[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

Article Types

Countries / Regions

Search Results (60)

Search Parameters:
Keywords = Pearson VI

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 5050 KiB  
Article
Yield Prediction Models for Rice Varieties Using UAV Multispectral Imagery in the Amazon Lowlands of Peru
by Diego Goigochea-Pinchi, Maikol Justino-Pinedo, Sergio S. Vega-Herrera, Martín Sanchez-Ojanasta, Roiser H. Lobato-Galvez, Manuel D. Santillan-Gonzales, Jorge J. Ganoza-Roncal, Zoila L. Ore-Aquino and Alex I. Agurto-Piñarreta
AgriEngineering 2024, 6(3), 2955-2969; https://doi.org/10.3390/agriengineering6030170 - 20 Aug 2024
Viewed by 1114
Abstract
Rice is cataloged as one of the most widely cultivated crops globally, providing food for a large proportion of the global population. Integrating Geographic Information Systems (GISs), such as unmanned aerial vehicles (UAVs), into agricultural practices offers numerous benefits. UAVs, equipped with imaging [...] Read more.
Rice is cataloged as one of the most widely cultivated crops globally, providing food for a large proportion of the global population. Integrating Geographic Information Systems (GISs), such as unmanned aerial vehicles (UAVs), into agricultural practices offers numerous benefits. UAVs, equipped with imaging sensors and geolocation technology, enable precise crop monitoring and management, enhancing yield and efficiency. However, Peru lacks sufficient experience with the application of these technologies, making them somewhat unfamiliar in the context of modern agriculture. In this study, we conducted experiments involving four distinct rice varieties (n = 24) at various stages of growth to predict yield using vegetation indices (VIs). A total of nine VIs (NDVI, GNDVI, ReCL, CIgreen, MCARI, SAVI, CVI, LCI, and EVI) were assessed across four dates: 88, 103, 116, and 130 days after sowing (DAS). Pearson correlation analysis, principal component analysis (PCA), and multiple linear regression were used to build prediction models. The results showed a general prediction model (including all the varieties) with the best performance at 130 days after sowing (DAS) using NDVI, EVI, and SAVI, with a coefficient of determination (adjusted-R2 = 0.43). The prediction models by variety showed the best performance for Esperanza at 88 DAS (adjusted-R2 = 0.94) using EVI as the vegetation index. The other varieties showed their best performance using different indices at different times: Capirona (LCI and CIgreen, 130 DAS, adjusted-R2 = 0.62); Conquista Certificada (MCARI, 116 DAS, R2 = 0.52); and Conquista Registrada (CVI and LCI, 116 DAS, adjusted-R2 = 0.79). These results provide critical information for optimizing rice crop management and support the use of unmanned aerial vehicles (UAVs) to inform timely decision making and mitigate yield losses in Peruvian agriculture. Full article
(This article belongs to the Section Remote Sensing in Agriculture)
Show Figures

Figure 1

Figure 1
<p>Location of the experimental area in the El Porvenir Experimental Center in San Martin, Peru.</p>
Full article ">Figure 2
<p>Flowchart of the methodology framework.</p>
Full article ">Figure 3
<p>Comparison of the yield of the different rice cultivars. Different letters indicate significant differences according to the Tukey test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Evaluation of chlorophyll content for all rice cultivars in different DAS.</p>
Full article ">Figure 5
<p>Normalized difference vegetative index (NDVI) for four dates evaluated in rice crop.</p>
Full article ">Figure 6
<p>Vegetation indices at different evaluated days. (<b>a</b>) NDVI, (<b>b</b>) CIGREEN, (<b>c</b>) CVI, (<b>d</b>) EVI, (<b>e</b>) GNDVI, (<b>f</b>) LCI, (<b>g</b>) MCARI, (<b>h</b>) RECL, and (<b>i</b>) SAVI estimated by multispectral images from UAV for all rice cultivars.</p>
Full article ">Figure 7
<p>PCA for VIs during the evaluation days: (<b>a</b>) PCA for all the indices and DAS; (<b>b</b>) PCA for the indices and group of rice cultivars at 88 DAS; (<b>c</b>) PCA for the indices and group of rice cultivars at 103 DAS; (<b>d</b>) PCA for the indices and group of rice cultivars at 116 DAS; (<b>e</b>) PCA for the indices and group of rice cultivars at 130 DAS.</p>
Full article ">Figure 8
<p>Yield evaluated in the field and prediction models for the yield of four rice varieties with multiple linear regression, (<b>a</b>) 130 DAS, (<b>b</b>) 88 DAS, and (<b>c</b>,<b>d</b>) 116 DAS. ns. <span class="html-italic">p</span>-value &gt; 0.05, * <span class="html-italic">p</span>-value &lt; 0.05, and *** <span class="html-italic">p</span>-value &lt; 0.001.</p>
Full article ">
19 pages, 5408 KiB  
Article
Can Multi-Temporal Vegetation Indices and Machine Learning Algorithms Be Used for Estimation of Groundnut Canopy State Variables?
by Shaikh Yassir Yousouf Jewan, Ajit Singh, Lawal Billa, Debbie Sparkes, Erik Murchie, Deepak Gautam, Alessia Cogato and Vinay Pagay
Horticulturae 2024, 10(7), 748; https://doi.org/10.3390/horticulturae10070748 - 16 Jul 2024
Viewed by 887
Abstract
The objective of this research was to assess the feasibility of remote sensing (RS) technology, specifically an unmanned aerial system (UAS), to estimate Bambara groundnut canopy state variables including leaf area index (LAI), canopy chlorophyll content (CCC), aboveground biomass (AGB), and fractional vegetation [...] Read more.
The objective of this research was to assess the feasibility of remote sensing (RS) technology, specifically an unmanned aerial system (UAS), to estimate Bambara groundnut canopy state variables including leaf area index (LAI), canopy chlorophyll content (CCC), aboveground biomass (AGB), and fractional vegetation cover (FVC). RS and ground data were acquired during Malaysia’s 2018/2019 Bambara groundnut growing season at six phenological stages; vegetative, flowering, podding, podfilling, maturity, and senescence. Five vegetation indices (VIs) were determined from the RS data, resulting in single-stage VIs and cumulative VIs (∑VIs). Pearson’s correlation was used to investigate the relationship between canopy state variables and single stage VIs and ∑VIs over several stages. Linear parametric and non-linear non-parametric machine learning (ML) regressions including CatBoost Regressor (CBR), Random Forest Regressor (RFR), AdaBoost Regressor (ABR), Huber Regressor (HR), Multiple Linear Regressor (MLR), Theil-Sen Regressor (TSR), Partial Least Squares Regressor (PLSR), and Ridge Regressor (RR) were used to estimate canopy state variables using VIs/∑VIs as input. The best single-stage correlations between canopy state variables and VIs were observed at flowering (r > 0.50 in most cases). Moreover, ∑VIs acquired from vegetative to senescence stage had the strongest correlation with all measured canopy state variables (r > 0.70 in most cases). In estimating AGB, MLR achieved the best testing performance (R2 = 0.77, RMSE = 0.30). For CCC, RFR excelled with R2 of 0.85 and RMSE of 2.88. Most models performed well in FVC estimation with testing R2 of 0.98–0.99 and low RMSE. For LAI, MLR stood out in testing with R2 of 0.74, and RMSE of 0.63. Results demonstrate the UAS-based RS technology potential for estimating Bambara groundnut canopy variables. Full article
(This article belongs to the Special Issue Smart Horticulture: Latest Advances and Prospects)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study area at Field Research Centre of Crops for the Future. The experimental layout of plots was digitised on an image acquired with the integrated DJI Phantom 4 Pro camera at a height of 10 m on flowering stage. B1G1R1 means plot is in block 1; genotype is genotype 1, and replicate is the first replicate.</p>
Full article ">Figure 2
<p>This depicts the environmental parameters and irrigation levels throughout the 2018 cultivation period spanning from May to September. The arrows indicate distinct growth phases in the life cycle of Bambara groundnut. These stages include SOW (sowing), VEG (vegetative), FLO (flowering), POD (podding), PF (pod filling), MAT (maturity), SEN (senescence), and HAR (harvest). The asterisk (*) denotes data collection time points.</p>
Full article ">Figure 3
<p>Workflow for modelling Bambara groundnut canopy state variables.</p>
Full article ">Figure 4
<p>Correlation coefficients between crop variables and VIs were plotted across various growth stages: VEG (vegetative), FLO (flowering), POD (podding), PF (pod filling), MAT (maturity), and SEN (senescence). * indicates statistical significance at <span class="html-italic">p</span> &lt; 0.05, ** indicates statistical significance at <span class="html-italic">p</span> &lt; 0.01 and ns means non-significant.</p>
Full article ">Figure 5
<p>Plot comparing predicted versus observed values using the top-performing models for each canopy state variable. The solid black line represents the best-fit line, while the dashed grey line corresponds to the line y = x.</p>
Full article ">Figure 6
<p>Predictor importance plots ranking ΣVIs for estimating Bambara groundnut canopy state variables, where higher feature importance values indicate greater importance in the model.</p>
Full article ">
17 pages, 3940 KiB  
Article
Estimating Leaf Chlorophyll Content of Winter Wheat from UAV Multispectral Images Using Machine Learning Algorithms under Different Species, Growth Stages, and Nitrogen Stress Conditions
by Liyuan Zhang, Aichen Wang, Huiyue Zhang, Qingzhen Zhu, Huihui Zhang, Weihong Sun and Yaxiao Niu
Agriculture 2024, 14(7), 1064; https://doi.org/10.3390/agriculture14071064 - 1 Jul 2024
Cited by 1 | Viewed by 796
Abstract
The rapid and accurate estimation of leaf chlorophyll content (LCC), an important indicator of crop photosynthetic capacity and nutritional status, is of great significance for precise nitrogen fertilization management. To explore the existence of a versatile regression model that can be successfully used [...] Read more.
The rapid and accurate estimation of leaf chlorophyll content (LCC), an important indicator of crop photosynthetic capacity and nutritional status, is of great significance for precise nitrogen fertilization management. To explore the existence of a versatile regression model that can be successfully used to estimate the LCC for different varieties under different growth stages and nitrogen stress conditions, a study was conducted in 2023 across the growing season for winter wheat with five species and five nitrogen application levels. Two machine learning regression algorithms, support vector machine (SVM) and random forest (RF), were used to establish the bridge between UAV-derived multispectral vegetation indices and ground truth LCC (relative chlorophyll content, SPAD), taking the multivariate linear regression (MLR) algorithm as a reference. The results show that the visible atmospherically resistant index, vegetative index, and normalized difference vegetation index had the highest correlation with ground truth LCC, with a Pearson’s correlation coefficient of 0.95. All three regression algorithms (MLR, RF, and SVM) performed well on the training dataset (R2: 0.932–0.944, RMSE: 3.96 to 4.37), but performed differently on validation datasets with different growth stages, species, and nitrogen application levels. Compared to winter wheat species and nitrogen application levels, the growth stages had the greatest influence on the generalization ability of LCC estimation models, especially for the dough stage. At the dough stage, compared to MLR and RF, SVM performed best, with R2 increasing by 0.27 and 0.10, respectively, and RMSE decreasing by 1.13 and 0.46, respectively. Overall, this study demonstrated that the combination of UAV-derived multispectral VIs and the SVM regression algorithm could be successfully applied to map the LCC of winter wheat for different species, growth stages, and nitrogen stress conditions. Ultimately, this research is significant as it shows the successful application of UAV data for mapping the LCC of winter wheat across diverse conditions, offering valuable insights for precision nitrogen fertilization management. Full article
(This article belongs to the Section Digital Agriculture)
Show Figures

