan Chen, Lu Lu, Gang Li, Hua Yan and Junying Zhang
Int. J. Mol. Sci. 2024, 25(19), 10684; https://doi.org/10.3390/ijms251910684 - 4 Oct 2024
Viewed by 446
Abstract
Preeclampsia is a pregnancy syndrome characterized by complex symptoms which cause maternal and fetal problems and deaths. The aim of this study is to achieve preeclampsia risk prediction and early risk prediction in Xinjiang, China, based on the placental growth factor measured using [...] Read more.
Preeclampsia is a pregnancy syndrome characterized by complex symptoms which cause maternal and fetal problems and deaths. The aim of this study is to achieve preeclampsia risk prediction and early risk prediction in Xinjiang, China, based on the placental growth factor measured using the SiMoA or Elecsys platform. A novel reliable calibration modeling method and missing data imputing method are proposed, in which different strategies are used to adapt to small samples, training data, test data, independent features, and dependent feature pairs. Multiple machine learning algorithms were applied to train models using various datasets, such as single-platform versus bi-platform data, early pregnancy versus early plus non-early pregnancy data, and real versus real plus augmented data. It was found that a combination of two types of mono-platform data could improve risk prediction performance, and non-early pregnancy data could enhance early risk prediction performance when limited early pregnancy data were available. Additionally, the inclusion of augmented data resulted in achieving a high but unstable performance. The models in this study significantly reduced the incidence of preeclampsia in the region from 7.2% to 2.0%, and the mortality rate was reduced to 0%. Full article
(This article belongs to the Section Molecular Informatics)
Show Figures

Figure 1

Figure 1
<p>Framework of PE risk prediction based on RF and bi-platform calibration. (<b>a</b>) Collecting PE case group sample data and control group sample data; (<b>b</b>) coding features, imputing missing data with MLP networks, and normalizing features; (<b>c</b>) calibrating PlGF from the two platforms with an MLP model; and (<b>d</b>) constructing PE risk prediction model and predicting PE risk of test samples.</p>
Full article ">Figure 2
<p>Missing data imputation based on MLP (missing data are represented by imputed orange). (<b>a</b>) The training process: Take the pair of pre-pregnancy weight and current weight of case group as an example. (<b>b</b>) Training process: Other features are imputed by intra-class median. (<b>c</b>) Test process: The missing data of other features are imputed by the median.</p>
Full article ">Figure 3
<p>PlGF value calibration based on MLP. Since MSE<sub>2</sub> &lt; MSE<sub>1</sub>, the PlGF values detected from the SiMoA platform do not need to be calibrated, while the PlGF values detected from the Elecsys platform are to be calibrated with MLP<sub>4</sub>.</p>
Full article ">Figure 4
<p>Ranking of feature importance obtained from the model trained from (<b>a</b>) Simoa Set, (<b>b</b>) Elecsys Set, (<b>c</b>) Simoa_Elecsys Set, and (<b>d</b>) First_Trimester Set, where the features ranked in the top 5 are colored red, the features ranked from 6th to 10th are colored yellow, and those ranked from 11th to 22nd are colored green.</p>
Full article ">
22 pages, 7288 KiB  
Article
Research on Tree Flash Fault Localization of Hybrid Overhead–Underground Lines Based on Improved Double-Ended Traveling Wave Method
by Zukang Huang, Chunhua Fang, Quancai Jiang, Tao Hu and Junjie Lv
Appl. Sci. 2024, 14(11), 4739; https://doi.org/10.3390/app14114739 - 30 May 2024
Viewed by 719
Abstract
The occurrence of tree flash faults in hybrid overhead–underground lines presents a significant challenge to the smooth operation of power systems. However, research on localizing such faults is relatively scarce. This study conducts theoretical analyses on the formation of tree flash faults, constructs [...] Read more.
The occurrence of tree flash faults in hybrid overhead–underground lines presents a significant challenge to the smooth operation of power systems. However, research on localizing such faults is relatively scarce. This study conducts theoretical analyses on the formation of tree flash faults, constructs a tree flash fault discharge test platform, and simulates the discharge process. The tree flash fault discharge traveling wave signals were obtained through a high-frequency current acquisition system. Additionally, this paper establishes a model for the current traveling wave of tree flash faults and analyzes transmission attenuation. To enhance the bi-terminal traveling wave localization method, we introduce modal decomposition and the Hilbert–Huang transform. Modal decomposition is used to disentangle signals and derive the instantaneous frequencies of modal signal components through the Hilbert–Huang transform. This process helps determine the time at which the initial wavefront reaches the terminals of the mixed-line transmission. The simulation analysis carried out using PSCAD/EMTDC v4.6.3 demonstrates that this method effectively calibrates the wavefront timing of tree flash fault signals without requiring knowledge of their wave velocity along the mixed-line transmission. Therefore, this approach achieves precise localization of tree flash faults efficiently. Full article
(This article belongs to the Topic Power System Protection)
Show Figures

Figure 1

Figure 1
<p>Illustration of tree flash fault occurrence.</p>
Full article ">Figure 2
<p>Schematic diagram of the tree flash fault discharge test platform.</p>
Full article ">Figure 3
<p>Tree flash fault discharge process.</p>
Full article ">Figure 4
<p>Tree flash fault test discharge current waveform (voltage value of booster system: 93 kV). (<b>a</b>) Raw current waveform plot; (<b>b</b>) magnified view of waveform; (<b>c</b>) tree flash discharge waveform a.</p>
Full article ">Figure 5
<p>Tree flash fault test discharge current waveform (voltage value of booster system: 103 kV). (<b>a</b>) Raw current waveform plot; (<b>b</b>) magnified view of waveform; (<b>c</b>) tree flash discharge waveform b.</p>
Full article ">Figure 6
<p>Tree flash fault test discharge current waveform (voltage value of booster system: 110 kV). (<b>a</b>) Raw current waveform plot; (<b>b</b>) magnified view of waveform; (<b>c</b>) tree flash discharge waveform c.</p>
Full article ">Figure 7
<p>Tree flash fault test discharge current waveform (voltage value of booster system: 118 kV). (<b>a</b>) Raw current waveform plot; (<b>b</b>) magnified view of waveform; (<b>c</b>) tree flash discharge waveform d.</p>
Full article ">Figure 8
<p>Tree flash fault discharge waveform spectrum.</p>
Full article ">Figure 9
<p>Topological structure of hybrid overhead–underground line model.</p>
Full article ">Figure 10
<p>Presentation of the 220 kV underground cable line.</p>
Full article ">Figure 11
<p>The tree flash fault current simulation component in PSCAD.</p>
Full article ">Figure 12
<p>Characteristics of tree flash fault current waveform attenuation.</p>
Full article ">Figure 13
<p>Localized characteristics of tree flash fault current waveform attenuation.</p>
Full article ">Figure 14
<p>The measured zero-mode component at terminal M and its VMD decomposition results.</p>
Full article ">Figure 15
<p>The VMD-Hilbert transform results of the zero-mode component at measurement terminal M.</p>
Full article ">Figure 16
<p>The zero-mode signal at measurement terminal N and its VMD decomposition results.</p>
Full article ">Figure 17
<p>The VMD-Hilbert transform results of the zero-mode component at measurement terminal N.</p>
Full article ">Figure 18
<p>The line-mode component at measurement terminal M and its VMD decomposition results.</p>
Full article ">Figure 19
<p>The VMD-Hilbert transform results of the line-mode component at measurement terminal M.</p>
Full article ">Figure 20
<p>The line-mode component at measurement terminal N and its VMD decomposition results.</p>
Full article ">Figure 21
<p>The VMD-Hilbert transform results of the line-mode component at measurement terminal N.</p>
Full article ">
14 pages, 4752 KiB  
Article
Coordinating Etching Inspired Synthesis of Fe(OH)3 Nanocages as Mimetic Peroxidase for Fluorescent and Colorimetric Self-Tuning Detection of Ochratoxin A
by Hongshuai Zhu, Bingfeng Wang and Yingju Liu
Biosensors 2023, 13(6), 665; https://doi.org/10.3390/bios13060665 - 19 Jun 2023
Cited by 1 | Viewed by 1698
Abstract
The development of multifunctional biomimetic nanozymes with high catalytic activity and sensitive response is rapidly advancing. The hollow nanostructures, including metal hydroxides, metal-organic frameworks, and metallic oxides, possess excellent loading capacity and a high surface area-to-mass ratio. This characteristic allows for the exposure [...] Read more.
