[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 (160)

Search Parameters:
Keywords = precision viticulture

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 362 KiB  
Review
Satellite Solutions for Precision Viticulture: Enhancing Sustainability and Efficiency in Vineyard Management
by Ana Mucalo, Damir Matić, Antonio Morić-Španić and Marin Čagalj
Agronomy 2024, 14(8), 1862; https://doi.org/10.3390/agronomy14081862 - 22 Aug 2024
Viewed by 545
Abstract
The priority problem in intensive viticulture is reducing pesticides, and fertilizers, and improving water-use efficiency. This is driven by global and EU regulatory efforts. This review, systematically examines 92 papers, focusing on progress in satellite solutions over time, and (pre)processing improvements of spatio-temporal [...] Read more.
The priority problem in intensive viticulture is reducing pesticides, and fertilizers, and improving water-use efficiency. This is driven by global and EU regulatory efforts. This review, systematically examines 92 papers, focusing on progress in satellite solutions over time, and (pre)processing improvements of spatio-temporal and spectral resolution. The importance of the integration of satellites with ground truth data is highlighted. The results provide precise on-field adaptation strategies through the generation of prescription maps and variable rate application. This enhances sustainability and efficiency in vineyard management and reduces the environmental footprint of vineyard techniques. The effectiveness of different vegetation indices in capturing spatial and temporal variations in vine health, water content, chlorophyll levels, and overall vigor is discussed. The challenges in the use of satellite data in viticulture are addressed. Advanced satellite technologies provide detailed vineyard monitoring, offering insights into spatio-temporal variability, soil moisture, and vine health. These are crucial for optimizing water-use efficiency and targeted management practices. By integrating satellite data with ground-based measurements, viticulturists can enhance precision viticulture, reduce reliance on chemical interventions, and improve overall vineyard sustainability and productivity. Full article
(This article belongs to the Special Issue Precision Viticulture for Vineyard Management)
21 pages, 2709 KiB  
Article
Integrating Spectral Sensing and Systems Biology for Precision Viticulture: Effects of Shade Nets on Grapevine Leaves
by Renan Tosin, Igor Portis, Leandro Rodrigues, Igor Gonçalves, Catarina Barbosa, Jorge Teixeira, Rafael J. Mendes, Filipe Santos, Conceição Santos, Rui Martins and Mário Cunha
Horticulturae 2024, 10(8), 873; https://doi.org/10.3390/horticulturae10080873 - 18 Aug 2024
Viewed by 661
Abstract
This study investigates how grapevines (Vitis vinifera L.) respond to shading induced by artificial nets, focusing on physiological and metabolic changes. Through a multidisciplinary approach, grapevines’ adaptations to shading are presented via biochemical analyses and hyperspectral data that are then combined with [...] Read more.
This study investigates how grapevines (Vitis vinifera L.) respond to shading induced by artificial nets, focusing on physiological and metabolic changes. Through a multidisciplinary approach, grapevines’ adaptations to shading are presented via biochemical analyses and hyperspectral data that are then combined with systems biology techniques. In the study, conducted in a ‘Moscatel Galego Branco’ vineyard in Portugal’s Douro Wine Region during post-veraison, shading was applied and predawn leaf water potential (Ψpd) was then measured to assess water stress. Biochemical analyses and hyperspectral data were integrated to explore adaptations to shading, revealing higher chlorophyll levels (chlorophyll a-b 117.39% higher) and increased Reactive Oxygen Species (ROS) levels in unshaded vines (52.10% higher). Using a self-learning artificial intelligence algorithm (SL-AI), simulations highlighted ROS’s role in stress response and accurately predicted chlorophyll a (R2: 0.92, MAPE: 24.39%), chlorophyll b (R2: 0.96, MAPE: 17.61%), and ROS levels (R2: 0.76, MAPE: 52.17%). In silico simulations employing flux balance analysis (FBA) elucidated distinct metabolic phenotypes between shaded and unshaded vines across cellular compartments. Integrating these findings provides a systems biology approach for understanding grapevine responses to environmental stressors. The leveraging of advanced omics technologies and precise metabolic models holds immense potential for untangling grapevine metabolism and optimizing viticultural practices for enhanced productivity and quality. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Representation of an integrated methodology combining the genomics, metabolomics, and systems biology approaches, aimed at establishing connections between laboratory experiments and field observations. Also, represents the integration of sensors for detecting molecular components and monitoring the physiological state of plants to feed the systems biology.</p>
Full article ">Figure 2
<p>Mean spectra absorbance from leaves on vines exposed to unshaded conditions and shaded conditions. a.u.: arbitrary units.</p>
Full article ">Figure 3
<p>Predawn leaf water potential variation in grapevine leaves under unshaded and shaded conditions. * Statistically significant (<span class="html-italic">p</span> &lt; 0.05) according to <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>Results of the use of foliar spectral data combined with self-learning artificial intelligence (SL-AI) in modelling: (<b>a</b>) chlorophyll <span class="html-italic">a</span> (mg/gFM), (<b>b</b>) chlorophyll <span class="html-italic">b</span> (mg/gFM), and (<b>c</b>) reactive oxygen species (ROS—ABS/gFM). N = number of samples considered. Coefficient of determination (R<sup>2</sup>), root mean square error (RMSE), and mean absolute percentage error (MAPE—%).</p>
Full article ">Figure 5
<p>Photorespiration cycle under (<b>a</b>) shaded and (<b>b</b>) unshaded conditions, showcasing the Photorespiration III hypothesis with the objective reactions in the flux balance analyses of Peroxisomal Catalase and Peroxisomal Glycolate Oxidase. The red and blue arrows represent respectively the flux balance analysis (FBA) of the reactions involved under shaded and unshaded conditions, offering a dynamic regulation of photorespiratory metabolism in response to environmental light conditions. The arrows’ sizes indicates each pathway’s intensity under the respective condition. Figure adapted from Huma, Kundu, Poolman, Kruger and Fell [<a href="#B46-horticulturae-10-00873" class="html-bibr">46</a>].</p>
Full article ">Figure 6
<p>Panel (<b>a</b>) shows the phenotype spaces for the three Flux Balance Analysis (FBA) models under different photorespiration conditions. For unshaded conditions, FBA models include FBA_P_CT_U (Peroxisomal Catalase Reaction as objective), FBA_P_96_U (Peroxisomal Glycolate Oxidase Reaction), and FBA_B3_U (Peroxisomal Catalase Reaction as objective and Peroxisomal Glycolate Oxidase Reaction as objective). Similarly, shaded conditions are represented by FBA_P_CT_S, FBA_P_96_S and FBA_B3_S, respectively. Panel (<b>b</b>) demonstrates the phenotype spaces resulting from Monte Carlo (MC) simulations, illustrating the range of variations in chlorophyll and reactive oxygen species (ROS) levels assessed in laboratory experiments. MC simulations encompass MC_B3 (Peroxisomal Catalase Reaction as objective and Peroxisomal Glycolate Oxidase Reaction as objective), MC_P_96 (Peroxisomal Glycolate Oxidase Reaction as objective), and MC_P_CT (Peroxisomal Catalase Reaction as objective). In the case of MC_P_96 and MC_P_CT, 1 represents higher ROS (unshaded condition) and 10, lower ROS (shaded condition). MC_B3 did not result in a valid MC simulation.</p>
Full article ">
24 pages, 5084 KiB  
Article
Comparative Analysis of TLS and UAV Sensors for Estimation of Grapevine Geometric Parameters
by Leilson Ferreira, Joaquim J. Sousa, José. M. Lourenço, Emanuel Peres, Raul Morais and Luís Pádua
Sensors 2024, 24(16), 5183; https://doi.org/10.3390/s24165183 - 11 Aug 2024
Viewed by 824
Abstract
Understanding geometric and biophysical characteristics is essential for determining grapevine vigor and improving input management and automation in viticulture. This study compares point cloud data obtained from a Terrestrial Laser Scanner (TLS) and various UAV sensors including multispectral, panchromatic, Thermal Infrared (TIR), RGB, [...] Read more.