Figure 1

Figure 1
<p>Geographical location of the experimental site and overview of winter wheat experimental design. R1–R5 present repetition areas of 1 to 5, N0–N4 represent nitrogen levels of 0 to 4, and S1–S5 represent winter wheat species of 1 to 5. GCP is the abbreviation of ground control point.</p>
Full article ">Figure 2
<p>The flowsheet for mapping LCC of winter wheat based on UAV-derived vegetation indices (VIs). (<b>a</b>) The establishment of LCC estimation models; (<b>b</b>) the evaluation of three LCC estimation models; (<b>c</b>) LCC maps derived based on the optimal model and UAV-derived VIs. MLR, RF, and SVM represent multivariate linear, random forest, and support vector machine regression algorithm, respectively.</p>
Full article ">Figure 3
<p>Temporal variations in LCC of winter wheat for different nitrogen application levels and winter wheat species. “N_levels” represents the five nitrogen application levels at 0, 90, 180, 270, and 360 kg/ha.</p>
Full article ">Figure 4
<p>Pearson’s correlation coefficient (r) between LCC and each individual UAV-derived VI. (<b>a</b>) VIs with only the visible bands; (<b>b</b>) VIs with the NIR band but excluding the red-edge band; (<b>c</b>) VIs including the red-edge band.</p>
Full article ">Figure 5
<p>The distributions of (<b>a</b>) coefficient of determination (R<sup>2</sup>) and (<b>b</b>) root mean square error (RMSE) derived using five-fold cross validation method for three machine learning algorithms.</p>
Full article ">Figure 6
<p>R<sup>2</sup> and RMSE values obtained for three regression models based on validation datasets, which were divided based on the growth stages of winter wheat. The dashed line is the corresponding estimation accuracy observed in the training process.</p>
Full article ">Figure 7
<p>R<sup>2</sup> and RMSE values obtained for three regression models based on validation datasets, which were divided based on the species of winter wheat. The dashed line is the corresponding estimation accuracy observed in the training process.</p>
Full article ">Figure 8
<p>R<sup>2</sup> and RMSE values obtained for three regression models based on validation datasets, which were divided based on the nitrogen application levels of winter wheat. The dashed line is the corresponding estimation accuracy observed in the training process.</p>
Full article ">Figure 9
<p>LCC (SPAD value) maps of winter wheat obtained based on UAV-derived VIs and SVM regression algorithm.</p>
Full article ">
14 pages, 4404 KiB  
Article
Hyperspectral Reflectance-Based High Throughput Phenotyping to Assess Water-Use Efficiency in Cotton
by Sahila Beegum, Muhammad Adeel Hassan, Purushothaman Ramamoorthy, Raju Bheemanahalli, Krishna N. Reddy, Vangimalla Reddy and Kambham Raja Reddy
Agriculture 2024, 14(7), 1054; https://doi.org/10.3390/agriculture14071054 - 29 Jun 2024
Viewed by 1193
Abstract
Cotton is a pivotal global commodity underscored by its economic value and widespread use. In the face of climate change, breeding resilient cultivars for variable environmental conditions becomes increasingly essential. However, the process of phenotyping, crucial to breeding programs, is often viewed as [...] Read more.
Cotton is a pivotal global commodity underscored by its economic value and widespread use. In the face of climate change, breeding resilient cultivars for variable environmental conditions becomes increasingly essential. However, the process of phenotyping, crucial to breeding programs, is often viewed as a bottleneck due to the inefficiency of traditional, low-throughput methods. To address this limitation, this study utilizes hyperspectral remote sensing, a promising tool for assessing crucial crop traits across forty cotton varieties. The results from this study demonstrated the effectiveness of four vegetation indices (VIs) in evaluating these varieties for water-use efficiency (WUE). The prediction accuracy for WUE through VIs such as the simple ratio water index (SRWI) and normalized difference water index (NDWI) was higher (up to R2 = 0.66), enabling better detection of phenotypic variations (p < 0.05) among the varieties compared to physiological-related traits (from R2 = 0.21 to R2 = 0.42), with high repeatability and a low RMSE. These VIs also showed high Pearson correlations with WUE (up to r = 0.81) and yield-related traits (up to r = 0.63). We also selected high-performing varieties based on the VIs, WUE, and fiber quality traits. This study demonstrated that the hyperspectral-based proximal sensing approach helps rapidly assess the in-season performance of varieties for imperative traits and aids in precise breeding decisions. Full article
(This article belongs to the Special Issue Smart Agriculture Sensors and Monitoring Systems for Field Detection)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Experimental setup and (<b>b</b>) weather parameters observed during the experimental duration. A total of 40 varieties with three replications and three plants per replication were planted (a total of 360 pots). The picture (<b>a</b>) was taken 37 days after emergence. Weather data (the rainfall, incident solar radiation, air temperature, and wind) were obtained from a nearby weather station (Delta Agricultural Center, Mississippi State University Extension, North Farm) <a href="http://deltaweather.extension.msstate.edu/stations" target="_blank">http://deltaweather.extension.msstate.edu/stations</a> accessed on 10 May 2024.</p>
Full article ">Figure 2
<p>Mean hyperspectral reflectance of all 40 varieties (V1–V40).</p>
Full article ">Figure 3
<p>Repeatability of hyperspectral vegetation indices (SRWI, NDWI, NDWI1640, and NDWI2130) and handheld-instrument-based physiology-related traits (PhiPS2, ETR, Fv′/Fm′, A, E, and WUE) of cotton cultivars. Abbreviations: SRWI, simple ratio water index; NDWI, normalized difference water index; NDWI1640, normalized difference water index centered at 1640 nm; NDWI2130, normalized difference water index centered at 2130 nm; PhiPS2, PSII actual photochemical quantum yield; ETR, electron transport rate; Fv′/Fm′, PSII effective chlorophyll fluorescence; A, photosynthesis; E, transpiration; WUE, water-use efficiency.</p>
Full article ">Figure 4
<p>Linear regressions among hyperspectral vegetative indices (<b>a</b>) NDWI, (<b>b</b>) SRWI, (<b>c</b>) NDWI2130, and (<b>d</b>) NDWI1640 and water-use efficiency. Abbreviations: SRWI, simple ratio water index; NDWI, normalized difference water index; NDWI1640, normalized difference water index centered at 1640 nm; NDWI2130, normalized difference water index centered at 2130 nm; WUE, water-use efficiency (µmol CO<sub>2</sub>/mmol H<sub>2</sub>O); <span class="html-italic">R</span><sup>2</sup>, coefficient of determination; RMSE, root mean squared error.</p>
Full article ">Figure 5
<p>Linear regressions among photosynthesis-related traits (<b>a</b>) Fv′/Fm′, (<b>b</b>) PhiPS2, and (<b>c</b>) ETR and water-use efficiency. Abbreviations: WUE, water-use efficiency (µmol CO<sub>2</sub>/mmol H<sub>2</sub>O); PhiPS2, PSII actual photochemical quantum yield (µmol µmol<sup>−1</sup>); ETR, electron transport rate (µmol m<sup>−2</sup> s<sup>−1</sup>); Fv′/Fm′, PSII effective chlorophyll fluorescence; R<sup>2</sup>, coefficient of determination; RMSE, root mean squared error.</p>
Full article ">Figure 6