The development of multifunctional biomimetic nanozymes with high catalytic activity and sensitive response is rapidly advancing. The hollow nanostructures, including metal hydroxides, metal-organic frameworks, and metallic oxides, possess excellent loading capacity and a high surface area-to-mass ratio. This characteristic allows for the exposure of more active sites and reaction channels, resulting in enhanced catalytic activity of nanozymes. In this work, based on the coordinating etching principle, a facile template-assisted strategy for synthesizing Fe(OH)3 nanocages by using Cu2O nanocubes as the precursors was proposed. The unique three-dimensional structure of Fe(OH)3 nanocages endows it with excellent catalytic activity. Herein, in the light of Fe(OH)3-induced biomimetic nanozyme catalyzed reactions, a self-tuning dual-mode fluorescence and colorimetric immunoassay was successfully constructed for ochratoxin A (OTA) detection. For the colorimetric signal, 2,2′-azino-bis (3-ethylbenzothiazoline-6-sulfonic acid) diammonium salt (ABTS) can be oxidized by Fe(OH)3 nanocages to form a color response that can be preliminarily identified by the human eye. For the fluorescence signal, the fluorescence intensity of 4-chloro-1-naphthol (4-CN) can be quantitatively quenched by the valence transition of Ferric ion in Fe(OH)3 nanocages. Due to the significant self-calibration, the performance of the self-tuning strategy for OTA detection was substantially enhanced. Under the optimized conditions, the developed dual-mode platform accomplishes a wide range of 1 ng/L to 5 μg/L with a detection limit of 0.68 ng/L (S/N = 3). This work not only develops a facile strategy for the synthesis of highly active peroxidase-like nanozyme but also achieves promising sensing platform for OTA detection in actual samples. Full article
(This article belongs to the Section Nano- and Micro-Technologies in Biosensors)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Schematic diagram of Fe(OH)<sub>3</sub> nanocages by the coordinating etching principle, CEP, coordinating etching and precipitating; CE, coordinating etching. (<b>B</b>) TEM images of the time-dependence of Fe(OH)<sub>3</sub> nanocages for different times.</p>
Full article ">Figure 2
<p>(<b>A</b>) SEM, (<b>B</b>) N<sub>2</sub> absorption–desorption isotherm, (<b>C</b>) pore size distribution, (<b>D</b>) elemental mappings, (<b>E</b>) EDS, and (<b>F</b>) XRD of Fe(OH)<sub>3</sub> nanocages. (<b>G</b>) UV–vis spectra of Fe(OH)<sub>3</sub>@Ab<sub>2</sub>.</p>
Full article ">Figure 3
<p>XPS survey (<b>A</b>), and the deconvolution of Fe 2p (<b>B</b>) and O 1s (<b>C</b>) of Fe(OH)<sub>3</sub> nanocages.</p>
Full article ">Figure 4
<p>Lineweaver−Burk plots of Fe(OH)<sub>3</sub> nanocages and Cu<sub>2</sub>O nanocubes by tuning the substrate of ABTS at a fixed 4.84 mM H<sub>2</sub>O<sub>2</sub> (<b>A</b>), and tuning the H<sub>2</sub>O<sub>2</sub> concentration at a fixed 3.87 mM ABTS (<b>B</b>), respectively.</p>
Full article ">Figure 5
<p>Fe(OH)<sub>3</sub> catalyst reaction mechanism and energy change (insets are the structure of relevant molecules (ions); the center is iron ion, purple is ghost atom, red is oxygen atom, and white is hydrogen atom).</p>
Full article ">Figure 6
<p>Fluorescence spectra (<b>A</b>) and UV–vis spectra (<b>B</b>) for different concentrations of OTA, (a) 0, (b) 0.001, (c) 0.0025, (d) 0.005, (e) 0.01, (f) 0.025, (g) 0.05, (h) 0.1, (i) 0.25, (j) 0.5, (k) 1, and (l) 5 μg/L (from bottom to top in A and B correspond to a–l, respectively). Inset in B was the photographs of the colorimetric method at different concentrations of OTA. (<b>C</b>,<b>D</b>) Calibration curve based on the fluorescence spectra (<b>A</b>) and UV–vis spectra (<b>B</b>), respectively. (<b>E</b>,<b>F</b>) Pearson correlation analysis of the self–tuning immunosensor (<b>E</b>) and corresponding calibration curve (<b>F</b>).</p>
Full article ">Figure 7
<p>The stability (<b>A</b>,<b>B</b>) and reproducibility (<b>C</b>,<b>D</b>) of the self–tuning immunosensor, (<b>E</b>) selectivity of the self–tuning dual–modal immunosensor, and (<b>F</b>) the recoveries of the OTA detection in corn and millet samples.</p>
Full article ">Scheme 1
<p>Schematic illustration of the self–tuning dual–mode immunosensor for OTA detection: fluorescence and colorimetry.</p>
Full article ">
11 pages, 826 KiB  
Article
Effect of the Matrix and Target on the Accurate Quantification of Genomic and Plasmid DNA by Digital Polymerase Chain Reaction
by Nengwu Si, Jun Li, Hongfei Gao, Yunjing Li, Shanshan Zhai, Fang Xiao, Li Zhang, Gang Wu and Yuhua Wu
Agriculture 2023, 13(1), 127; https://doi.org/10.3390/agriculture13010127 - 3 Jan 2023
Cited by 1 | Viewed by 1819
Abstract
In polymerase chain reaction (PCR)-based nucleic acid quantification, the DNA template type, primer/probe sequence, and instrument platform such as real-time quantitative PCR (qPCR) and digital PCR (dPCR) affect the accuracy and reliability of quantitative results. In this study, a plasmid DNA (pDNA) pBI121-screening, [...] Read more.