Understanding geometric and biophysical characteristics is essential for determining grapevine vigor and improving input management and automation in viticulture. This study compares point cloud data obtained from a Terrestrial Laser Scanner (TLS) and various UAV sensors including multispectral, panchromatic, Thermal Infrared (TIR), RGB, and LiDAR data, to estimate geometric parameters of grapevines. Descriptive statistics, linear correlations, significance using the F-test of overall significance, and box plots were used for analysis. The results indicate that 3D point clouds from these sensors can accurately estimate maximum grapevine height, projected area, and volume, though with varying degrees of accuracy. The TLS data showed the highest correlation with grapevine height (r = 0.95, p < 0.001; R2 = 0.90; RMSE = 0.027 m), while point cloud data from panchromatic, RGB, and multispectral sensors also performed well, closely matching TLS and measured values (r > 0.83, p < 0.001; R2 > 0.70; RMSE < 0.084 m). In contrast, TIR point cloud data performed poorly in estimating grapevine height (r = 0.76, p < 0.001; R2 = 0.58; RMSE = 0.147 m) and projected area (r = 0.82, p < 0.001; R2 = 0.66; RMSE = 0.165 m). The greater variability observed in projected area and volume from UAV sensors is related to the low point density associated with spatial resolution. These findings are valuable for both researchers and winegrowers, as they support the optimization of TLS and UAV sensors for precision viticulture, providing a basis for further research and helping farmers select appropriate technologies for crop monitoring. Full article
(This article belongs to the Section Smart Agriculture)
Show Figures

Figure 1

Figure 1
<p>Overview of studied vineyard plot subjected to unmanned aerial vehicle data acquisition (<b>a</b>) and the area scanned by terrestrial laser scanner along with the studied grapevines (<b>b</b>).</p>
Full article ">Figure 2
<p>Equipment used and data acquisition tasks: (<b>a</b>) BLK360 G1; (<b>b</b>) Matrice 300 RTK; (<b>c</b>) acquisition of coordinates from deployed targets; (<b>d</b>) D-RTK 2 high precision GNSS mobile station; and (<b>e</b>) measurement of grapevine height.</p>
Full article ">Figure 3
<p>View of ground control points in the different imagery data types acquired by the sensors on the unmanned aerial vehicle. NIR: near-infrared; PAN: panchromatic; TIR: thermal infrared.</p>
Full article ">Figure 4
<p>Comparative view of the point clouds generated from the different sensors in part of a grapevine row in the study area. TLS: terrestrial laser scanner; MSP: multispectral; PAN: panchromatic; TIR: thermal infrared.</p>
Full article ">Figure 5
<p>Top perspective of the point distribution along the grapevine canopy from point clouds generated by different sensors in part of a grapevine row. TLS: terrestrial laser scanner; MSP: multispectral; PAN: panchromatic; TIR: thermal infrared.</p>
Full article ">Figure 6
<p>Spectral and thermal behavior of acquired data in different vineyard elements (grapevine, bare soil, other vegetation): (<b>a</b>) visual representation of the raster products, (<b>b</b>) reflectance in the five spectral bands, and (<b>c</b>) mean values of normalized difference vegetation index (NDVI) and land surface temperature (LST).</p>
Full article ">Figure 7
<p>Box plot distribution of geometrical parameters of the analyzed grapevines obtained using different measurement methods: (<b>a</b>) height metrics, (<b>b</b>) projected area, and (<b>c</b>) canopy volume.</p>
Full article ">Figure 8
<p>Correlation matrix between height variables obtained from different sensors for (<b>a</b>) maximum height and field-measured height; (<b>b</b>) height at the 95th percentile; and (<b>c</b>) height at the 90th percentile. The diagonal line is intentionally omitted.</p>
Full article ">Figure 9
<p>Correlation matrix for grapevine projected area (<b>a</b>) and canopy volume (<b>b</b>) from the different sensors. The diagonal line is intentionally omitted.</p>
Full article ">
17 pages, 3815 KiB  
Article
Site-Independent Mapping of Clay Content in Vineyard Soils via Mobile Proximal Gamma-Ray Spectrometry and Machine Learning Calibrations
by Ralf Wehrle and Stefan Pätzold
Sensors 2024, 24(14), 4528; https://doi.org/10.3390/s24144528 - 12 Jul 2024
Viewed by 506
Abstract
Vineyards hold considerable soil variability between regions and plots, and there is frequently large soil heterogeneity within plots. Clay content in vineyard soils is of interest with respect to soil management, environmental monitoring, and wine quality. However, spatially resolved clay mapping is laborious [...] Read more.