<p>Pearson correlations between vegetation indices, physiological and yield-related traits. Abbreviations: SRWI, simple ratio water index; NDWI, normalized difference water index; NDWI1640, normalized difference water index centered at 1640 nm; NDWI2130, normalized difference water index centered at 2130 nm; WUE, water-use efficiency; PhiPS2, PSII actual photochemical quantum yield; ETR, electron transport rate; Fv′/Fm′, PSII effective chlorophyll fluorescence; ISW, individual seed weight per plant; SNB, seed number per boll; A, photosynthesis; E, transpiration.</p>
Full article ">Figure 7
<p>Principal component analysis (PCA) biplot on 40 cultivars (V1 to V40) for (<b>a</b>) hyperspectral remote-sensing-based vegetation indices, physiological and yield-related traits, and (<b>b</b>) fiber quality traits of cotton cultivars. Abbreviations: SRWI, simple ratio water index; NDWI, normalized difference water index; NDWI1640, normalized difference water index centered at 1640 nm; NDWI2130, normalized difference water index centered at 2130 nm; WUE, water-use efficiency (µmol CO<sub>2</sub>/mmol H<sub>2</sub>O); PhiPS2, PSII actual photochemical quantum yield (µmol µmol<sup>−1</sup>); ETR, electron transport rate (µmol m<sup>−2</sup> s<sup>−1</sup>); Fv′/Fm′, PSII effective chlorophyll fluorescence; A, photosynthesis (µmol m<sup>−2</sup> s<sup>−1</sup>); E, transpiration (mmol m<sup>−2</sup> s<sup>−1</sup>); ISW, individual seed weight per plant (mg); SNB, seed number per boll; Len, fiber length (inch); Str, fiber strength (g/tex); Unif, fiber uniformity (%); Elo, fiber elongation (%); Mic, micronaire (-).</p>
Full article ">Figure 8
<p>Comparison of 40 cotton cultivars (V1 to V40) for (<b>a</b>) hyperspectral vegetation indices, (<b>b</b>) physiological traits, (<b>c</b>) yield, and (<b>d</b>) fiber-quality-related traits, (<b>e</b>) photosynthesis, water use efficiency and transpiration. Different colored dots indicate the average values of each trait, and bars indicate standard deviations. Abbreviations: SRWI, simple ratio water index; NDWI, normalized difference water index; NDWI1640, normalized difference water index centered at 1640 nm; NDWI2130, normalized difference water index centered at 2130 nm; WUE, water-use efficiency (µmol CO<sub>2</sub>/mmolH<sub>2</sub>O); PhiPS2, PSII actual photochemical quantum yield (µmol µmol<sup>−1</sup>); ETR, electron transport rate (µmol m<sup>−2</sup> s<sup>−1</sup>); Fv′/Fm′, PSII effective chlorophyll fluorescence; A, photosynthesis (µmol m<sup>−2</sup> s<sup>−1</sup>); E, transpiration (mmol m<sup>−2</sup> s<sup>−1</sup>); ISW, individual seed weight per plant (mg); SNB, seed number per boll; Len, fiber length (inch); Str, fiber strength (g/tex); Unif, fiber uniformity (%); Elo, fiber elongation (%); Mic, micronaire (-).</p>
Full article ">
16 pages, 339 KiB  
Article
Exploring Determinants of Health-Related Quality of Life in Emerging Adults with Type 1 Diabetes Mellitus: A Cross-Sectional Analysis
by María-Ángeles Núñez-Baila, Anjhara Gómez-Aragón, Armando-Manuel Marques-Silva and José Rafael González-López
Nutrients 2024, 16(13), 2059; https://doi.org/10.3390/nu16132059 - 28 Jun 2024
Viewed by 1253
Abstract
(1) Background: Emerging adulthood, from the age of 18 to 29 years, is a crucial phase for individuals with Type 1 Diabetes Mellitus, as it affects their Health-Related Quality of Life. (2) Methods: This cross-sectional study analyzes the influence of the Mediterranean diet, [...] Read more.
(1) Background: Emerging adulthood, from the age of 18 to 29 years, is a crucial phase for individuals with Type 1 Diabetes Mellitus, as it affects their Health-Related Quality of Life. (2) Methods: This cross-sectional study analyzes the influence of the Mediterranean diet, Diabetes duration, Hemoglobin A1c, and sleep disorders on Health-Relate Quality of Life in Type 1 Diabetes Mellitus. In this study, conducted in Andalusia, Spain, 362 emerging adults with Type 1 Diabetes Mellitus completed the Oviedo Sleep Questionnaire, the Adaptation of Mediterranean Diet Adherence Screener, and the Vida con Diabetes Tipo 1 (ViDa1) Health-Related Quality of Life questionnaire between October 2021 and July 2022. Pearson correlation coefficients and a multiple regression analysis were conducted for each Health-Related Quality of Life in Type 1 Diabetes Mellitus dimension (Interference with Life, Well-being, Self-care, and Concern about the Condition) for overall sample and separately for males and females. (3) Results: Different and significant correlations are found among factors such as Age, Body Mass Index, Currently being a student, Hemoglobin A1c, Sleep satisfaction, Insomnia, Hypersomnolence, and Adherence to Mediterranean diet. Notably, Insomnia is a main predictor for Interference with Life, Well-being, and Concern about the Condition, especially for females. (4) Conclusions: Insomnia is the main predictor of Health-Related Quality of Life in Type 1 Diabetes Mellitus among Andalusian emerging adults with this condition. Consequently, a regular assessment of sleep and Health-Related Quality of Life from a gender perspective in this age group is crucial. Full article
(This article belongs to the Special Issue Advances in Nutrition and Lifestyle Interventions for Type 1 Diabetes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
13 pages, 4563 KiB  
Article
Biomass Estimation of Milk Vetch Using UAV Hyperspectral Imagery and Machine Learning
by Hao Hu, Hongkui Zhou, Kai Cao, Weidong Lou, Guangzhi Zhang, Qing Gu and Jianhong Wang
Remote Sens. 2024, 16(12), 2183; https://doi.org/10.3390/rs16122183 - 16 Jun 2024
Cited by 1 | Viewed by 899
Abstract
Milk vetch (Astragalus sinicus L.) is a winter-growing plant that can enhance soil fertility and provide essential nutrients for subsequent season crops. The fertilizing capacity of milk vetch is closely related to its above-ground biomass. Compared to the manual measurement methods of [...] Read more.
Milk vetch (Astragalus sinicus L.) is a winter-growing plant that can enhance soil fertility and provide essential nutrients for subsequent season crops. The fertilizing capacity of milk vetch is closely related to its above-ground biomass. Compared to the manual measurement methods of milk vetch biomass, remote sensing-based estimation methods have the advantages of rapid, noninvasive, and large-scale measurement. However, few studies have been conducted on remote sensing-based estimation of milk vetch biomass. To address this shortcoming, this study proposes combining unmanned aerial vehicle (UAV)-based hyperspectral imagery and machine learning algorithms for accurate estimation of milk vetch biomass. Through the analysis of hyperspectral images and feature selection based on the Pearson correlation and principal component analysis, vegetation indices (VIs), including near-infrared reflectance (NIR), red-edge spectral transform index (RE), and difference vegetation index (DVI), are selected as estimation metrics of the model development process. Four machine learning methods, including random forest (RF), multiple linear regression (MLR), deep neural network (DNN), and support vector machine (SVM), are used to construct the biomass models. The results show that the RF estimation model exhibits the highest coefficient of determination (R2) of 0.950 and the lowest relative root-mean-squared error (RRMSE) of 14.86% among all the models. Notably, the DNN model demonstrates promising performance on the test set, with the R2 and RRMSE values slightly superior and inferior to those of the RF, respectively. The proposed method based on UAV imagery and machine learning can provide an accurate and reliable large-scale estimation of milk vetch biomass. Full article
(This article belongs to the Special Issue Application of Satellite and UAV Data in Precision Agriculture)
Show Figures