In polymerase chain reaction (PCR)-based nucleic acid quantification, the DNA template type, primer/probe sequence, and instrument platform such as real-time quantitative PCR (qPCR) and digital PCR (dPCR) affect the accuracy and reliability of quantitative results. In this study, a plasmid DNA (pDNA) pBI121-screening, genetically modified (GM) rice SDrice genomic DNA (gDNA), and GM rapeseed SDrape gDNA, all carrying the same 11 screening elements, were used to prepare samples of different levels of gDNA and pDNA in a non-GM gDNA background. The comparison of the dPCR assays targeting the 11 screening elements revealed that the primer/probe set is a key factor that affects the accuracy of dPCR quantification. The optimal PCR method for the 11 screening elements was screened out from among the validated qPCR methods. The accuracy of the qPCR quantification of the low-level pDNA and gDNA test samples was low when pDNA was used as a calibrator, whereas that of the dPCR quantification was high and not affected by variations in template type and detection target. The validated dPCR assays targeting one or two elements can be randomly selected to characterize multiple-target pDNA reference materials (RMs). Low-level pDNA RMs with certified values can be used as quality controls for dPCR assays to avoid significant bias in gDNA quantification. Full article
(This article belongs to the Section Agricultural Product Quality and Safety)
Show Figures

Figure 1

Figure 1
<p>Relative bias of the quantitative results of the test samples from the expected value by the quantitative polymerase chain reaction assays targeting T-NOS.</p>
Full article ">Figure 2
<p>Distribution of measurement average together with a standard deviation by ddPCR assays targeting 11 different elements for each test sample. 1–11 correspond to the dPCR assays of <span class="html-italic">Bar</span>, P-<span class="html-italic">CaMV</span>35S, P-<span class="html-italic">FMV</span>35S, T-NOS, <span class="html-italic">HPT</span>, <span class="html-italic">NPTII</span>, P-NOS, T-35S, <span class="html-italic">Pmi</span>, T-e9 and T-g7.</p>
Full article ">
16 pages, 1471 KiB  
Article
Do Mass Spectrometry-Derived Metabolomics Improve the Prediction of Pregnancy-Related Disorders? Findings from a UK Birth Cohort with Independent Validation
by Nancy McBride, Paul Yousefi, Ulla Sovio, Kurt Taylor, Yassaman Vafai, Tiffany Yang, Bo Hou, Matthew Suderman, Caroline Relton, Gordon C. S. Smith and Deborah A. Lawlor
Metabolites 2021, 11(8), 530; https://doi.org/10.3390/metabo11080530 - 10 Aug 2021
Cited by 9 | Viewed by 4514
Abstract
Many women who experience gestational diabetes (GDM), gestational hypertension (GHT), pre-eclampsia (PE), have a spontaneous preterm birth (sPTB) or have an offspring born small/large for gestational age (SGA/LGA) do not meet the criteria for high-risk pregnancies based upon certain maternal risk factors. Tools [...] Read more.
Many women who experience gestational diabetes (GDM), gestational hypertension (GHT), pre-eclampsia (PE), have a spontaneous preterm birth (sPTB) or have an offspring born small/large for gestational age (SGA/LGA) do not meet the criteria for high-risk pregnancies based upon certain maternal risk factors. Tools that better predict these outcomes are needed to tailor antenatal care to risk. Recent studies have suggested that metabolomics may improve the prediction of these pregnancy-related disorders. These have largely been based on targeted platforms or focused on a single pregnancy outcome. The aim of this study was to assess the predictive ability of an untargeted platform of over 700 metabolites to predict the above pregnancy-related disorders in two cohorts. We used data collected from women in the Born in Bradford study (BiB; two sub-samples, n = 2000 and n = 1000) and the Pregnancy Outcome Prediction study (POPs; n = 827) to train, test and validate prediction models for GDM, PE, GHT, SGA, LGA and sPTB. We compared the predictive performance of three models: (1) risk factors (maternal age, pregnancy smoking, BMI, ethnicity and parity) (2) mass spectrometry (MS)-derived metabolites (n = 718 quantified metabolites, collected at 26–28 weeks’ gestation) and (3) combined risk factors and metabolites. We used BiB for the training and testing of the models and POPs for independent validation. In both cohorts, discrimination for GDM, PE, LGA and SGA improved with the addition of metabolites to the risk factor model. The models’ area under the curve (AUC) were similar for both cohorts, with good discrimination for GDM (AUC (95% CI) BiB 0.76 (0.71, 0.81) and POPs 0.76 (0.72, 0.81)) and LGA (BiB 0.86 (0.80, 0.91) and POPs 0.76 (0.60, 0.92)). Discrimination was improved for the combined models (compared to the risk factors models) for PE and SGA, with modest discrimination in both studies (PE-BiB 0.68 (0.58, 0.78) and POPs 0.66 (0.60, 0.71); SGA-BiB 0.68 (0.63, 0.74) and POPs 0.64 (0.59, 0.69)). Prediction for sPTB was poor in BiB and POPs for all models. In BiB, calibration for the combined models was good for GDM, LGA and SGA. Retained predictors include 4-hydroxyglutamate for GDM, LGA and PE and glycerol for GDM and PE. MS-derived metabolomics combined with maternal risk factors improves the prediction of GDM, PE, LGA and SGA, with good discrimination for GDM and LGA. Validation across two very different cohorts supports further investigation on whether the metabolites reflect novel causal paths to GDM and LGA. Full article
(This article belongs to the Special Issue Metabolomics of Complex Traits II)
Show Figures

Figure 1

Figure 1
<p>Predictive discrimination of models for each outcome. AUC and 95% confidence intervals are shown for established risk factor prediction models (red), metabolite models (green) and combined risk factor and metabolite models (yellow) trained in the Born in Bradford 2000, tested in the Born in Bradford 1000 (triangles) and externally validated in the Pregnancy Outcome Prediction study (circles). POPs did not have sufficient data on gestational hypertension for validation. Abbreviations: BiB, Born in Bradford; POPs, Pregnancy Outcome Prediction study; GDM, gestational diabetes; GHT, gestational hypertension; PE, pre-eclampsia; SGA, small for gestational age; LGA, large for gestational age; sPTB, spontaneous preterm birth.</p>
Full article ">Figure 2
<p>Calibration slope for GDM combined model in BiB 1000 testing sample.</p>
Full article ">Figure 3
<p>Calibration slope for LGA combined model in BiB 1000 testing sample.</p>
Full article ">Figure 4
<p>Calibration slope for SGA combined model in BiB 1000 testing sample.</p>
Full article ">Figure 5