Vineyards hold considerable soil variability between regions and plots, and there is frequently large soil heterogeneity within plots. Clay content in vineyard soils is of interest with respect to soil management, environmental monitoring, and wine quality. However, spatially resolved clay mapping is laborious and expensive. Gamma-ray spectrometry (GS) is a suitable tool for predicting clay content in precision agriculture when locally calibrated, but it has scarcely been tested site-independently and in vineyards. This study evaluated GS to predict clay content with a site-independent calibration and four machine learning algorithms (Support Vector Machines, Random Forest, k-Nearest Neighbors, and Bayesian regulated neuronal networks) in eight vineyards from four German vine-growing regions. Clay content in the studied soils ranged from 62 to 647 g kg−1. The Random Forest calibration was most suitable. Test set evaluation revealed good model performance for the entire dataset with RPIQ = 4.64, RMSEP = 56.7 g kg−1, and R2 = 0.87; however, prediction quality varied between the sites. Overall, GS with the Random Forest model calibration was appropriate to predict the clay content and its spatial distribution, even for heterogeneous geopedological settings and in individual plots. Therefore, GS is considered a valuable tool for soil mapping in vineyards, where clay content and product quality are closely linked. Full article
(This article belongs to the Special Issue Precision Agriculture and Sensor Systems—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Tractor-mounted gamma spectrometer.</p>
Full article ">Figure 2
<p>Clay content (g kg<sup>−1</sup>) versus gamma ROIs for the entire dataset: (<b>a</b>) total counts (cps); (<b>b</b>) K-40 (cps); (<b>c</b>) Th-232 (cps); (<b>d</b>) Th-232/K-40-ratios.</p>
Full article ">Figure 3
<p>Predicted and observed values of cross-validation (cal) and test set validation (val) for clay content from site-independent regression models. SVM: Support Vector Machine; KNN: k-Nearest Neighbor, BNN: Bayesian regularized neuronal network; RF: Random Forest.</p>
Full article ">Figure 4
<p>Prediction results of the site-independent RF calibration model when separately applied to individual vineyards.</p>
Full article ">Figure 5
<p>(<b>a</b>) Total counts (cps) of on-the-go gamma spectrometric measurements and ground truth sampling points of the examined vineyards. (<b>b</b>) Total counts (cps) of on-the-go gamma spectrometric measurements and ground truth sampling points of the examined vineyards.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) Total counts (cps) of on-the-go gamma spectrometric measurements and ground truth sampling points of the examined vineyards. (<b>b</b>) Total counts (cps) of on-the-go gamma spectrometric measurements and ground truth sampling points of the examined vineyards.</p>
Full article ">Figure 6
<p>Gamma spectrometric on-the-go predicted soil clay maps of Random Forest model and ground truth data (large dots) for the vineyards in (<b>a</b>) Leiw H, (<b>b</b>) Leiw K, (<b>c</b>) Sieb N, and (<b>d</b>) Spre B.</p>
Full article ">
22 pages, 6267 KiB  
Article
Long-Term Evolution of the Climatic Factors and Its Influence on Grape Quality in Northeastern Romania
by Roxana Mihaela Filimon, Claudiu Ioan Bunea, Răzvan Vasile Filimon, Florin Dumitru Bora and Doina Damian
Horticulturae 2024, 10(7), 705; https://doi.org/10.3390/horticulturae10070705 - 3 Jul 2024
Viewed by 530
Abstract
Climate change is currently the greatest threat to the environment as we know it today. The present study aimed to highlight the changes in the main climatic elements during the last five decades (1971–2020) in northeastern Romania (Copou-Iaşi wine-growing center) and their impact [...] Read more.
Climate change is currently the greatest threat to the environment as we know it today. The present study aimed to highlight the changes in the main climatic elements during the last five decades (1971–2020) in northeastern Romania (Copou-Iaşi wine-growing center) and their impact on grape quality, as part of precision viticulture strategies and efficient management of grapevine plantations. Data analysis revealed a constant and significant increase in the average air temperature in the last 50 years (+1.70 °C), more pronounced in the last 10 years (+0.61 °C), with a number of days with extreme temperatures (>30 °C) of over 3.5-fold higher, in parallel with a fluctuating precipitation regime. The increase in average temperatures in the last 40 years was highly correlated with the advancement of the grape harvest date (up to 12 days), a significant increase in Vitis vinifera L. white grape sugar concentration (+15–25 g/L), and a drastic decrease in total acidity (−2.0–3.5 g/L tartaric acid). The significant increase in the values of the bioclimatic indices require the reclassification of the wine-growing area in higher classes of favorability, raising the opportunity to grow cultivars that are more suited to warmer climates, ensuring the efficiency of the plantation, and meeting current consumer expectations. Full article
(This article belongs to the Special Issue Orchard Management under Climate Change)
Show Figures

Figure 1

Figure 1
<p>The location of the Copou-Iaşi wine-growing center (NE of Romania). Source: Google Earth [<a href="#B28-horticulturae-10-00705" class="html-bibr">28</a>].</p>
Full article ">Figure 2
<p>Evolution of annual average temperatures (<b>a</b>) and growing season average temperature (<b>b</b>) in the Copou-Iaşi wine-growing center, NE of Romania (1971–2020). Note: The mean values of the decades are presented as the average of the annual data (n = 10) with standard deviation (±). Values with the same letter are not statistically significant (<span class="html-italic">p</span> &gt; 0.05) using Tukey’s test.</p>
Full article ">Figure 3
<p>Changes in the annual average precipitation (<b>a</b>) and growing season average precipitation (April–September) (<b>b</b>) in the Copou-Iaşi wine-growing center, NE of Romania (1971–2020). Note: The mean values of the decades are presented as the average of the annual data (n = 10) with standard deviation (±). Values with the same letter are not statistically significant (<span class="html-italic">p</span> &gt; 0.05) using Tukey’s test.</p>
Full article ">Figure 4
<p>The average number of days with temperatures above 30 °C (<b>a</b>) and above 35 °C (<b>b</b>) in the Copou-Iaşi wine-growing center, NE of Romania (1971–2020). Note: The mean values of the decades are presented as the average of the annual data (n = 10) with standard deviation (±). Values with the same letter are not statistically significant (<span class="html-italic">p</span> &gt; 0.05) using Tukey’s test.</p>
Full article ">Figure 5
<p>The average number of days with temperatures below −15 °C (December, January, and February; winter frost) (<b>a</b>) and below −2 °C (March, April, May; spring frost) (<b>b</b>) in the Copou-Iaşi wine-growing center, NE of Romania (1971–2020). Note: The mean values of the decades are presented as the average of the annual data (n = 10), with standard deviation (±). Values with the same letter are not statistically significant (<span class="html-italic">p</span> &gt; 0.05) using Tukey’s test.</p>
Full article ">Figure 6
<p>The comparative presentation of the grape harvest intervals (by decade), in the Copou-Iasi wine-growing center, NE of Romania (1981–2020). Note: The decade average was calculated as the mean value of the annual data (n = 10), for each cultivar.</p>
Full article ">Figure 7
<p>Changes in sugar amount (g/L) and total acidity (g/L as tartaric acid) in mature grapes of Fetească Albă (<b>a</b>), Fetească Regală (<b>b</b>), Aligoté (<b>c</b>), and Muscat Ottonel (<b>d</b>) cultivars in the period 1981–2020, in the Copou-Iaşi wine-growing center (NE of Romania). Note: The mean values of the decades are presented as the average of the annual data (n = 10), with standard deviation (±). Values with the same letter are not statistically significant (<span class="html-italic">p</span> &gt; 0.05) using Tukey’s test.</p>
Full article ">Figure 7 Cont.