Figure 1

Figure 1
<p>Study area (<b>A</b>,<b>D</b>) and a UAV used in this study; (<b>B</b>): DJI Matrice 600 Pro equipped with a Pika XC2 hyperspectral imager; (<b>C</b>): DJI Phantom 4 RTK.</p>
Full article ">Figure 2
<p>Design workflow of the biomass estimation model.</p>
Full article ">Figure 3
<p>DNN network structure used in this study.</p>
Full article ">Figure 4
<p>Correlations between the VIs and milk vetch biomass in 2022 (<b>A</b>) and 2023 (<b>B</b>). * indicates significant at <span class="html-italic">p</span> &lt; 0.05 with two-tailed test.</p>
Full article ">Figure 5
<p>PCA plots of the milk vetch VIs in 2022 (<b>A</b>) and 2023 (<b>B</b>).</p>
Full article ">Figure 6
<p>Training and test results of milk vetch in 2022 (<b>A</b>) and 2023 (<b>B</b>). A1 (B1), A2 (B2), A3 (B3), and A4 (B4) were plots of predicted biomass to measured biomass, respectively.</p>
Full article ">Figure 7
<p>Training and test results of milk vetch for two-year data (2022 and 2023).</p>
Full article ">Figure 8
<p>Estimated and measured biomass data in 2022 (<b>A</b>) and 2023 (<b>B</b>).</p>
Full article ">
20 pages, 27585 KiB  
Article
Non-Destructive Monitoring of Peanut Leaf Area Index by Combing UAV Spectral and Textural Characteristics
by Dan Qiao, Juntao Yang, Bo Bai, Guowei Li, Jianguo Wang, Zhenhai Li, Jincheng Liu and Jiayin Liu
Remote Sens. 2024, 16(12), 2182; https://doi.org/10.3390/rs16122182 - 16 Jun 2024
Cited by 1 | Viewed by 806
Abstract
The leaf area index (LAI) is a crucial metric for indicating crop development in the field, essential for both research and the practical implementation of precision agriculture. Unmanned aerial vehicles (UAVs) are widely used for monitoring crop growth due to their rapid, repetitive [...] Read more.
The leaf area index (LAI) is a crucial metric for indicating crop development in the field, essential for both research and the practical implementation of precision agriculture. Unmanned aerial vehicles (UAVs) are widely used for monitoring crop growth due to their rapid, repetitive capture ability and cost-effectiveness. Therefore, we developed a non-destructive monitoring method for peanut LAI, combining UAV vegetation indices (VI) and texture features (TF). Field experiments were conducted to capture multispectral imagery of peanut crops. Based on these data, an optimal regression model was constructed to estimate LAI. The initial computation involves determining the potential spectral and textural characteristics. Subsequently, a comprehensive correlation study between these features and peanut LAI is conducted using Pearson’s product component correlation and recursive feature elimination. Six regression models, including univariate linear regression, support vector regression, ridge regression, decision tree regression, partial least squares regression, and random forest regression, are used to determine the optimal LAI estimation. The following results are observed: (1) Vegetation indices exhibit greater correlation with LAI than texture characteristics. (2) The choice of GLCM parameters for texture features impacts estimation accuracy. Generally, smaller moving window sizes and higher grayscale quantization levels yield more accurate peanut LAI estimations. (3) The SVR model using both VI and TF offers the utmost precision, significantly improving accuracy (R2 = 0.867, RMSE = 0.491). Combining VI and TF enhances LAI estimation by 0.055 (VI) and 0.541 (TF), reducing RMSE by 0.093 (VI) and 0.616 (TF). The findings highlight the significant improvement in peanut LAI estimation accuracy achieved by integrating spectral and textural characteristics with appropriate parameters. These insights offer valuable guidance for monitoring peanut growth. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>Overview of the study site, where (<b>a</b>) illustrates location of Tai’an in China and (<b>b</b>) illustrates field measurements.</p>
Full article ">Figure 2
<p>UAV remote sensing platform. (<b>a</b>): DJI Phantom 4 multispectral quadcopter; (<b>b</b>): The calibration panel.</p>
Full article ">Figure 3
<p>Visualization of fifteen vegetation indices that are rendered from red to green according to the values of the associated vegetation index. (<b>a</b>) NDVI; (<b>b</b>) NDVIre; (<b>c</b>) Cire; (<b>d</b>) DVI; (<b>e</b>) MSAVI; (<b>f</b>) OSAVI; (<b>g</b>) TVI; (<b>h</b>) GRVI; (<b>i</b>) SAVI; (<b>j</b>) RESR; (<b>k</b>) MCARI; (<b>l</b>) RDVI; (<b>m</b>) MSR; (<b>n</b>) EVI; and (<b>o</b>) GNDVI.</p>
Full article ">Figure 4
<p>Visualization of eight GLCM-based texture features that are rendered from red to green according to the values of associated texture feature. (<b>a</b>) MEA; (<b>b</b>) VAR; (<b>c</b>) HOM; (<b>d</b>) CON; (<b>e</b>) DIS; (<b>f</b>) ENT; (<b>g</b>) SEC; and (<b>h</b>) COR.</p>
Full article ">Figure 5
<p>Diagram of the multivariate regression models. (<b>a</b>) SVR model; (<b>b</b>) RR model; (<b>c</b>) DTR model; (<b>d</b>) PLSR model; and (<b>e</b>) RFR model.</p>
Full article ">Figure 6
<p>Experimental methodology and statistical analysis procedure used in this work.</p>
Full article ">Figure 7
<p>Correlations between LAI and vegetation indices. Above the diagonal: Pearson correlation coefficients; Below the diagonal: scatter plots representing the linear relationships between variables; Diagonal: the data distribution of each variable The circle represents discrete points, the red line represents the fitting curve, and *** indicates that the p-value is less than 0.001.</p>
Full article ">Figure 8
<p>The ranking of each feature’s importance.</p>
Full article ">Figure 9
<p>The accuracy of models based on different parameters for calculating GLCM features.</p>
Full article ">Figure 10
<p>Estimation results of different models for peanut LAI using a combination of VI and TF. (<b>a</b>) SVR model; (<b>b</b>) RR model; (<b>c</b>) DTR model; (<b>d</b>) PLSR model; and (<b>e</b>) RFR model.</p>
Full article ">Figure 11
<p>Different models’ coefficients of determination (R<sup>2</sup>) for the LAI estimation of peanut from spectral variables, texture features, and their combination at different window sizes and grayscale quantization levels in GLCM. (<b>a</b>) SVR model; (<b>b</b>) RR model; (<b>c</b>) DTR model; (<b>d</b>) PLSR model; and (<b>e</b>) RFR model.</p>
Full article ">
13 pages, 1156 KiB  
Article
Classification of Soybean Genotypes as to Calcium, Magnesium, and Sulfur Content Using Machine Learning Models and UAV–Multispectral Sensor
by Dthenifer Cordeiro Santana, Izabela Cristina de Oliveira, Sâmela Beutinger Cavalheiro, Paulo Henrique Menezes das Chagas, Marcelo Carvalho Minhoto Teixeira Filho, João Lucas Della-Silva, Larissa Pereira Ribeiro Teodoro, Cid Naudi Silva Campos, Fábio Henrique Rojo Baio, Carlos Antonio da Silva Junior and Paulo Eduardo Teodoro
AgriEngineering 2024, 6(2), 1581-1593; https://doi.org/10.3390/agriengineering6020090 - 1 Jun 2024
Viewed by 866
Abstract
Making plant breeding programs less expensive, fast, practical, and accurate, especially for soybeans, promotes the selection of new soybean genotypes and contributes to the emergence of new varieties that are more efficient in absorbing and metabolizing nutrients. Using spectral information from soybean genotypes [...] Read more.
Making plant breeding programs less expensive, fast, practical, and accurate, especially for soybeans, promotes the selection of new soybean genotypes and contributes to the emergence of new varieties that are more efficient in absorbing and metabolizing nutrients. Using spectral information from soybean genotypes combined with nutritional information on secondary macronutrients can help genetic improvement programs select populations that are efficient in absorbing and metabolizing these nutrients. In addition, using machine learning algorithms to process this information makes the acquisition of superior genotypes more accurate. Therefore, the objective of the work was to verify the classification performance of soybean genotypes regarding secondary macronutrients by ML algorithms and different inputs. The experiment was conducted in the experimental area of the Federal University of Mato Grosso do Sul, municipality of Chapadão do Sul, Brazil. Soybean was sown in the 2019/20 crop season, with the planting of 103 F2 soybean populations. The experimental design used was randomized blocks, with two replications. At 60 days after crop emergence (DAE), spectral images were collected with a Sensifly eBee RTK fixed-wing remotely piloted aircraft (RPA), with autonomous takeoff control, flight plan, and landing. At the reproductive stage (R1), three leaves were collected per plant to determine the macronutrients calcium (Ca), magnesium (Mg), and sulfur (S) levels. The data obtained from the spectral information and the nutritional values of the genotypes in relation to Ca, Mg, and S were subjected to a Pearson correlation analysis; a PC analysis was carried out with a k-means algorithm to divide the genotypes into clusters. The clusters were taken as output variables, while the spectral data were used as input variables for the classification models in the machine learning analyses. The configurations tested in the models were spectral bands (SBs), vegetation indices (VIs), and a combination of both. The combination of machine learning algorithms with spectral data can provide important biological information about soybean plants. The classification of soybean genotypes according to calcium, magnesium, and sulfur content can maximize time, effort, and labor in field evaluations in genetic improvement programs. Therefore, the use of spectral bands as input data in random forest algorithms makes the process of classifying soybean genotypes in terms of secondary macronutrients efficient and important for researchers in the field. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the experimental area in Chapadão do Sul-MS, Brazil; photographic area of the experimental area.</p>
Full article ">Figure 2
<p>Pearson correlation scatterplot with spectral and secondary macronutrients.</p>
Full article ">Figure 3
<p>Principal Component (PC) for clusters based on Ca, M, and S contents of soybean genotypes based on k-means.</p>
Full article ">Figure 4
<p>Boxplot with Ca, Mg, and S means for clustered data. Means followed by the same letters do not differ for the cluster by the Scott–Knott test at 5% probability.</p>
Full article ">Figure 5
<p>Boxplot with clustering means for percent correct classification regarding the machine learning models. Means followed by the same uppercase letters do not differ for the inputs tested by the Scott–Knott test at 5% probability; means followed by the same lowercase letters do not differ for the algorithms tested by the Scott–Knott test at 5% probability.</p>
Full article ">Figure 6
<p>Boxplot with clustering means for kappa regarding machine learning models. Means followed by the same uppercase letters do not differ for the inputs tested by the Scott–Knott test at 5% probability; means followed by the same lowercase letters do not differ for the algorithms tested by the Scott–Knott test at 5% probability.</p>
Full article ">Figure 7
<p>Boxplot with clustering means for F-score regarding the machine learning models tested. Means followed by the same uppercase letters do not differ for the inputs tested by the Scott–Knott test at 5% probability; means followed by the same lowercase letters do not differ for the algorithms tested by the Scott–Knott test at 5% probability.</p>
Full article ">
26 pages, 4686 KiB  
Article
Winter Durum Wheat Disease Severity Detection with Field Spectroscopy in Phenotyping Experiment at Leaf and Canopy Level
by Dessislava Ganeva, Lachezar Filchev, Eugenia Roumenina, Rangel Dragov, Spasimira Nedyalkova and Violeta Bozhanova
Remote Sens. 2024, 16(10), 1762; https://doi.org/10.3390/rs16101762 - 16 May 2024
Viewed by 975
Abstract
Accurate disease severity assessment is critical for plant breeders, as it directly impacts crop yield. While hyperspectral remote sensing has shown promise for disease severity assessment in breeding experiments, most studies have focused on either leaf or canopy levels, neglecting the valuable insights [...] Read more.
Accurate disease severity assessment is critical for plant breeders, as it directly impacts crop yield. While hyperspectral remote sensing has shown promise for disease severity assessment in breeding experiments, most studies have focused on either leaf or canopy levels, neglecting the valuable insights gained from a combined approach. Moreover, many studies have centered on experiments involving a single disease and a few genotypes. However, this approach needs to accurately represent the challenges encountered in field conditions, where multiple diseases could occur simultaneously. To address these gaps, our current study analyses a combination of diseases, yellow rust, brown rust, and yellow leaf spots, collectively evaluated as the percentage of the diseased leaf area relative to the total leaf area (DA) at both leaf and canopy levels, using hyperspectral data from an ASD field spectrometer. We quantitatively estimate overall disease severity across fifty-two winter durum wheat genotypes categorized into early (medium milk) and late (late milk) groups based on the phenophase. Chlorophyll content (CC) within each group is studied concerning infection response, and a correlation analysis is conducted for each group with nine vegetation indices (VI) known for their sensitivity to rust and leaf spot infection in wheat. Subsequent parametric (linear and polynomial) and nonparametric (partial least squares and kernel ridge) regression analyses were performed using all available spectral bands. We found a significant reduction in Leaf CC (>30%) in the late group and Canopy CC (<10%) for both groups. YROI and LRDSI_1 are the VIs that exhibited notable and strong negative correlations with Leaf CC in the late group, with a Pearson coefficient of −0.73 and −0.72, respectively. Interestingly, spectral signatures between the early and late disease groups at both leaf and canopy levels exhibit opposite trends. The regression analysis showed we could retrieve leaf CC only for the late group, with R2 of 0.63 and 0.42 for the cross-validation and test datasets, respectively. Canopy CC retrieval required separate models for each group: the late group achieved R2 of 0.61 and 0.37 (cross-validation and test), while the early group achieved R2 of 0.48 and 0.50. Similar trends were observed for canopy DA, with separate models for early and late groups achieving comparable R2 values of 0.53 and 0.51 (cross-validation) and 0.35 and 0.36 (test), respectively. All of our models had medium accuracy and tended to overfit. In this study, we analyzed the spectral response mechanism associated with durum wheat diseases, offering a novel crop disease severity assessment approach. Additionally, our findings serve as a foundation for detecting resistant wheat varieties, which is the most economical and environmentally friendly management strategy for wheat leaf diseases on a large scale in the future. Full article
(This article belongs to the Special Issue Spectral Imaging Technology for Crop Disease Detection)
Show Figures