<p>Born in Bradford flowchart: the selection of participants for mass spectrometry metabolomic profiling in the Born in Bradford 1000 (<b>A</b>) and 2000 (<b>B</b>). Abbreviations: MS, mass spectrometry; BiB, Born in Bradford; GWAS, genome-wide association study; EDTA, ethylenediaminetetraacetic acid; HDP, hypertensive disorders of pregnancy; GD, gestational diabetes; GHT, gestational hypertension; PE, pre-eclampsia, PTB, preterm birth; CA, congenital anomaly; SB, still birth. (<b>C</b>) Illustrating the flow of participants into the Metabolon datasets ((<b>A</b>) BiB 1000, (<b>B</b>) BiB 2000 and (<b>C</b>) POPs (<span class="html-italic">n</span> = 923) cohorts). Abbreviations: MS, mass spectrometry; BiB, Born in Bradford; GWAS, genome-wide association study; EDTA, ethylenediaminetetraacetic acid; HDP, hypertensive disorders of pregnancy; GDM, gestational diabetes; GHT, gestational hypertension; PE, pre-eclampsia, PTB, preterm birth; sPTB, spontaneous preterm birth; CA, congenital anomaly; SB, still birth; FGR, fetal growth restriction; GA, gestational age. (<b>A</b>,<b>B</b>) were taken from our data note by Taylor et al. [<a href="#B33-metabolites-11-00530" class="html-bibr">33</a>] with permission.</p>
Full article ">
15 pages, 6040 KiB  
Review
Sub-THz and THz SiGe HBT Electrical Compact Modeling
by Bishwadeep Saha, Sebastien Fregonese, Anjan Chakravorty, Soumya Ranjan Panda and Thomas Zimmer
Electronics 2021, 10(12), 1397; https://doi.org/10.3390/electronics10121397 - 10 Jun 2021
Cited by 1 | Viewed by 2583
Abstract
From the perspectives of characterized data, calibrated TCAD simulations and compact modeling, we present a deeper investigation of the very high frequency behavior of state-of-the-art sub-THz silicon germanium heterojunction bipolar transistors (SiGe HBTs) fabricated with 55-nm BiCMOS process technology from STMicroelectronics. The TCAD [...] Read more.
From the perspectives of characterized data, calibrated TCAD simulations and compact modeling, we present a deeper investigation of the very high frequency behavior of state-of-the-art sub-THz silicon germanium heterojunction bipolar transistors (SiGe HBTs) fabricated with 55-nm BiCMOS process technology from STMicroelectronics. The TCAD simulation platform is appropriately calibrated with the measurements in order to aid the extraction of a few selected high-frequency (HF) parameters of the state-of-the-art compact model HICUM, which are otherwise difficult to extract from traditionally prepared test-structures. Physics-based strategies of extracting the HF parameters are elaborately presented followed by a sensitivity study to see the effects of the variations of HF parameters on certain frequency-dependent characteristics until 500 GHz. Finally, the deployed HICUM model is evaluated against the measured s-parameters of the investigated SiGe HBT until 500 GHz. Full article
(This article belongs to the Section Microelectronics)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Block diagram of the 140 GHz to 500 GHz measurement set-up, (<b>b</b>) photograph of the probe station for the 140 GHz to 500 GHz measurements, photo taken from [<a href="#B17-electronics-10-01397" class="html-bibr">17</a>].</p>
Full article ">Figure 2
<p>SiGe HBT device structure: (<b>a</b>) TEM image [<a href="#B21-electronics-10-01397" class="html-bibr">21</a>] and (<b>b</b>) structure made in sentaurus TCAD.</p>
Full article ">Figure 3
<p>(<b>a</b>) Gummel plots and (<b>b</b>) collector current dependent transit frequency characteristics for a 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V and <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> varying from 0.4 V to 1 V: comparison between measured data (“o” symbol), calibrated <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> simulation (“+” symbol) and quasi-static HICUM model (solid lines).</p>
Full article ">Figure 4
<p>Large-signal equivalent circuit of HICUM with an improved substrate network [<a href="#B18-electronics-10-01397" class="html-bibr">18</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Customized <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> structure (gray: silicite, blue: p-type poly-Si) and (<b>b</b>) RC equivalent circuit representing the circled region in (<b>a</b>) for the determination of base-emitter parasitic capacitance partitioning factor (<math display="inline"><semantics> <mrow> <mi>f</mi> <mi>b</mi> <mi>e</mi> <mi>p</mi> <mi>a</mi> <mi>r</mi> </mrow> </semantics></math>). Here, <span class="html-italic">B</span> is the external base node.</p>
Full article ">Figure 6
<p>Frequency dependent capacitance characteristic for <math display="inline"><semantics> <mrow> <mi>f</mi> <mi>b</mi> <mi>e</mi> <mi>p</mi> <mi>a</mi> <mi>r</mi> </mrow> </semantics></math> extraction following the customized <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> structure shown in <a href="#electronics-10-01397-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 7
<p>Customized <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> structure (<b>a</b>) and RC equivalent circuit (<b>b</b>) representing the left circled region for the determination of <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>j</mi> <mi>c</mi> <mi>x</mi> </mrow> </semantics></math>. <span class="html-italic">B</span> is the external base node.</p>
Full article ">Figure 8
<p>Customized <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> structure (<b>a</b>) and RC equivalent circuit (<b>b</b>) representing the left circled region for the determination of <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>j</mi> <mi>c</mi> <mi>x</mi> <mo>+</mo> <mi>c</mi> <mi>b</mi> <mi>c</mi> <mi>p</mi> <mi>a</mi> <mi>r</mi> </mrow> </semantics></math>. <span class="html-italic">B</span> is the external base node.</p>
Full article ">Figure 9
<p>Frequency dependent capacitance characteristic following the customized <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> structure shown in <a href="#electronics-10-01397-f007" class="html-fig">Figure 7</a>a (<b>a</b>) and <a href="#electronics-10-01397-f008" class="html-fig">Figure 8</a>a (<b>b</b>) for <math display="inline"><semantics> <mrow> <mi>f</mi> <mi>b</mi> <mi>c</mi> <mi>p</mi> <mi>a</mi> <mi>r</mi> </mrow> </semantics></math> extraction.</p>
Full article ">Figure 10
<p>(<b>a</b>) Turn-on and (<b>b</b>) turn-off characteristics for the collector current <math display="inline"><semantics> <mrow> <msub> <mi>i</mi> <mi>C</mi> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>: comparison between <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (symbols) and HICUM (solid line with <math display="inline"><semantics> <mrow> <mi>a</mi> <mi>l</mi> <mi>i</mi> <mi>t</mi> </mrow> </semantics></math> = 1) for a 0.09 μm × 4.8 μm SiGe HBT biased at constant <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>C</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.9 V (circles) and 0.85 V (plus). <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> <mi>s</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> have been shown as a dashed line to the right <span class="html-italic">Y</span>-axis.</p>