<p>Changes in sugar amount (g/L) and total acidity (g/L as tartaric acid) in mature grapes of Fetească Albă (<b>a</b>), Fetească Regală (<b>b</b>), Aligoté (<b>c</b>), and Muscat Ottonel (<b>d</b>) cultivars in the period 1981–2020, in the Copou-Iaşi wine-growing center (NE of Romania). Note: The mean values of the decades are presented as the average of the annual data (n = 10), with standard deviation (±). Values with the same letter are not statistically significant (<span class="html-italic">p</span> &gt; 0.05) using Tukey’s test.</p>
Full article ">Figure 8
<p>Principal Component Analysis (PCA) biplot combining the output variables (<b>a</b>) and the Agglomerative Hierarchical Clustering (AHC) of the decades (<b>b</b>) for the interval 1981–2020, in the Copou-Iaşi wine-growing center. Note: FA—Fetească Albă cv.; FR—Fetească Regală cv.; Alig.—Aligoté; MO—Muscat Ottonel cv.; GS—growing season; T—temperature; Harvest—harvest date; Σt°u—the sum of active temperatures; HC—hydrothermal coefficient; IDM—De Martonne aridity index; IHr—actual heliotermal index; Ibcv—grapevine bioclimatic index; IAOe—oenoclimate aptitude index; HI—Huglin index; Wi—Winkler index; CNI—cool night index.</p>
Full article ">
22 pages, 10908 KiB  
Article
Synthetic Aperture Radar in Vineyard Monitoring: Examples, Demonstrations, and Future Perspectives
by Matus Bakon, Ana Cláudia Teixeira, Luís Pádua, Raul Morais, Juraj Papco, Lukas Kubica, Martin Rovnak, Daniele Perissin and Joaquim J. Sousa
Remote Sens. 2024, 16(12), 2106; https://doi.org/10.3390/rs16122106 - 11 Jun 2024
Viewed by 1081
Abstract
Synthetic aperture radar (SAR) technology has emerged as a pivotal tool in viticulture, offering unique capabilities for various applications. This study provides a comprehensive overview of the current state-of-the-art applications of SAR in viticulture, highlighting its significance in addressing key challenges and enhancing [...] Read more.
Synthetic aperture radar (SAR) technology has emerged as a pivotal tool in viticulture, offering unique capabilities for various applications. This study provides a comprehensive overview of the current state-of-the-art applications of SAR in viticulture, highlighting its significance in addressing key challenges and enhancing viticultural practices. The historical evolution and motivations behind SAR technology are also provided, along with a demonstration of its applications within viticulture, showcasing its effectiveness in various aspects of vineyard management, including delineating vineyard boundaries, assessing grapevine health, and optimizing irrigation strategies. Furthermore, future perspectives and trends in SAR applications in viticulture are discussed, including advancements in SAR technology, integration with other remote sensing techniques, and the potential for enhanced data analytics and decision support systems. Through this article, a comprehensive understanding of the role of SAR in viticulture is provided, along with inspiration for future research endeavors in this rapidly evolving field, contributing to the sustainable development and optimization of vineyard management practices. Full article
(This article belongs to the Special Issue Applications of SAR for Environment Observation Analysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geometry of side-looking radar (adapted from [<a href="#B1-remotesensing-16-02106" class="html-bibr">1</a>,<a href="#B2-remotesensing-16-02106" class="html-bibr">2</a>]).</p>
Full article ">Figure 2
<p>Time series of precipitation and Sentinel-1A <math display="inline"><semantics> <msubsup> <mi>σ</mi> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mn>0</mn> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mi>σ</mi> <mrow> <mi>V</mi> <mi>H</mi> </mrow> <mn>0</mn> </msubsup> </semantics></math> backscatter responses for four vineyards in the Douro wine region.</p>
Full article ">Figure 3
<p>Time series of Sentinel-1A <math display="inline"><semantics> <msubsup> <mi>σ</mi> <mrow> <mi>V</mi> <mi>H</mi> </mrow> <mn>0</mn> </msubsup> </semantics></math> backscatter and CR for two vineyards (vineyards 3 and 4).</p>
Full article ">Figure 4
<p>Comparison of predicted and sensor soil moisture values.</p>
Full article ">Figure 5
<p>Mean line-of-sight (LOS) velocities from Sentinel-1A/B’s ascending track No. 147 in the Douro Demarcated Region.</p>
Full article ">Figure 6
<p>Examples from persistent scatterer interferometry (PSI) deformation monitoring of vineyard sites (<b>b</b>,<b>d</b>) and surrounding infrastructure (<b>a</b>,<b>c</b>).</p>
Full article ">
29 pages, 23398 KiB  
Article
Classification of Grapevine Varieties Using UAV Hyperspectral Imaging
by Alfonso López, Carlos J. Ogayar, Francisco R. Feito and Joaquim J. Sousa
Remote Sens. 2024, 16(12), 2103; https://doi.org/10.3390/rs16122103 - 10 Jun 2024
Viewed by 665
Abstract
Classifying grapevine varieties is crucial in precision viticulture, as it allows for accurate estimation of vineyard row growth for different varieties and ensures authenticity in the wine industry. This task can be performed with time-consuming destructive methods, including data collection and analysis in [...] Read more.
Classifying grapevine varieties is crucial in precision viticulture, as it allows for accurate estimation of vineyard row growth for different varieties and ensures authenticity in the wine industry. This task can be performed with time-consuming destructive methods, including data collection and analysis in the laboratory. In contrast, unmanned aerial vehicles (UAVs) offer a markedly more efficient and less restrictive method for gathering hyperspectral data, even though they may yield data with higher levels of noise. Therefore, the first task is the processing of these data to correct and downsample large amounts of data. In addition, the hyperspectral signatures of grape varieties are very similar. In this study, we propose the use of a convolutional neural network (CNN) to classify seventeen different varieties of red and white grape cultivars. Instead of classifying individual samples, our approach involves processing samples alongside their surrounding neighborhood for enhanced accuracy. The extraction of spatial and spectral features is addressed with (1) a spatial attention layer and (2) inception blocks. The pipeline goes from data preparation to dataset elaboration, finishing with the training phase. The fitted model is evaluated in terms of response time, accuracy and data separability and is compared with other state-of-the-art CNNs for classifying hyperspectral data. Our network was proven to be much more lightweight by using a limited number of input bands (40) and a reduced number of trainable weights (560 k parameters). Hence, it reduced training time (1 h on average) over the collected hyperspectral dataset. In contrast, other state-of-the-art research requires large networks with several million parameters that require hours to be trained. Despite this, the evaluated metrics showed much better results for our network (approximately 99% overall accuracy), in comparison with previous works barely achieving 81% OA over UAV imagery. This notable OA was similarly observed over satellite data. These results demonstrate the efficiency and robustness of our proposed method across different hyperspectral data sources. Full article
(This article belongs to the Section Engineering Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Overview of the areas surveyed using UAV hyperspectral imaging for the classification task. Two different vineyard crops are depicted according to their main variety: (<b>a</b>) red and (<b>b</b>) white.</p>
Full article ">Figure 2
<p>Conversion of (<b>a</b>) hyperspectral DNs into (<b>b</b>) reflectance using white and dark references. The three grey levels are sampled in (<b>a</b>).</p>
Full article ">Figure 3
<p>Three-dimensional distribution of hyperspectral samples from 17 classes, obtained by narrowing 50 components calculated with PCA into three components estimated by uMAP.</p>
Full article ">Figure 4
<p>Results of experiments conducted to compare different feature transformation algorithms with different numbers of components. PCA, FA, NMF and LSA (truncated SVD) are evaluated using the DSI metric and the OA obtained by training an SVM model. The default DSI and accuracy are obtained from the original data with 140 features.</p>