Figure 1

Figure 1
<p>Examples of flag leaf diseases observed in the studied wheat genotypes and brief descriptions of each disease. (<b>a</b>) Yellow rust. (<b>b</b>) Brown rust. (<b>c</b>) Leaf spots [<a href="#B12-remotesensing-16-01762" class="html-bibr">12</a>,<a href="#B13-remotesensing-16-01762" class="html-bibr">13</a>].</p>
Full article ">Figure 2
<p>Workflow for disease severity assessment in the study. The process involved data acquisition and pre-processing, followed by data analysis and presentation of results. Various scenarios were explored based on the input data and the type of regression, whether parametric or nonparametric. An additional analysis was conducted to investigate the influence of disease and phenophase on chlorophyll content (CC). These steps were carried out independently at both leaf and canopy levels.</p>
Full article ">Figure 3
<p>Location of the study area. The phenotyping experiment in the Field Crops Institute, Chirpan, South Bulgaria.</p>
Full article ">Figure 4
<p>Canopy and leaf images of four winter durum wheat genotypes at two growth stages and different disease severity levels (B). Within each genotype, leaves are numbered sequentially from top to bottom at the leaf level.</p>
Full article ">Figure 4 Cont.
<p>Canopy and leaf images of four winter durum wheat genotypes at two growth stages and different disease severity levels (B). Within each genotype, leaves are numbered sequentially from top to bottom at the leaf level.</p>
Full article ">Figure 5
<p>Plot of average reflectance spectra per wavelength of (<b>a</b>) the leaf early group with 140 samples, (<b>b</b>) the leaf late group with 68 samples, (<b>c</b>) the canopy early group with 129 samples, and (<b>d</b>) the canopy late group with 65 samples.</p>
Full article ">Figure A1
<p>Visualization with Q-Q plot of the data for normality assumptions and results of Shapiro–Wilk test at leaf level.</p>
Full article ">Figure A2
<p>Visualization with boxplot and violin of the data for equal variance assumptions and results of Levene’s test at leaf level.</p>
Full article ">Figure A3
<p>Visualization with Q-Q plot of the data for normality assumptions and results of Shapiro–Wilk test at canopy level.</p>
Full article ">Figure A4
<p>Visualization with boxplot and violin of the data for equal variance assumptions and results of Levene’s test at canopy level.</p>
Full article ">Figure A5
<p>Spectra used for the regression analysis for (<b>a</b>) the leaf early group, (<b>b</b>) the leaf late group, (<b>c</b>) the canopy early group, and (<b>d</b>) the canopy late group.</p>
Full article ">
15 pages, 913 KiB  
Article
Sociodemographic and Clinical Determinants on Health-Related Quality of Life in Emerging Andalusian Adults with Type 1 Diabetes: A Cross-Sectional Study
by María-Ángeles Núñez-Baila, Anjhara Gómez-Aragón and José Rafael González-López
J. Clin. Med. 2024, 13(1), 240; https://doi.org/10.3390/jcm13010240 - 31 Dec 2023
Cited by 1 | Viewed by 1464
Abstract
(1) Background: Having type 1 diabetes during emerging adulthood can impact quality of life due to the challenge of balancing optimal glycemic blood levels with a period of transition and exploration. The purpose of this study was to characterize the quality of life [...] Read more.
(1) Background: Having type 1 diabetes during emerging adulthood can impact quality of life due to the challenge of balancing optimal glycemic blood levels with a period of transition and exploration. The purpose of this study was to characterize the quality of life of emerging adults aged 18 to 29 years with type 1 diabetes and to determine the associations between dimensions of Health-Related Quality of Life in type 1 diabetes and sociodemographic and diabetes-related variables. (2) Methods: This cross-sectional descriptive study was conducted in Andalusia, Spain, from October 2021 to July 2022. A total of 362 emerging adults with type 1 diabetes (67.4% women, mean age 22.8 ± 3.4 years) participated. Data were gathered via sociodemographic information form and the ViDa1 scale. Statistical evaluations, encompassing descriptive analyses, t-tests, ANOVA, Pearson correlations, and logistic regression, were conducted using SPSSv26, adhering to STROBE guidelines. (3) Results: Among the participants, 52.1% have a glycosylated hemoglobin level over 7%. Interference with Life is correlated with sex, age, and age at diagnosis, with age being the only predictor. Self-Care is correlated with and predicted by glycosylated hemoglobin levels. Well-being is correlated with and predicted by sex, Body Mass Index, and glycosylated hemoglobin levels. Concern about the Condition is correlated with and predicted by sex and glycosylated hemoglobin levels. (4) Conclusions: Despite concerns about their disease, participants generally maintain optimal levels of Health-Related Quality of Life in type 1 diabetes. Predictive factors for Health-Related Quality of Life in type 1 diabetes in this group include sex, age, Body Mass Index, and glycosylated hemoglobin. Full article
(This article belongs to the Section Endocrinology & Metabolism)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flow chart of participant enrollment process.</p>
Full article ">Figure 2
<p>Key factors reducing Health-Related Quality of Life in Type 1 diabetes mellitus—insights from <span class="html-italic">t</span>-Student and ANOVA. <sup>1</sup> A1C: Glycosylated hemoglobin A1C. <sup>2</sup> T1DM: Type 1 diabetes mellitus. <sup>3</sup> BMI: Body Mass Index. <sup>4</sup> HRQoL: Health-Related to Quality of Life.↑: Increase in variable value.</p>
Full article ">
18 pages, 2984 KiB  
Article
Failure Identification Method of Sound Signal of Belt Conveyor Rollers under Strong Noise Environment
by Yuxuan Ban, Chunyang Liu, Fang Yang, Nan Guo, Xiqiang Ma, Xin Sui and Yan Huang
Electronics 2024, 13(1), 34; https://doi.org/10.3390/electronics13010034 - 20 Dec 2023
Cited by 2 | Viewed by 994
Abstract
Accurately extracting faulty sound signals from belt conveyor rollers within the high-noise environment of coal mine operations presents a formidable challenge. To address this issue, this study introduces an innovative fault diagnosis method that merges the variational modal de-composition (VMD) model with the [...] Read more.
Accurately extracting faulty sound signals from belt conveyor rollers within the high-noise environment of coal mine operations presents a formidable challenge. To address this issue, this study introduces an innovative fault diagnosis method that merges the variational modal de-composition (VMD) model with the Swin Transformer deep learning network model. First, the study employed the adaptive VMD method to eliminate intense noise from the original signal of the rollers, while also assessing the reconstruction accuracy of the VMD signal across different modal components. Subsequently, we delved into the impact of the parameter structure of the Swin Transformer network model on the fault diagnosis accuracy. Finally, the accuracy of the method was validated using a sound test dataset from the rollers. The results indicated that optimizing the K-value of the VMD method effectively reduced the noise in the reconstructed signal, and the Swin Transformer excelled in extracting both local and global features. Specifically, on the conveyor roller sound dataset, it was shown that, after the VMD reconstruction of the signal so that the highest Pearson correlation coefficient corresponded to a modal component of 3 and adjusting the parameters of the Swin Transformer coding layer, the combination of the VMD+Swin-S model achieved an accuracy of 99.36%, while the VMD+Swin-T model achieved an accuracy of 98.6%. Meanwhile, the accuracy of the VMD+Swin-S model was higher than that of the VMD + CNN model combination, with 95.4% accuracy, and the VMD+ViT model, with 97.68% accuracy. In the example application experiments, compared with other models the VMD+Swin-S model achieved the highest accuracy rate at all three speeds, with 98.67%, 98.32%, and 97.65%, respectively. Overall, this approach demonstrated high accuracy and robustness, rendering it an optimal choice for diagnosing conveyor belt roller faults within environments characterized by strong noise. Full article
Show Figures