Full article ">Figure 11
<p>Variation of electric field (left axis) and carrier density (right axis) captured at 20 ps (solid line) and 26 ps (solid line with symbols). The bias voltages <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>C</mi> <mi>E</mi> </mrow> </msub> </semantics></math> are ramped up from 0 V to 0.9 V at 18 ps with a rise time of 2 ps. The value ‘0’ on the <span class="html-italic">x</span>-axis refers to the position where the poly-emitter and mono-emitter meet.</p>
Full article ">Figure 12
<p>Time dependent turn-on stored minority charge (in the emitter and base): comparison between <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (symbols) and HICUM (solid line with <math display="inline"><semantics> <mrow> <mi>a</mi> <mi>l</mi> <mi>q</mi> <mi>f</mi> </mrow> </semantics></math> = 1) for the 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.9 V (circle) and <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.85 V (plus), <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> <mi>s</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> have been shown as a dashed line to the right <span class="html-italic">Y</span>-axis.</p>
Full article ">Figure 13
<p>Sensitivity of <math display="inline"><semantics> <mrow> <mi>f</mi> <mi>b</mi> <mi>e</mi> <mi>p</mi> <mi>a</mi> <mi>r</mi> </mrow> </semantics></math> on frequency dependent <math display="inline"><semantics> <mrow> <mo>ℑ</mo> <msub> <mi>y</mi> <mn>11</mn> </msub> </mrow> </semantics></math> for the 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V with <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.85 V: comparison between <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (circles) and Hicum L2v2.4 (solid line).</p>
Full article ">Figure 14
<p>Sensitivity of <math display="inline"><semantics> <mrow> <mi>f</mi> <mi>b</mi> <mi>c</mi> <mi>p</mi> <mi>a</mi> <mi>r</mi> </mrow> </semantics></math> on frequency dependent <math display="inline"><semantics> <mrow> <mo>ℜ</mo> <msub> <mi>y</mi> <mn>12</mn> </msub> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <mo>ℑ</mo> <msub> <mi>y</mi> <mn>12</mn> </msub> </mrow> </semantics></math> (<b>b</b>) and <math display="inline"><semantics> <mrow> <mo>ℑ</mo> <msub> <mi>y</mi> <mn>11</mn> </msub> </mrow> </semantics></math> (<b>c</b>) for the 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V with <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.85 V: comparison between <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (circles) and Hicum L2v2.4 (solid line).</p>
Full article ">Figure 15
<p>Sensitivity of <math display="inline"><semantics> <mrow> <mi>a</mi> <mi>l</mi> <mi>i</mi> <mi>t</mi> </mrow> </semantics></math> on frequency dependent <math display="inline"><semantics> <mrow> <mo>ℜ</mo> <msub> <mi>y</mi> <mn>21</mn> </msub> </mrow> </semantics></math> (<b>a</b>) and <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>h</mi> <mi>a</mi> <mi>s</mi> <mi>e</mi> <msub> <mi>h</mi> <mn>21</mn> </msub> </mrow> </semantics></math> (<b>b</b>) for the 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V with <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.85 V: comparison between <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (circles) and Hicum L2v2.4 (solid line).</p>
Full article ">Figure 16
<p>Sensitivity of <math display="inline"><semantics> <mrow> <mi>a</mi> <mi>l</mi> <mi>q</mi> <mi>f</mi> </mrow> </semantics></math> on frequency dependent <math display="inline"><semantics> <mrow> <mo>ℜ</mo> <msub> <mi>y</mi> <mn>11</mn> </msub> </mrow> </semantics></math> (<b>a</b>) and <math display="inline"><semantics> <mrow> <mo>ℑ</mo> <msub> <mi>y</mi> <mn>11</mn> </msub> </mrow> </semantics></math> (<b>b</b>) for the 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V with <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.85 V: comparison between <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (circles) and Hicum L2v2.4 (solid line).</p>
Full article ">Figure 17
<p>Sensitivity of <math display="inline"><semantics> <mrow> <mi>f</mi> <mi>c</mi> <mi>r</mi> <mi>b</mi> <mi>i</mi> </mrow> </semantics></math> on frequency dependent <math display="inline"><semantics> <mrow> <mo>ℜ</mo> <msub> <mi>z</mi> <mn>11</mn> </msub> </mrow> </semantics></math> (<b>a</b>) and <math display="inline"><semantics> <mrow> <mo>ℑ</mo> <msub> <mi>z</mi> <mn>11</mn> </msub> </mrow> </semantics></math> (<b>b</b>) for the 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V with <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.85 V: comparison between <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (circles) and Hicum L2v2.4 (solid line).</p>
Full article ">Figure 18
<p>Frequency dependent magnitude of scattering parameters for the 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V with <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.8 V and 0.85 V: comparison between measured data (rectangles and circles), <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (triangle and cross) and HICUM (solid lines).</p>
Full article ">Figure 19
<p>Frequency dependent phase of scattering parameters for the 0.09 μm × 4.8 μm SiGe HBT biased at <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> </msub> </semantics></math> = 0 V with <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>B</mi> <mi>E</mi> </mrow> </msub> </semantics></math> = 0.8 V and 0.85 V: comparison between measured data (rectangles and circles), <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>C</mi> <mi>A</mi> <mi>D</mi> </mrow> </semantics></math> (triangle and cross) and HICUM (solid lines).</p>
Full article ">
30 pages, 11570 KiB  
Article
Coastal Sand Dunes Monitoring by Low Vegetation Cover Classification and Digital Elevation Model Improvement Using Synchronized Hyperspectral and Full-Waveform LiDAR Remote Sensing
by Giovanni Frati, Patrick Launeau, Marc Robin, Manuel Giraud, Martin Juigner, Françoise Debaine and Cyril Michon
Remote Sens. 2021, 13(1), 29; https://doi.org/10.3390/rs13010029 - 23 Dec 2020
Cited by 7 | Viewed by 4155
Abstract
Due to the coastal morphodynamic being impacted by climate change there is a need for systematic and large-scale monitoring. The monitoring of sandy dunes in Pays-de-la-Loire (France) requires a simultaneous mapping of (i) its morphology, allowing to assess the sedimentary stocks and (ii) [...] Read more.