Full article ">Figure 5
<p>Workflow for manually labelling HSI swaths. First, the false RGB image is displayed. Then, the NDVI is extracted, followed by thresholding and marking with polygons using the Sensarea software. Finally, a Boolean operation, <span class="html-italic">AND</span>, is performed between the polygon and binary masks to obtain the final labelled regions.</p>
Full article ">Figure 6
<p>Overview of dataset preparation. First, a binary mask was generated using the NDVI and rows were organized into different groups to distinguish vineyard classes. Once pixels were processed as described in <a href="#sec3dot3-remotesensing-16-02103" class="html-sec">Section 3.3</a>, both reflectance and labels were split into patches. The signatures on the right side show the original and transformed reflectance, including the variance per feature. Blue lines show the averaged ground spectral signature, whereas orange represents the pixels labelled as vegetation.</p>
Full article ">Figure 7
<p>Scheme of the proposed CNN, highlighting four different parts as well as the structure of inception blocks.</p>
Full article ">Figure 8
<p>From top to bottom: initial distribution of samples per label and after using the proposed narrowing, with only three groups being downsampled.</p>
Full article ">Figure 9
<p>Confusion matrix for classifying red and white varieties altogether.</p>
Full article ">Figure 10
<p>Training and validation accuracy and loss during training.</p>
Full article ">Figure 11
<p>Response time for training the network (left axis) as well as the number of parameters (right axis) for every compared network, including ours [<a href="#B59-remotesensing-16-02103" class="html-bibr">59</a>].</p>
Full article ">Figure 12
<p>Clustering of samples according to the feature transformation performed by uMAP over (<b>a</b>) the starting hyperspectral features and (<b>b</b>) features extracted by the CNN before transferring it to the final softmax layer.</p>
Full article ">Figure 13
<p>Overall accuracy obtained for patches of different sizes, from 5 to 31.</p>
Full article ">Figure 14
<p>Training time (in minutes) and the number of parameters as the window size increases. The blue bars correspond to the left axis, representing training time, while the orange bars represent the number of parameters.</p>
Full article ">Figure 15
<p>Errors observed in the classification of red varieties and the hyperspectral signature of a few samples concerning different surfaces. FP indicates a false-positive sample mistakenly labelled as vegetation during dataset preparation, since it reflects a human-made structure in the false RGB image. It led to a few prediction errors in close grapevine samples. However, it is not trivial to mask them out during dataset preparation.</p>
Full article ">Figure 16
<p>OA observed by training the proposed CNN network with a subset (percentage) of the training dataset. The dashed line represents the expected results for intermediate values.</p>
Full article ">
19 pages, 10565 KiB  
Article
Detection of Leak Areas in Vineyard Irrigation Systems Using UAV-Based Data
by Luís Pádua, Pedro Marques, Lia-Tânia Dinis, José Moutinho-Pereira, Joaquim J. Sousa, Raul Morais and Emanuel Peres
Drones 2024, 8(5), 187; https://doi.org/10.3390/drones8050187 - 8 May 2024
Viewed by 1688
Abstract
Water is essential for maintaining plant health and optimal growth in agriculture. While some crops depend on irrigation, others can rely on rainfed water, depending on regional climatic conditions. This is exemplified by grapevines, which have specific water level requirements, and irrigation systems [...] Read more.
Water is essential for maintaining plant health and optimal growth in agriculture. While some crops depend on irrigation, others can rely on rainfed water, depending on regional climatic conditions. This is exemplified by grapevines, which have specific water level requirements, and irrigation systems are needed. However, these systems can be susceptible to damage or leaks, which are not always easy to detect, requiring meticulous and time-consuming inspection. This study presents a methodology for identifying potential damage or leaks in vineyard irrigation systems using RGB and thermal infrared (TIR) imagery acquired by unmanned aerial vehicles (UAVs). The RGB imagery was used to distinguish between grapevine and non-grapevine pixels, enabling the division of TIR data into three raster products: temperature from grapevines, from non-grapevine areas, and from the entire evaluated vineyard plot. By analyzing the mean temperature values from equally spaced row sections, different threshold values were calculated to estimate and map potential leaks. These thresholds included the lower quintile value, the mean temperature minus the standard deviation (Tmeanσ), and the mean temperature minus two times the standard deviation (Tmean2σ). The lower quintile threshold showed the best performance in identifying known leak areas and highlighting the closest rows that need inspection in the field. This approach presents a promising solution for inspecting vineyard irrigation systems. By using UAVs, larger areas can be covered on-demand, improving the efficiency and scope of the inspection process. This not only reduces water wastage in viticulture and eases grapevine water stress but also optimizes viticulture practices. Full article
(This article belongs to the Special Issue Advances of UAV in Precision Agriculture)
Show Figures

Figure 1

Figure 1
<p>Overview of the vineyard plots analyzed along with land surface temperature data of Vineyard A (<b>a</b>) and Vineyard B (<b>b</b>).</p>
Full article ">Figure 2
<p>Overview of the data processing pipeline employed for the detection of leaks in vineyard irrigation systems. TIR: thermal infrare; DSM: digital surface model; DTM: digital terrain model; LST: land surface temperature; <math display="inline"><semantics> <msub> <mi>G</mi> <mi>n</mi> </msub> </semantics></math>: normalized green value; CSM: crop surface model.</p>
Full article ">Figure 3
<p>Land surface temperature (LST) behavior in a cross-row profiles in Vineyard A (<b>a</b>) and Vineyard B (<b>b</b>). Height values from the crop surface model (CSM) are also provided. Gray areas correspond grapevine shadows.</p>
Full article ">Figure 4
<p>Temperature distribution of Vineyard A and Vineyard B when considering temperature from the entire vineyard area, grapevine vegetation, and non-grapevine areas.</p>
Full article ">Figure 5
<p>Row section histograms for Vineyard A (<b>a</b>) and Vineyard B (<b>b</b>) when considering the entire vineyard area, grapevine temperature only, and non-grapevine areas.</p>
Full article ">Figure 6
<p>Temperature distribution classes in Vineyard A (<b>a</b>) and Vineyard B (<b>b</b>) categorized by the entire vineyard area, grapevine temperature exclusively, and non-grapevine areas.</p>
Full article ">Figure 7
<p>Areas of potential leaks in Vineyard A when using (<b>a</b>) the lower quintile value, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>m</mi> <mi>e</mi> <mi>a</mi> <mi>n</mi> </mrow> </msub> <mo>−</mo> <mi>σ</mi> </mrow> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>m</mi> <mi>e</mi> <mi>a</mi> <mi>n</mi> </mrow> </msub> <mo>−</mo> <mn>2</mn> <mi>σ</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Areas of potential leaks in Vineyard B when using (<b>a</b>) the lower quintile value, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>m</mi> <mi>e</mi> <mi>a</mi> <mi>n</mi> </mrow> </msub> <mo>−</mo> <mi>σ</mi> </mrow> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>m</mi> <mi>e</mi> <mi>a</mi> <mi>n</mi> </mrow> </msub> <mo>−</mo> <mn>2</mn> <mi>σ</mi> </mrow> </semantics></math>.</p>
Full article ">
21 pages, 64151 KiB  
Article
A Cobot in the Vineyard: Computer Vision for Smart Chemicals Spraying
by Claudio Tomazzoli, Andrea Ponza, Matteo Cristani, Francesco Olivieri and Simone Scannapieco
Appl. Sci. 2024, 14(9), 3777; https://doi.org/10.3390/app14093777 - 28 Apr 2024
Viewed by 1019
Abstract
Precision agriculture (PA) is a management concept that makes use of digital techniques to monitor and optimise agricultural production processes and represents a field of growing economic and social importance. Within this area of knowledge, there is a topic not yet fully explored: [...] Read more.