Figure 1

Figure 1
<p>VMD+Swin Transformer fault diagnosis process.</p>
Full article ">Figure 2
<p>VMD decomposition and reconstruction signal effect. (<b>a</b>) Original time-domain signal waveform. (<b>b</b>) Original signal time-domain characteristics. (<b>c</b>) Time-domain signal waveform after adding noise. (<b>d</b>) Time–frequency domain characteristics after adding noise. (<b>e</b>) Time-domain signal waveform after reconstruction. (<b>f</b>) Time–frequency domain characteristics after reconstruction.</p>
Full article ">Figure 3
<p>Swin Transformer structure diagram.</p>
Full article ">Figure 4
<p>Different Transformer models encoder structures. (<b>a</b>) Traditional Transformer Encoder. (<b>b</b>) Swin Transformer Block.</p>
Full article ">Figure 5
<p>Mechanism of patch merging.</p>
Full article ">Figure 6
<p>Shifted window procedure. (<b>a</b>) W-MSA in layer l and SW-MSA in layer l + 1. (<b>b</b>) Shift configuration batch calculations.</p>
Full article ">Figure 7
<p>Influence of different modal component selections on diagnostic accuracy.</p>
Full article ">Figure 8
<p>Training accuracy of various Swin Transformer models.</p>
Full article ">Figure 9
<p>Diagnostic accuracy of various Swin Transformer models.</p>
Full article ">Figure 10
<p>Comparison of diagnosis accuracy between models.</p>
Full article ">Figure 11
<p>Comparison of fault identification accuracy of models at different rotational speeds.</p>
Full article ">
20 pages, 5209 KiB  
Article
Using Remote and Proximal Sensing Data and Vine Vigor Parameters for Non-Destructive and Rapid Prediction of Grape Quality
by Hongyi Lyu, Miles Grafton, Thiagarajah Ramilan, Matthew Irwin, Hsiang-En Wei and Eduardo Sandoval
Remote Sens. 2023, 15(22), 5412; https://doi.org/10.3390/rs15225412 - 19 Nov 2023
Cited by 5 | Viewed by 1709
Abstract
The traditional method for determining wine grape total soluble solid (TSS) is destructive laboratory analysis, which is time consuming and expensive. In this study, we explore the potential of using different predictor variables from various advanced techniques to predict the grape TSS in [...] Read more.
The traditional method for determining wine grape total soluble solid (TSS) is destructive laboratory analysis, which is time consuming and expensive. In this study, we explore the potential of using different predictor variables from various advanced techniques to predict the grape TSS in a non-destructive and rapid way. Calculating Pearson’s correlation coefficient between the vegetation indices (VIs) obtained from UAV multispectral imagery and grape TSS resulted in a strong correlation between OSAVI and grape TSS with a coefficient of 0.64. Additionally, seven machine learning models including ridge regression and lasso regression, k-Nearest neighbor (KNN), support vector regression (SVR), random forest regression (RFR), extreme gradient boosting (XGBoost), and artificial neural network (ANN) are used to build the prediction models. The predictor variables include the unmanned aerial vehicles (UAV) derived VIs, and other ancillary variables including normalized difference vegetation index (NDVI_proximal) and soil electrical conductivity (ECa) measured by proximal sensors, elevation, slope, trunk circumference, and day of the year for each sampling date. When using 23 VIs and other ancillary variables as input variables, the results show that ensemble learning models (RFR, and XGBoost) outperform other regression models when predicting grape TSS, with the average of root mean square error (RMSE) of 1.19 and 1.2 °Brix, and coefficient of determination (R2) of 0.52 and 0.52, respectively, during the 20 times testing process. In addition, this study examines the prediction performance of using optimized soil adjusted vegetation index (OSAVI) or normalized green-blue difference index (NGBDI) as the main input for different machine learning models with other ancillary variables. When using OSAVI-based models, the best prediction model is RFR with an average R2 of 0.51 and RMSE of 1.19 °Brix, respectively. For NGBDI-based model, the RFR model showed the best average result of predicting TSS were a R2 of 0.54 and a RMSE of 1.16 °Brix, respectively. The approach proposed in this study provides an opportunity to grape growers to estimate the whole vineyard grape TSS in a non-destructive way. Full article
(This article belongs to the Special Issue Advanced Artificial Intelligence for Environmental Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The location of sampling vines in PN (<b>a</b>); HN (<b>b</b>). (Points represent the location of sampling vines).</p>
Full article ">Figure 2
<p>Total daily precipitation, average temperature, and irradiance recorded by on-site weather station (blue bar represents the precipitation).</p>
Full article ">Figure 3
<p>Vine row segmentation workflow.</p>
Full article ">Figure 4
<p>The boxplot of grape TSS during study period (diamond records outliers).</p>
Full article ">Figure 5
<p>Pearson’s correlation coefficient between VIs and grape TSS (Different colors represent different sampling date).</p>
Full article ">Figure 6
<p>Spatial interpolation map of soil EC<sub>a</sub> (<b>a</b>); elevation (<b>b</b>); trunk circumference (<b>c</b>); NDVI (<b>d</b>); PCD (<b>e</b>) in PN vineyard.</p>
Full article ">Figure 7
<p>Spatial interpolation map of soil EC<sub>a</sub> (<b>a</b>); elevation (<b>b</b>); trunk circumference (<b>c</b>); NDVI (<b>d</b>); PCD (<b>e</b>) in HN vineyard.</p>
Full article ">Figure 8
<p>The boxplot of different machine learning model performance in R<sup>2</sup> (<b>a</b>); RMSE (<b>b</b>). (Different letters between any two groups represents significant difference between them, if two groups have the same letter then this indicates that they are not statistically different).</p>
Full article ">Figure 9
<p>The boxplot of different OSAVI-based model performance in R<sup>2</sup> (<b>a</b>); RMSE (<b>b</b>). (Different letters between any two groups represents significant difference between them, if two groups have the same letter then this indicates that they are not statistically different).</p>
Full article ">Figure 10
<p>The boxplot of different NGBDI-based model performance in R<sup>2</sup> (<b>a</b>); RMSE (<b>b</b>). (Different letters between any two groups represents significant difference between them, if two groups have the same letter then this indicates that they are not statistically different).</p>
Full article ">Figure A1
<p>Measured TSS value comparison against predicted TSS value for the best NGBDI-based RFR model.</p>
Full article ">
17 pages, 2217 KiB  
Article
Multispectral Vegetation Indices and Machine Learning Approaches for Durum Wheat (Triticum durum Desf.) Yield Prediction across Different Varieties
by Giuseppe Badagliacca, Gaetano Messina, Salvatore Praticò, Emilio Lo Presti, Giovanni Preiti, Michele Monti and Giuseppe Modica
AgriEngineering 2023, 5(4), 2032-2048; https://doi.org/10.3390/agriengineering5040125 - 2 Nov 2023
Cited by 8 | Viewed by 3256
Abstract
Durum wheat (Triticum durum Desf.) is one of the most widely cultivated cereal species in the Mediterranean basin, supporting pasta, bread and other typical food productions. Considering its importance for the nutrition of a large population and production of high economic value, [...] Read more.
Durum wheat (Triticum durum Desf.) is one of the most widely cultivated cereal species in the Mediterranean basin, supporting pasta, bread and other typical food productions. Considering its importance for the nutrition of a large population and production of high economic value, its supply is of strategic significance. Therefore, an early and accurate crop yield estimation may be fundamental to planning the purchase, storage, and sale of this commodity on a large scale. Multispectral (MS) remote sensing (RS) of crops using unpiloted aerial vehicles (UAVs) is a powerful tool to assess crop status and productivity with a high spatial–temporal resolution and accuracy level. The object of this study was to monitor the behaviour of thirty different durum wheat varieties commonly cultivated in Italy, taking into account their spectral response to different vegetation indices (VIs) and assessing the reliability of this information to estimate their yields by Pearson’s correlation and different machine learning (ML) approaches. VIs allowed us to separate the tested wheat varieties into different groups, especially when surveyed in April. Pearson’s correlations between VIs and grain yield were good (R2 > 0.7) for a third of the varieties tested; the VIs that best correlated with grain yield were CVI, GNDVI, MTVI, MTVI2, NDRE, and SR RE. Implementing ML approaches with VIs data highlighted higher performance than Pearson’s correlations, with the best results observed by random forest (RF) and support vector machine (SVM) models. Full article
(This article belongs to the Section Remote Sensing in Agriculture)
Show Figures