Due to the coastal morphodynamic being impacted by climate change there is a need for systematic and large-scale monitoring. The monitoring of sandy dunes in Pays-de-la-Loire (France) requires a simultaneous mapping of (i) its morphology, allowing to assess the sedimentary stocks and (ii) its low vegetation cover, which constitutes a significant proxy of the dune dynamics. The synchronization of hyperspectral imaging (HSI) with full-waveform (FWF) LiDAR is possible with an airborne platform. For a more intimate combination, we aligned the 1064 nm laser beam of a bi-spectral Titan FWF LiDAR with 401 bands and the 15 cm range resolution on the Hyspex VNIR camera with 160 bands and a 4.2 nm spectral resolution, making both types of data follow the same emergence angle. A ray tracing procedure permits to associate the data while keeping the acquisition angles. Stacking multiple shifted FWFs, which are linked to the same pixel, enables reaching a 5 cm range resolution grid. The objectives are (i) to improve the accuracy of the digital terrain models (DTM) obtained from an FWF analysis by calibrating it on dGPS field measurements and correcting it from local deviations induced by vegetation and (ii) in combination with airborne reflectances obtained with PARGE and ATCOR-4 corrections, to implement a supervised hierarchic classification of the main foredune vegetation proxies independently of the acquisition year and the physiological state. The normalization of the FWF LiDAR range to a dry sand reference waveform and the centering on their top canopy echoes allows to isolate Ammophilia arenaria from other vegetation types using two FWF indices, without confusion with slope effects. Fourteen HSI reflectance indices and 19 HSI Spectral Angle Mapping (SAM) indices based on 2017 spectral field measurements performed with the same Hyspex VNIR camera were stacked with both FWF indices into a single co-image for each acquisition year. A simple straightforward hierarchical classification of all 35 pre-classified co-image bands was successfully applied along 20 km, out of the 250 km of coastline acquired from 2017 to 2019, prefiguring its systematic application to the whole 250 km every year. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A typical western French sandy coast profile. The morphological terminology is written in blue and typical vegetation cover is written in green. Note that the distance scale from the foreshore is only indicative and can vary from one area to another.</p>
Full article ">Figure 2
<p>Location of the study areas: Noirmoutier Island and the north part of Pays-de-Monts, in western France. Here are noted: the working area “Tresson” and the validation areas “Barbâtre” (available in the appendix) and the north part of “Pays-de-Monts”. (<b>a</b>) Represents the dGPS field sampling location of the working area, (<b>b</b>) represents Dame-de-Monts subsidiary validation area 1, with b1 and b2 corresponding to the dGPS field sampling location of the validation area and (<b>c</b>) Dame-de-Monts subsidiary validation area 2. Image from <a href="http://geoportail.gouv.fr" target="_blank">geoportail.gouv.fr</a>.</p>
Full article ">Figure 3
<p><span class="html-italic">Ammophila arenaria</span> mean spectra. (<b>a</b>) The light green curve represents the <span class="html-italic">Ammophilia arenaria</span> mean spectrum, acquired from a 1500 m distance with a 90<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>± 8.5<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math> angle, whereas the yellow curve represents a dry sand mean spectrum, acquired with the same camera; (<b>b</b>) the dark green curve represents the <span class="html-italic">Ammophila arenaria</span> field sampled mean spectrum, acquired at a distance of 1 m, following a 50<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>± 5<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math> angle, whereas the yellow curve represents the dry sand mean spectrum acquired using an ASD field spectrometer.</p>
Full article ">Figure 4
<p>Dune vegetation spectral range. (<b>a</b>) Represents the field-acquired mean spectrum of the 17 main plants of the scene; (<b>b</b>) represents mean spectra (<b>thick lines</b>) plus or minus one standard deviation (<b>colored areas</b>) for <span class="html-italic">Ammophila arenaria</span> (<b>green</b>) and <span class="html-italic">Elymus farctus</span> (<b>pink</b>). This shows that the major part of their spectral variation ranges overlaps, making a multi-temporal systematic discrimination based on a spectral argument almost impossible.</p>
Full article ">Figure 5
<p>(x, y) plane of the scan-line geometry giving an along-track row spacing of 0.5 m at a nadir to 1 m on the side strip image with a minimum along-track point density of 1 m. At the nadir, the 0.50 m footprint allows placing two footprints in a pixel in the along-track direction. However, at the edges, the PRF allows obtaining at least one footprint in the middle of a pixel.</p>
Full article ">Figure 6
<p>Procedures of FWF signal quality improvement: (<b>a</b>) raw 15 cm resolution FWF of beach dry sand (<b>grey</b>), <span class="html-italic">Ammophila arenaria</span> (<b>magenta</b>) and dune slope (<b>blue</b>); (<b>b</b>) FWF low-pass filtering with red green and blue channel selection of (<b>c</b>) the corresponding color composite image; (<b>d</b>) same raw FWF signals with a median filter discussed in the text; (<b>e</b>) FWF median results with low-pass and (<b>f</b>) its color composite display; (<b>g</b>) sketch of the 15 to 5 cm procedure discussed in the text; (<b>h</b>) raw 5 cm resolution FWF; (<b>i</b>) with low-pass and (<b>j</b>) its corresponding color composite display; (<b>k</b>) combination of 5 cm resolution and median filtering of the FWF; (<b>l</b>) with low-pass filter and (<b>m</b>) its corresponding color composite display.</p>
Full article ">Figure 7
<p>Shape analysis methodology. (<b>a</b>–<b>c</b>) represent the dNCCFWF signal, whereas (<b>d</b>–<b>f</b>) represent the deviation from the dry sandy surface reference used for surface identification, as it emphasizes the shape variations. The black curve represents the dry sandy reference signal and the red curve represents the signal of the considered object (<span class="html-italic">Ammophila arenaria</span> for (<b>a</b>,<b>d</b>), tree for (<b>b</b>,<b>e</b>), and slope effect for (<b>c</b>,<b>f</b>)). The colors represent the main shape variation of the considered signal from the dry sand reference: blue is the upward damping, corresponding to the FWF marram index measured area, yellow is the downward damping, corresponding to the FWF down index, and green is the relative intensity difference at the position of the center of the echo.</p>
Full article ">Figure 8
<p>Decision tree used for the classification. The ellipses are the resulting vegetation groups with the color code used in Figures 11, 13a,b, 14a, 15a,b and 16a,b.</p>
Full article ">Figure 9
<p>Deviation between the dGPS data and (<b>a</b>) the d3NCFWF’s first echo and (<b>b</b>) DTM data for 2017 (<b>green points</b>), 2018 (<b>red points</b>), and 2019 (<b>blue points</b>), as a function of the field sampling point number. The colored areas represent the different types of environments: crest sand paths (<b>brown</b>), beach top (<b>yellow</b>) and <span class="html-italic">Ammophila arenaria</span> (<b>green</b>).</p>
Full article ">Figure 10
<p>Deviation between the dGPS data and corrected data: (<b>a</b>) d3NCFWF’s first echo and (<b>b</b>) DTM data for 2017 (<b>green points</b>), 2018 (<b>red points</b>), and 2019 (<b>blue points</b>), as a function of the field sampling point number. Years’ and environments’ color codes are identical to <a href="#remotesensing-13-00029-f009" class="html-fig">Figure 9</a>.</p>
Full article ">Figure 11