Precision agriculture (PA) is a management concept that makes use of digital techniques to monitor and optimise agricultural production processes and represents a field of growing economic and social importance. Within this area of knowledge, there is a topic not yet fully explored: outlining a road map towards the definition of an affordable cobot solution (i.e., a low-cost robot able to safely coexist with humans) able to perform automatic chemical treatments. The present study narrows its scope to viticulture technologies, and targets small/medium-sized winemakers and producers, for whom innovative technological advancements in the production chain are often precluded by financial factors. The aim is to detail the realization of such an integrated solution and to discuss the promising results achieved. The results of this study are: (i) The definition of a methodology for integrating a cobot in the process of grape chemicals spraying under the constraints of a low-cost apparatus; (ii) the realization of a proof-of-concept of such a cobotic system; (iii) the experimental analysis of the visual apparatus of this system in an indoor and outdoor controlled environment as well as in the field. Full article
(This article belongs to the Special Issue Application of Machine Learning in Industry 4.0)
Show Figures

Figure 1

Figure 1
<p>The orchestrator allows hardware and software intercommunication.</p>
Full article ">Figure 2
<p>Franka Emika Panda’s end effector. Source: <a href="https://pkj-robotics.dk/wp-content/uploads/2020/09/Franka_Emika_Hand_01.jpg" target="_blank">https://pkj-robotics.dk/wp-content/uploads/2020/09/Franka_Emika_Hand_01.jpg</a>, accessed on 25 April 2024.</p>
Full article ">Figure 3
<p>Example of a sprayer end effector. Source: <a href="https://www.researchgate.net/publication/313497501/figure/fig18/AS:614364813983767@1523487394998/Variable-color-output-from-the-foam-spray-end-effector-is-seen.png" target="_blank">https://www.researchgate.net/publication/313497501/figure/fig18/AS:614364813983767@1523487394998/Variable-color-output-from-the-foam-spray-end-effector-is-seen.png</a>, accessed on 25 April 2024.</p>
Full article ">Figure 4
<p>Examples of image augmentation for the WGISD dataset.</p>
Full article ">Figure 5
<p>Average precision/loss curve for <b>608@VOC+WA-WGISD</b> training.</p>
Full article ">Figure 6
<p>Grape cluster replicas used during simulations.</p>
Full article ">Figure 7
<p>Bounding Box Inference from <b>608@VOC+WA-WGISD</b> in Open-Space Controlled Environment (BBI-OSCE).</p>
Full article ">Figure 8
<p>Bounding Box Inference of <b>608@VOC+WA-WGISD</b> in Vineyard (BBI-V)-Part 1/2.</p>
Full article ">Figure 9
<p>Bounding Box Inference of <b>608@VOC+WA-WGISD</b> in Vineyard (BBI-V)-Part 2/2.</p>
Full article ">Figure 10
<p>Cobotic preliminary experiment: spraying test in indoor controlled environment.</p>
Full article ">Figure 11
<p>Cobotic preliminary experiment: spraying test in vineyard.</p>
Full article ">Figure 12
<p>Examples of buds/<span class="html-italic">veraison</span> bunch detection of <b>608@VOC+WA-WGISD</b> for GBDDv2 test images (right in each image), compared with corresponding ground truth (left in each image).</p>
Full article ">
23 pages, 2892 KiB  
Article
Estimation of Intercepted Solar Radiation and Stem Water Potential in a Table Grape Vineyard Covered by Plastic Film Using Sentinel-2 Data: A Comparison of OLS-, MLR-, and ML-Based Methods
by Alessandro Farbo, Nicola Gerardo Trombetta, Laura de Palma and Enrico Borgogno-Mondino
Plants 2024, 13(9), 1203; https://doi.org/10.3390/plants13091203 - 25 Apr 2024
Viewed by 1336
Abstract
In the framework of precision viticulture, satellite data have been demonstrated to significantly support many tasks. Specifically, they enable the rapid, large-scale estimation of some viticultural parameters like vine stem water potential (Ψstem) and intercepted solar radiation (ISR) that traditionally require time-consuming ground [...] Read more.