Figure 1

Figure 1
<p>On the left is the location of the experimental site, while on the right is the list of the thirty tested wheat cultivars and the orthomosaic with the field plots highlighted in yellow (RGB composition of April 2022 UAV flight). Below, the crop cycle of the durum wheat.</p>
Full article ">Figure 2
<p>The workflow of the adopted methodology. The first part shows the UAV survey and the data pre-processing. The second part shows the image data processing and the selected vegetation indices (VIs). The third part shows the statistics and machine learning (ML) analysis.</p>
Full article ">Figure 3
<p>Principal component analysis (PCA) of the 30 wheat cultivars calculated from their vegetation indices (VIs) responses on the April survey. PC1 is the first principal component, and PC2 is the second principal component.</p>
Full article ">Figure 4
<p>Principal component analysis of the 30 wheat cultivars calculated from their VI responses on the May survey. PC1 is the first principal component, and PC2 is the second principal component.</p>
Full article ">
17 pages, 5508 KiB  
Article
Quantitative Analysis of Soil Cd Content Based on the Fusion of Vis-NIR and XRF Spectral Data in the Impacted Area of a Metallurgical Slag Site in Gejiu, Yunnan
by Zhenlong Zhang, Zhe Wang, Ying Luo, Jiaqian Zhang, Xiyang Feng, Qiuping Zeng, Duan Tian, Chao Li, Yongde Zhang, Yuping Wang, Shu Chen and Li Chen
Processes 2023, 11(9), 2714; https://doi.org/10.3390/pr11092714 - 11 Sep 2023
Viewed by 1014
Abstract
Vis-NIR and XRF spectroscopy are widely used in monitoring heavy metals in soil due to their advantages of being fast, non-destructive, cost-effective, and non-polluting. However, when used individually, XRF and vis-NIR may not meet the accuracy requirements for Cd determination. In this study, [...] Read more.
Vis-NIR and XRF spectroscopy are widely used in monitoring heavy metals in soil due to their advantages of being fast, non-destructive, cost-effective, and non-polluting. However, when used individually, XRF and vis-NIR may not meet the accuracy requirements for Cd determination. In this study, we focused on the impact area of a non-ferrous metal smelting slag site in Gejiu City, Yunnan Province, fused the pre-selected vis-NIR and XRF spectra using the Pearson correlation coefficient (PCC), and identified the characteristic spectra using the competitive adaptive reweighted sampling (CARS) method. Based on this, a quantitative model for soil Cd concentration was established using partial least squares regression (PLSR). The results showed that among the four fusion spectral quantitative models constructed, the model combining vis-NIR spectral second-order derivative transformation and XRF spectral first-order derivative transformation (D2(vis-NIR) + D1(XRF)) had the highest coefficient of determination (R2 = 0.9505) and the smallest root mean square error (RMSE = 0.1174). Compared to the estimation models built using vis-NIR and XRF spectra alone, the average computational time of the fusion models was reduced by 68.19% and 63.92%, respectively. This study provides important technical means for real-time and large-scale on-site rapid estimation of Cd content using multi-source spectral fusion. Full article
(This article belongs to the Special Issue Advances in Remediation of Contaminated Sites: Volume II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geographical location of the study area.</p>
Full article ">Figure 2
<p>Geological map.</p>
Full article ">Figure 3
<p>PLSR model with D2 transformation of vis-NIR (the rest of the transformations are in <a href="#app1-processes-11-02714" class="html-app">Figure S6</a>).</p>
Full article ">Figure 4
<p>PLSR model with different transformations of XRF (the remaining transformations are in the attached <a href="#app1-processes-11-02714" class="html-app">Figure S7</a>).</p>
Full article ">Figure 5
<p>PLSR model with different transformation methods of FS.</p>
Full article ">Figure 6
<p>Comparison of <span class="html-italic">R</span><sup>2</sup> for each of the 5 spectra with 5 operations.</p>
Full article ">
27 pages, 11829 KiB  
Article
Non−Invasive Assessment, Classification, and Prediction of Biophysical Parameters Using Reflectance Hyperspectroscopy
by Renan Falcioni, Glaucio Leboso Alemparte Abrantes dos Santos, Luis Guilherme Teixeira Crusiol, Werner Camargos Antunes, Marcelo Luiz Chicati, Roney Berti de Oliveira, José A. M. Demattê and Marcos Rafael Nanni
Plants 2023, 12(13), 2526; https://doi.org/10.3390/plants12132526 - 2 Jul 2023
Cited by 6 | Viewed by 2100
Abstract
Hyperspectral technology offers significant potential for non-invasive monitoring and prediction of morphological parameters in plants. In this study, UV−VIS−NIR−SWIR reflectance hyperspectral data were collected from Nicotiana tabacum L. plants using a spectroradiometer. These plants were grown under different light and gibberellic acid (GA [...] Read more.
Hyperspectral technology offers significant potential for non-invasive monitoring and prediction of morphological parameters in plants. In this study, UV−VIS−NIR−SWIR reflectance hyperspectral data were collected from Nicotiana tabacum L. plants using a spectroradiometer. These plants were grown under different light and gibberellic acid (GA3) concentrations. Through spectroscopy and multivariate analyses, key growth parameters, such as height, leaf area, energy yield, and biomass, were effectively evaluated based on the interaction of light with leaf structures. The shortwave infrared (SWIR) bands, specifically SWIR1 and SWIR2, showed the strongest correlations with these growth parameters. When classifying tobacco plants grown under different GA3 concentrations in greenhouses, artificial intelligence (AI) and machine learning (ML) algorithms were employed, achieving an average accuracy of over 99.1% using neural network (NN) and gradient boosting (GB) algorithms. Among the 34 tested vegetation indices, the photochemical reflectance index (PRI) demonstrated the strongest correlations with all evaluated plant phenotypes. Partial least squares regression (PLSR) models effectively predicted morphological attributes, with R2CV values ranging from 0.81 to 0.87 and RPDP values exceeding 2.09 for all parameters. Based on Pearson’s coefficient XYZ interpolations and HVI algorithms, the NIR−SWIR band combination proved the most effective for predicting height and leaf area, while VIS−NIR was optimal for optimal energy yield, and VIS−VIS was best for predicting biomass. To further corroborate these findings, the SWIR bands for certain morphological characteristic wavelengths selected with s−PLS were most significant for SWIR1 and SWIR2, while i−PLS showed a more uniform distribution in VIS−NIR−SWIR bands. Therefore, SWIR hyperspectral bands provide valuable insights into developing alternative bands for remote sensing measurements to estimate plant morphological parameters. These findings underscore the potential of remote sensing technology for rapid, accurate, and non-invasive monitoring within stationary high-throughput phenotyping systems in greenhouses. These insights align with advancements in digital and precision technology, indicating a promising future for research and innovation in this field. Full article
(This article belongs to the Special Issue Integration of Spectroscopic and Photosynthetic Analyses in Plants)
Show Figures