<p>dGPS field measurements versus classified foredune proxies. The extension of the classes 3 to 7 is shown in magenta, containing <span class="html-italic">Ammophila arenaria</span> classified using the FWF argument; the extension of classes 20 to 24 is shown in green, containing <span class="html-italic">Elymus farctus</span>. Sparse vegetation over dry sand, which corresponds to class 17, is represented by light yellow. Yellow points represent dGPS points acquired over the white dune crest, and red points correspond to those sampled on <span class="html-italic">Ammophila arenaria</span>, making 8 different numerated profiles oriented perpendicularly to the coastline; blue points correspond to other dGPS points.</p>
Full article ">Figure 12
<p>LiDAR elevation profiles of the fourth profile of <a href="#remotesensing-13-00029-f011" class="html-fig">Figure 11</a>. The black line represents field dGPS, the red line is the 2019 discrete echo, the blue one is the 2019 FWF single echo without <span class="html-italic">Ammophila arenaria</span> correction, and the green one is the 2019 FWF single echo with <span class="html-italic">Ammophila arenaria</span> correction. The grey line represents IGN Litto3D DTM.</p>
Full article ">Figure 13
<p>Tresson training area of d3NCFWF 1st echoes DSM with marram height correction in (<b>a</b>) 2018 and (<b>b</b>) 2019 with maps of <span class="html-italic">Ammophila arenaria</span> (<b>magenta</b>), <span class="html-italic">Elymus farctus</span> (<b>green</b>), and sparse vegetation over dry sand (<b>light yellow</b>). Light blue dotted lines represent the position of the profile represented in (<b>c</b>), which are corrected FWF last echoes of 2018 (<b>dark red</b>) and 2019 (<b>light red</b>), discrete echoes DTM of 2018 (<b>dark blue</b>) and 2019 (<b>light blue</b>), and IGN Litto3D data of 2010 (<b>grey</b>) and 2019 (<b>black</b>).</p>
Full article ">Figure 14
<p>Pays-de-Monts validation area (b in <a href="#remotesensing-13-00029-f002" class="html-fig">Figure 2</a>) of true color hyperspectral display with maps of <span class="html-italic">Ammophila arenaria</span> (<b>magenta</b>), <span class="html-italic">Elymus farctus</span> (<b>green</b>), and sparse vegetation over dry sand (<b>light yellow</b>). Locations of dGPS foot points of <span class="html-italic">Ammophila arenaria</span> are shown in red, <span class="html-italic">Elymus farctus</span> in addition to those classified as sparse vegetation over dry sand are shown in blue, and the yellow ones are beach-top samples; (<b>a</b>) displays the 6 first profiles (b1 <a href="#remotesensing-13-00029-f002" class="html-fig">Figure 2</a>) cutting <span class="html-italic">Ammophila arenaria</span> polygons larger than 3 pixels; (<b>b</b>) displays the last 4 profiles (b2 <a href="#remotesensing-13-00029-f002" class="html-fig">Figure 2</a>) cutting narrower <span class="html-italic">Ammophila arenaria</span> polygons.</p>
Full article ">Figure 15
<p>Notre-Dame-de-Monts validation area with the same caption as <a href="#remotesensing-13-00029-f013" class="html-fig">Figure 13</a>. See location b in <a href="#remotesensing-13-00029-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 16
<p>Barre-de-Monts validation area with the same caption as <a href="#remotesensing-13-00029-f013" class="html-fig">Figure 13</a>. See location c in <a href="#remotesensing-13-00029-f002" class="html-fig">Figure 2</a>.</p>
Full article ">
20 pages, 4083 KiB  
Article
Study on a Dynamic Numerical Model of an Underground Air Tunnel System for Cooling Applications—Experimental Validation and Multidimensional Parametrical Analysis
by Liang Tang, Zhengxuan Liu, Yuekuan Zhou, Di Qin and Guoqiang Zhang
Energies 2020, 13(5), 1236; https://doi.org/10.3390/en13051236 - 6 Mar 2020
Cited by 11 | Viewed by 2520
Abstract
The underground air tunnel system shows promising potentials for reducing energy consumption of buildings and for improving indoor thermal comfort, whereas the existing dynamic models using the computational fluid dynamic (CFD) method show computational complexity and are user-unfriendly for parametrical analysis. In this [...] Read more.
The underground air tunnel system shows promising potentials for reducing energy consumption of buildings and for improving indoor thermal comfort, whereas the existing dynamic models using the computational fluid dynamic (CFD) method show computational complexity and are user-unfriendly for parametrical analysis. In this study, a dynamic numerical model was developed with the on-site experimental calibration. Compared to the traditional CFD method with high computational complexity, the mathematical model on the MATLAB/SIMULINK platform is time-saving in terms of the real-time thermal performance prediction. The experimental validation results indicated that the maximum absolute relative deviation was 3.18% between the model-driven results and the data from the on-site experiments. Parametrical analysis results indicated that, with the increase of the tube length, the outlet temperature decreases with an increase of the cooling capacity whereas the increasing/decreasing magnitude slows down. In addition, the system performance is independent on the tube materials. Furthermore, the outlet air temperature and cooling capacity are dependent on the tube diameter and air velocity, i.e., a larger tube diameter and a higher air velocity are more suitable to improve the system’s cooling capacity, and a smaller tube diameter and a lower air velocity will produce a more stable and lower outlet temperature. Further studies need to be conducted for the trade-off solutions between air velocity and tube diameter for the bi-criteria performance enhancement between outlet temperature and cooling capacity. This study proposed an experimentally validated mathematical model to accurately predict the thermal performance of the underground air tunnel system with high computational efficiency, which can provide technical guidance to multi-combined solutions through geometrical designs and operating parameters for the optimal design and robust operation. Full article
(This article belongs to the Special Issue Optimization of Solar Thermal Systems for Buildings)
Show Figures

Figure 1

Figure 1
<p>The parameter calculation of system’s thermal performance through the MATLAB/SIMULINK platform using S-Function.</p>
Full article ">Figure 2
<p>The divided volume elements of the underground air tunnel system.</p>
Full article ">Figure 3
<p>Demonstration of the nodes for the air, tube wall, and soil layer.</p>
Full article ">Figure 4
<p>Schematic diagram of underground air tunnel (UAT) system.</p>
Full article ">Figure 5
<p>The comparison between experiment and simulation data for outlet air temperature and cooling capacity.</p>
Full article ">Figure 6
<p>The effect of different tube lengths on the cooling capacity and the outlet temperature.</p>
Full article ">Figure 7
<p>The effect of different tube conductivities on the outlet temperature and cooling capacity.</p>
Full article ">Figure 8
<p>Evolution of outlet temperature and cooling capacity with respect to tube diameters.</p>
Full article ">Figure 9
<p>The effect of different air velocities on the outlet temperature and cooling capacity.</p>
Full article ">
20 pages, 3665 KiB  
Article
O&M Models for Ocean Energy Converters: Calibrating through Real Sea Data
by Tianna Bloise Thomaz, David Crooks, Encarni Medina-Lopez, Leonore van Velzen, Henry Jeffrey, Joseba Lopez Mendia, Raul Rodriguez Arias and Pablo Ruiz Minguela
Energies 2019, 12(13), 2475; https://doi.org/10.3390/en12132475 - 27 Jun 2019
Cited by 10 | Viewed by 3942
Abstract
Of the cost centres that combine to result in Levelised Cost of Energy (LCOE), O&M costs play a significant part. Several developers have calculated component costs, demonstrating how they can become commercially competitive with other forms of renewable energy. However, there are uncertainties [...] Read more.