In the framework of precision viticulture, satellite data have been demonstrated to significantly support many tasks. Specifically, they enable the rapid, large-scale estimation of some viticultural parameters like vine stem water potential (Ψstem) and intercepted solar radiation (ISR) that traditionally require time-consuming ground surveys. The practice of covering table grape vineyards with plastic films introduces an additional challenge for estimation, potentially affecting vine spectral responses and, consequently, the accuracy of estimations from satellites. This study aimed to address these challenges with a special focus on the exploitation of Sentinel-2 Level 2A and meteorological data to monitor a plastic-covered vineyard in Southern Italy. Estimates of Ψstem and ISR were obtained using different algorithms, namely, Ordinary Least Square (OLS), Multivariate Linear Regression (MLR), and machine learning (ML) techniques, which rely on Random Forest Regression, Support Vector Regression, and Partial Least Squares. The results proved that, despite the potential spectral interference from the plastic coverings, ISR and Ψstem can be locally estimated with a satisfying accuracy. In particular, (i) the OLS regression-based approach showed a good performance in providing accurate ISR estimates using the near-infrared spectral bands (RMSE < 8%), and (ii) the MLR and ML algorithms could estimate both the ISR and vine water status with a higher accuracy (RMSE < 7 for ISR and RMSE < 0.14 MPa for Ψstem). These results encourage the adoption of medium–high resolution multispectral satellite imagery for deriving satisfying estimates of key crop parameters even in anomalous situations like the ones where plastic films cover the monitored vineyard, thus marking a significant advancement in precision viticulture. Full article
(This article belongs to the Section Crop Physiology and Crop Production)
Show Figures

Figure 1

Figure 1
<p>Test vineyard located in Southern Italy, Apulia Region (area with green lines). Black squares represent the spatial distribution of surveyed replicates (a square of 4 contiguous vines/replicate). The reference system is WGS 84/UTM 32N, EPSG:32633.</p>
Full article ">Figure 2
<p>Test vineyard covered with plastic sheets.</p>
Full article ">Figure 3
<p>Temporal trends of satellite-derived spectral features collected during the year (DOY): (<b>a</b>) reflectance values of visible bands (B2, B3, B4); (<b>b</b>) reflectance values of red-edge and NIR bands (B5, B6, B7, B8, B8A); (<b>c</b>) reflectance values of shortwave infrared bands (B11, B12); (<b>d</b>) vegetative indices (NDVI, GDVI, EVI, NDRE). Vertical dashed lines correspond to the time series change points identified by the Pettit test (dashed line colors match the temporal profile colors).</p>
Full article ">Figure 4
<p>ISR (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) and Ψstem (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>) prediction maps of the studied vineyard derived from the RFR models at (<b>a</b>,<b>b</b>) DOY 150 (‘berries are groat-sized’); (<b>c</b>,<b>d</b>) DOY 181 (‘all berries are touching’); (<b>e</b>,<b>f</b>) DOY 192 (‘beginning of ripening’); (<b>g</b>,<b>h</b>) DOY 217 (‘after harvest’). Red dots correspond to the ground surveyed areas. The reference system is WGS 84/UTM 32N, EPSG:32633.</p>
Full article ">
18 pages, 47387 KiB  
Article
Time Series Analysis of Multisensor Data for Precision Viticulture—Assessing Microscale Variations in Plant Development with Respect to Irrigation and Topography
by Melanie Brandmeier, Daniel Heßdörfer, Philipp Siebenlist, Adrian Meyer-Spelbrink and Anja Kraus
Remote Sens. 2024, 16(8), 1419; https://doi.org/10.3390/rs16081419 - 17 Apr 2024
Viewed by 1077
Abstract
In the context of climate change, vineyard monitoring to better understand spatiotemporal patterns of grapevine development is of utter importance for precision viticulture. We present a time series analysis of hyperspectral in situ and multispectral UAV data for different irrigation systems in Lower [...] Read more.
In the context of climate change, vineyard monitoring to better understand spatiotemporal patterns of grapevine development is of utter importance for precision viticulture. We present a time series analysis of hyperspectral in situ and multispectral UAV data for different irrigation systems in Lower Franconia and correlate results with sensor data for soil moisture, temperature, and precipitation. Analysis of Variance (ANOVA) and a Tukey’s HSD test were performed to see whether Vegetation Indices (VIs) are significantly different with respect to irrigation systems as well as topographic position in the vineyard. Correlation between in situ measurements and UAV data for selected VIs is also investigated for upscaling analysis. We find significant differences with respect to irrigation, as well as for topographic position for most of the VIs investigated, highlighting the importance of adapted water management. Correlation between in situ and UAV data is significant only for some indices (NDVI and CIRedEdge, r2 of 0.33 and 0.49, respectively), while shallow soil moisture patterns correlate well with in situ-derived VIs such as the CIRedEdge and RG index (r2 of 0.34 and 0.46). Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overall workflow of this study. For more information, refer to the text.</p>
Full article ">Figure 2
<p>Study area near Himmelstadt, Bavaria. Different irrigation systems are shown in red, green, and blue. Locations for hyperspectral measurements on plants are shown depending on their location in the vineyard (yellow outline). Locations of tubes for soil moisture readings are shown in blue. RGB of the flight on 13 July as basemap. The climate station is visible north of the vineyard.</p>
Full article ">Figure 3
<p>Example of one NDVI calculation (17 July 2023) showing the plant locations and zonal statistics areas used for correlation with spectroradiometer measurements. As vines are aligned along wires, it is important to automatically derive the maximum VI value locations (for some indices minimum) and avoid sampling erroneous pixels (such as bare soil or stem areas), as can be seen in the zoomed-in location.</p>
Full article ">Figure 4
<p>Time series of (<b>A</b>) NDVI, (<b>B</b>) RG Index, (<b>C</b>) NGRDI, (<b>D</b>) MSI. NDVI and NGRDI show lower values for nonirrigated plants from June through to September (NGRDI) and are more pronounced in August for the NDVI. The RG index follows the same pattern with higher values for nonirrigated plants. The moisture stress index, MSI, on the other hand, shows slightly higher values for irrigated plants, especially after strong rainfall in August. Additional plots for VIs not shown here can be found in <a href="#app1-remotesensing-16-01419" class="html-app">Appendix A</a>.</p>
Full article ">Figure 5
<p>Time series of (<b>A</b>) NDVI, (<b>B</b>) CIRedEdge Index, (<b>C</b>) RG Index, (<b>D</b>) GLI index. For NDVI, there are no significant differences with respect to topography, while the CIRedEdge index is less favorable for the lower part of the vineyard. The RG index is lower for the lower part of the vineyard, while the GLI is higher for most of the vegetation period. Additional plots for VIs not shown here can be found in <a href="#app1-remotesensing-16-01419" class="html-app">Appendix A</a>.</p>
Full article ">Figure 6
<p>Matrix showing results of the Tukey HSD test after performing ANOVA analysis for all indices with respect to topographic and irrigation classes. Test results are shown for in situ data as well as UAV-derived indices. Significant test results are highlighted in green.</p>
Full article ">Figure 7
<p>Time series of (<b>A</b>) NDVI, (<b>B</b>) RG index, (<b>C</b>) NGRD index, (<b>D</b>) GNDVI index. Significant differences for nonirrigated plots are clearly pronounced for all indices shown. The steep drop (NDVI, NGRDI, GNDVI) at the end of September is not clearly visible in the spectrometer data, as the last measurement took place on the 14th of September. Additional plots for VIs not shown here can be found in <a href="#app1-remotesensing-16-01419" class="html-app">Appendix A</a>.</p>
Full article ">Figure 8
<p>Correlation analysis of in situ data with UAV-derived VIs. (<b>A</b>) CIRedEdge, (<b>B</b>) NDVI.</p>
Full article ">Figure 9
<p>(<b>A</b>) Mean soil moisture at different depth levels over the vegetation period. Stdv. is plotted in shaded colors. Rainfall events are shown in blue. We observe high fluctuations in shallow soil levels, while deeper levels mainly react to prolonged rainfall in July to August. (<b>B</b>) GNDVI mean values for plants located at soil moisture measurement tubes. Stdv. is shown in shaded colors. We observe a steep rise in GNDVI values after prolonged rainfall at the end of June and again in August.</p>
Full article ">Figure A1
<p>Additional plots for <a href="#remotesensing-16-01419-f004" class="html-fig">Figure 4</a> in the paper: Time series of spectrometer data with respect to irrigation systems for GLI and CIRedEdge.</p>
Full article ">Figure A2
<p>Additional plots for <a href="#remotesensing-16-01419-f005" class="html-fig">Figure 5</a> in the paper: Time series of spectrometer data with respect to topography for MSI and NGRDI.</p>
Full article ">Figure A3
<p>Additional plot for <a href="#remotesensing-16-01419-f007" class="html-fig">Figure 7</a> in the paper: Time series of UAV data with respect to irrigation for the CIRedEdge (MSI not calculated).</p>
Full article ">
18 pages, 13828 KiB  
Article
Automated Derivation of Vine Objects and Ecosystem Structures Using UAS-Based Data Acquisition, 3D Point Cloud Analysis, and OBIA
by Stefan Ruess, Gernot Paulus and Stefan Lang
Appl. Sci. 2024, 14(8), 3264; https://doi.org/10.3390/app14083264 - 12 Apr 2024
Viewed by 867
Abstract
This study delves into the analysis of a vineyard in Carinthia, Austria, focusing on the automated derivation of ecosystem structures of individual vine parameters, including vine heights, leaf area index (LAI), leaf surface area (LSA), and the geographic positioning of single plants. For [...] Read more.