Figure 1

Figure 1
<p>Box plot of the morphological and efficiency parameters in <span class="html-italic">Nicotiana tabacum</span> L. leaves of plants grown in high irradiance (yellow; full light) and low light (grey; 8.5% of full light; shade) environments and submitted to different gibberellin (gibberellic acid–GA<sub>3</sub>) concentration. (<b>A</b>) Height (cm). (<b>B</b>) Leaf area (m<sup>2</sup>). (<b>C</b>) Yield energetic (m<sup>3</sup>). (<b>D</b>) Biomass (g). CONT (Control; full light); GA10 (10 µM GA<sub>3</sub>; full light); GA100 (100 µM GA<sub>3</sub>; full light); PAC (50 mg L<sup>−1</sup> of paclobutrazol; full light); GA10P (combined GA<sub>3</sub> 10 µM + PAC; full light); GA100P (combined GA<sub>3</sub> 100 µM + PAC; full light); SCONT (Control; shade); SGA10 (10 µM GA<sub>3</sub>; shade); SGA100 (100 µM GA<sub>3</sub>; shade); SPAC (50 mg L<sup>−1</sup> of paclobutrazol; shade); SGA10P (combined GA<sub>3</sub> 10 µM + PAC; shade); SGA100P (combined GA<sub>3</sub> 100 µM + PAC; shade). Different letters above the box denote significant differences according to Duncan’s test (<span class="html-italic">p</span> &lt; 0.001). Mean ± SE. (<span class="html-italic">n</span> = 144).</p>
Full article ">Figure 2
<p>The factor of reflectance was calculated for spectral leaf curves spanning from 350 to 2500 nm in <span class="html-italic">Nicotiana tabacum</span> L. leaves. The plants were cultivated under two different light conditions, high irradiance (full light) and low light (8.5% of full light), and submitted to different concentrations of gibberellic acid (GA<sub>3</sub>). The determination of the accurate inflection points at 700 and 1300 nm was conducted (red dotted lines). Each repetition was generated by calculating the mean of measurements taken for leaves. For abbreviations pertaining to other treatments, see <a href="#plants-12-02526-f001" class="html-fig">Figure 1</a>. F-test (<span class="html-italic">p</span> &lt; 0.001). (<span class="html-italic">n</span> = 144). To enhance clarity, the standard error was omitted.</p>
Full article ">Figure 3
<p>Principal component analyses (PCAs) of spectral leaf reflectance curves ranging from 350 to 2500 nm in <span class="html-italic">Nicotiana tabacum</span> leaves from plants cultivated in high irradiance (yellow; full light) and low light (grey; 8.5% of full light) environments and submitted to different GA<sub>3</sub> concentrations. (<b>A</b>) 3D plot of the PCA scores for the PC1, PC2, and PC3 of hyperspectroscopy data. (<b>B</b>) Correlation of coefficients with three principal components (dark to light green lines; PC1, PC2, and PC3). The red line represents −0.70 and +0.70 correlation coefficients. For detailed information on treatments and their abbreviation, please refer to <a href="#plants-12-02526-f001" class="html-fig">Figure 1</a>. (<span class="html-italic">n</span> = 144).</p>
Full article ">Figure 4
<p>The cluster heatmap displayed the correlation between the hyperspectral bands for <span class="html-italic">Nicotiana tabacum</span> L. leaves from plants cultivated in high irradiance (full light) and low light (8.5% full light) environments and submitted to different GA<sub>3</sub> concentrations. The correlations are arranged by wavelength bands (VIS, NIR, SWIR), environment (sun, shade), gibberellin regimes (low, normal, high, very high), and leaf colors (very low, low, normal, dark). Positive correlations are shown in red, and negative correlations are shown in blue (Z-score, <span class="html-italic">p</span> &lt; 0.001). For detailed information on treatments and their abbreviation, please refer to <a href="#plants-12-02526-f001" class="html-fig">Figure 1</a>. (<span class="html-italic">n</span> = 144).</p>
Full article ">Figure 5
<p>Evaluation of models for precision and recall for 8 machine learning and artificial intelligence algorithms (AIAs) for <span class="html-italic">Nicotiana tabacum</span> L. leaves from plants cultivated in high irradiance (full light) and low light (8.5% of full light) environments and submitted to different GA<sub>3</sub> concentrations. Confusion matrix for neural network (NN), gradient boosting (GB), random forest (RF), supervision vector machine (SVM), K-nearest neighbors (KNN), naive Bayes (NB), logistic regression (LR), stochastic gradient descent (SGD). Inset shows LogLoss for error accumulation in models for classification. Boxes show overall accuracy/precision for correct classification (accepted in green) and mistake (error in red). A total of 100 training samples and 44 validation samples were used. For abbreviations V01−12, which indicate the order of treatments, see <a href="#plants-12-02526-f001" class="html-fig">Figure 1</a>. (<span class="html-italic">n</span> = 144).</p>
Full article ">Figure 6
<p>Relative contribution of narrowbands (VIs) to total variability for spectral leaf reflectance data from 350 to 2500 nm in <span class="html-italic">Nicotiana tabacum</span> L. leaves from plants cultivated in high irradiance (full light) and low light (8.5% of full light) environments and submitted to different GA<sub>3</sub> concentrations. For abbreviations of the vegetation indices, see <a href="#app1-plants-12-02526" class="html-app">Table S1</a>. Each vegetation index was separated for more correlation of specific differences in biochemical and structural compounds. Monitoring status (pink bars), photosynthetic pigments (yellow bars), photochemical efficiency (orange bars), water (blue bars), pigments (green bars), and structural (grey bars).</p>
Full article ">Figure 7
<p>Pearson’s correlation matrix between each of the morphological, efficiency, and vegetation indices parameters valuable for <span class="html-italic">Nicotiana tabacum</span> L. leaves from plants cultivated in high irradiance (full light) and low light (8.5% of full light) environments and submitted to different GA<sub>3</sub> concentrations. (<span class="html-italic">p</span> &lt; 0.001) Abbreviations are described in <a href="#app1-plants-12-02526" class="html-app">Table S1</a>.</p>
Full article ">Figure 8
<p>β−loadings and weighted-coefficient method-based reflectance hyperspectral data between 350 and 2500 nm for <span class="html-italic">Nicotiana tabacum</span> L. leaves from plants cultivated in high irradiance (full light) and low light (8.5% of full light) environments and submitted to different GA<sub>3</sub> concentrations. (<b>A</b>) Height (cm). (<b>B</b>) Leaf area (m<sup>2</sup>). (<b>C</b>) Yield energetic (m<sup>3</sup>). (<b>D</b>) Biomass (g).</p>
Full article ">Figure 9
<p>Count plot map by XYZ interpolates Pearson’s coefficient of correlation by inverse distance to a power math between morphological and efficiency parameters and wavelengths<sub>1</sub> vs. wavelength<sub>2</sub> for 350 to 2500 nm. (<b>A</b>) Height (cm). (<b>B</b>) Leaf area (m<sup>2</sup>). (<b>C</b>) Yield energetic (m<sup>3</sup>). (<b>D</b>) Biomass (g). White line, correlation 1:1. The displayed color gradient from dark blue to light red indicates an increase in associations.</p>
Full article ">Figure 10
<p>The most responsive variables were selected wavelength range of 350−2500 nm by the VIP, GA, <span class="html-italic">s</span>−PLS, <span class="html-italic">i</span>−PLS, <span class="html-italic">r</span>−PLS, and <span class="html-italic">n</span>−PLS algorithms for tobacco plant growth variables. (<b>A</b>) Height (cm). (<b>B</b>) Leaf area (m<sup>2</sup>). (<b>C</b>) Yield energetic (m<sup>3</sup>). (<b>D</b>) Biomass (g).</p>
Full article ">Figure 11
<p>The count plot map displays the coefficient of correlation (R<sup>2</sup>) obtained from linear regression analysis between morphological and phenotyping parameters and wavelengths<sub>1</sub> vs. wavelength<sub>2</sub> in the range of 350 to 2500 nm. (<b>A</b>) Height (cm). (<b>B</b>) Leaf area (m<sup>2</sup>). (<b>C</b>) Yield energetic (m<sup>3</sup>). (<b>D</b>) Biomass (g). White line, correlation 1:1. The displayed color gradient from dark blue to light red indicates an increase in associations.</p>
Full article ">
Back to TopTop