Of the cost centres that combine to result in Levelised Cost of Energy (LCOE), O&M costs play a significant part. Several developers have calculated component costs, demonstrating how they can become commercially competitive with other forms of renewable energy. However, there are uncertainties relating to the O&M figures that can only be reduced through lessons learned at sea. This work presents an O&M model calibrated with data from real sea experience of a wave energy device deployed at the Biscay Marine energy Platform (BiMEP): the OPERA O&M Model. Two additional case studies, utilising two other O&M calculation methodologies, are presented for comparison with the OPERA O&M Model. The second case study assumes the inexistence of an O&M model, utilising a Simplified Approach. The third case study applies DTOcean’s (a design tool for ocean energy arrays) O&M module. The results illustrate the potential advantages of utilising real sea data for the calibration and development of an O&M model. The Simplified Approach was observed to overestimate LCOE when compared to the OPERA O&M Model. This work also shows that O&M models can be used for the definition of optimal maintenance plans to assist with OPEX reduction. Full article
(This article belongs to the Section A3: Wind, Wave and Tidal Energy)
Show Figures

Figure 1

Figure 1
<p>Relationship between expected O&amp;M costs and Maintenance Effort. Green line show that expected failure costs reduce with the Maintenance Effort, whereas the blue line shows that the maintenance costs increase with the rise of Maintenance Effort. The sign (x) shows the optimal strategy point when the total costs are minimum [<a href="#B11-energies-12-02475" class="html-bibr">11</a>].</p>
Full article ">Figure 2
<p>OPEX model architecture.</p>
Full article ">Figure 3
<p>Maintenance actions of <span class="html-italic">OPERA O&amp;M Model</span> [<a href="#B37-energies-12-02475" class="html-bibr">37</a>].</p>
Full article ">Figure 4
<p>MARMOK-A-5 device. Device deployed at BiMEP (left-side) and device’s main components (right-side).</p>
Full article ">Figure 5
<p>Maintenance strategy of DTOcean [<a href="#B45-energies-12-02475" class="html-bibr">45</a>].</p>
Full article ">Figure 6
<p>Handling of automated inspection in case of condition-based maintenance [<a href="#B44-energies-12-02475" class="html-bibr">44</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>): Array of Three Turbines and bathymetry of the lease area; (<b>b</b>): Nova M100 tidal turbine.</p>
Full article ">Figure 8
<p>Normalised LCOE for the <span class="html-italic">OPERA O&amp;M Model</span> and <span class="html-italic">Simplified Approach</span>. Error bars show the uncertainty range between results.</p>
Full article ">Figure 9
<p>Normalised OPEX, installation and decommissioning costs calculated using the <span class="html-italic">OPERA O&amp;M Model</span> and the <span class="html-italic">Simplified Approach</span>.</p>
Full article ">Figure 10
<p>Variation of the number of operations a year for the <span class="html-italic">OPERA O&amp;M Model</span>. Graph shows the number of preventive and corrective activities, as well as the total number of operations for the Array of 18 MW (72 devices).</p>
Full article ">Figure 11
<p>Normalised OPEX (N<sub>OPEX</sub>) for different Maintenance Efforts (M<sub>e</sub>) for the <span class="html-italic">OPERA O&amp;M Model</span>, presented in (<b>upper</b>) line graph and (<b>lower</b>) bar graph. Error bar shows the uncertainty range between results with a probability of 68% of the results.</p>
Full article ">Figure 12
<p>Normalised OPEX (N<sub>OPEX</sub>) for different Maintenance Efforts (M<sub>e</sub>) calculated by the <span class="html-italic">DTOcean O&amp;M Module</span> case study. Error bars shows the uncertainty range between results with a probability of 68% of the results.</p>
Full article ">
392 KiB  
Article
Fatigue Reliability and Calibration of Fatigue Design Factors for Offshore Wind Turbines
by Sergio Márquez-Domínguez and John D. Sørensen
Energies 2012, 5(6), 1816-1834; https://doi.org/10.3390/en5061816 - 15 Jun 2012
Cited by 67 | Viewed by 8228 | Correction
Abstract
Consequences of failure of offshore wind turbines (OWTs) is in general lower than consequences of failure of, e.g., oil & gas platforms. It is reasonable that lower fatigue design factors can be applied for fatigue design of OWTs when compared to other fixed [...] Read more.
Consequences of failure of offshore wind turbines (OWTs) is in general lower than consequences of failure of, e.g., oil & gas platforms. It is reasonable that lower fatigue design factors can be applied for fatigue design of OWTs when compared to other fixed offshore structures. Calibration of appropriate partial safety factors/Fatigue Design Factors (FDF) for steel substructures for OWTs is the scope of this paper. A reliability-based approach is used and a probabilistic model has been developed, where design and limit state equations are established for fatigue failure. The strength and load uncertainties are described by stochastic variables. SN and fracture mechanics approaches are considered for to model the fatigue life. Further, both linear and bi-linear SN-curves are formulated and various approximations are investigated. The acceptable reliability level for fatigue failure of OWTs is discussed and results are presented for calibrated optimal fatigue design factors. Further, the influence of inspections is considered in order to extend and maintain a given target safety level. Full article
(This article belongs to the Special Issue Wind Turbines)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Representative model and linearized model for number of stress ranges as function of stress ranges (normalized).</p>
Full article ">Figure 2
<p>Illustration of updating of the reliability of a critical detail/component by inspection of the same component and by inspection of another component in the same wind turbine or in another wind turbine in a wind farm.</p>
Full article ">Figure 3
<p>Illustration of inspection plan with equidistant inspections.</p>
Full article ">Figure 4
<p>Illustration of inspection plan where inspections are performed when the annual probability of failure exceeds the maximum acceptable annual probability of failure.</p>
Full article ">Figure 5
<p>Annual reliability index as function of <span class="html-italic">FDF</span> for <span class="html-italic">T<sub>L</sub></span> = 25 years and base values of coefficients of variation in <a href="#energies-05-01816-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 6
<p>Cumulative reliability index as function of <span class="html-italic">FDF</span> for life <span class="html-italic">T<sub>L</sub></span> = 25 years and base values of coefficients of variation in <a href="#energies-05-01816-t001" class="html-table">Table 1</a>.</p>
Full article ">
Back to TopTop