This study delves into the analysis of a vineyard in Carinthia, Austria, focusing on the automated derivation of ecosystem structures of individual vine parameters, including vine heights, leaf area index (LAI), leaf surface area (LSA), and the geographic positioning of single plants. For the derivation of these parameters, intricate segmentation processes and nuanced UAS-based data acquisition techniques are necessary. The detection of single vines was based on 3D point cloud data, generated at a phenological stage in which the plants were in the absence of foliage. The mean distance from derived vine locations to reference measurements taken with a GNSS device was 10.7 cm, with a root mean square error (RMSE) of 1.07. Vine height derivation from a normalized digital surface model (nDSM) using photogrammetric data showcased a strong correlation (R2 = 0.83) with real-world measurements. Vines underwent automated classification through an object-based image analysis (OBIA) framework. This process enabled the computation of ecosystem structures at the individual plant level post-segmentation. Consequently, it delivered comprehensive canopy characteristics rapidly, surpassing the speed of manual measurements. With the use of uncrewed aerial systems (UAS) equipped with optical sensors, dense 3D point clouds were computed for the derivation of canopy-related ecosystem structures of vines. While LAI and LSA computations await validation, they underscore the technical feasibility of obtaining precise geometric and morphological datasets from UAS-collected data paired with 3D point cloud analysis and object-based image analysis. Full article
Show Figures

Figure 1

Figure 1
<p>Study area in Carinthia showing seven blocks of wine grapes across the 0.7 ha parcel.</p>
Full article ">Figure 2
<p>Three-dimensional point cloud together with the camera positions.</p>
Full article ">Figure 3
<p>Methodical workflow diagram showing the analysis steps that lead from the processed data sets ready for analysis to the derived vine-related parameters: LSA, LAI, vine height as indicated by the green boxes.</p>
Full article ">Figure 4
<p>Portion of the 3D point cloud recorded in April 2023 showing the vine in the phenological stage EL 1-3. Bird exclusion netting is placed around the trunks of the vines all year around.</p>
Full article ">Figure 5
<p>(<b>Left</b>): The maximal values of the mesh to cloud distance calculation represent the tips of the individual vine trunks. (<b>Right</b>): The nadir view of the calculated maximal values of the mesh to cloud distance calculation.</p>
Full article ">Figure 6
<p>The image objects created with the multi-threshold segmentation approach.</p>
Full article ">Figure 7
<p>The automatically derived vine heights in meters for possible nutrient supply calculation or other resource optimization problems, like pruning management.</p>
Full article ">Figure 8
<p>(<b>Left</b>): Three-dimensional point cloud collected 1 h before harvest. (<b>Right</b>): The filtered grape cluster points from the same point-cloud at the same location.</p>
Full article ">Figure 9
<p>The created mesh computed with the leaf points only, used for the calculation of canopy-related metrics like leave surface area and LAI.</p>
Full article ">Figure 10
<p>Parameters derived at the vine level.</p>
Full article ">Figure 11
<p>Comparison of the derived vine points and the reference vine points.</p>
Full article ">Figure 12
<p>Locations of the 23 vine height reference measurements.</p>
Full article ">Figure 13
<p>Linear regression of the vine height validation.</p>
Full article ">Figure 14
<p>The matrix displays the correlation coefficients between the variables. Strong correlations are shown in dark red shades.</p>
Full article ">
18 pages, 296 KiB  
Article
Case Studies on Sustainability-Oriented Innovations and Smart Farming Technologies in the Wine Industry: A Comparative Analysis of Pilots in Cyprus and Italy
by Aikaterini Kasimati, George Papadopoulos, Valentina Manstretta, Marianthi Giannakopoulou, George Adamides, Damianos Neocleous, Vassilis Vassiliou, Savvas Savvides and Andreas Stylianou
Agronomy 2024, 14(4), 736; https://doi.org/10.3390/agronomy14040736 - 2 Apr 2024
Viewed by 1814
Abstract
Addressing the urgent sustainability challenges in the wine industry, this study explores the efficacy of sustainability-oriented innovations (SOIs) and smart farming technologies (SFTs) across wine value chains in Cyprus and Italy. Utilising a mixed-methods approach that includes quantitative analysis through Key Performance Indicators [...] Read more.
Addressing the urgent sustainability challenges in the wine industry, this study explores the efficacy of sustainability-oriented innovations (SOIs) and smart farming technologies (SFTs) across wine value chains in Cyprus and Italy. Utilising a mixed-methods approach that includes quantitative analysis through Key Performance Indicators (KPIs) and qualitative assessments to understand stakeholders’ perspectives, this research delves into the environmental, economic, and social impacts of these technologies. In Cyprus, the integration of digital labelling and smart farming solutions led to a substantial reduction in pesticide usage by up to 75% and enhanced the perceived quality of wine by an average of 8%. A pilot study in Italy witnessed a 33.4% decrease in greenhouse gas emissions, with the additional benefit of a 5.3% improvement in intrinsic product quality. The pilot introduced a carbon credit system, potentially generating an average annual revenue of EUR 4140 per farm. These findings highlight the transformative potential of SOIs and SFTs in promoting sustainable practices within the wine industry, demonstrating significant advancements in reducing environmental impact, improving product quality, and enhancing economic viability. This study underscores the critical role of innovative technologies in achieving sustainability goals and provides a compelling case for their wider adoption within the agricultural sector. Full article
(This article belongs to the Section Precision and Digital Agriculture)
22 pages, 12087 KiB  
Article
A Cloud-Based Deep Learning Framework for Downy Mildew Detection in Viticulture Using Real-Time Image Acquisition from Embedded Devices and Drones
by Sotirios Kontogiannis, Myrto Konstantinidou, Vasileios Tsioukas and Christos Pikridas
Information 2024, 15(4), 178; https://doi.org/10.3390/info15040178 - 24 Mar 2024
Cited by 1 | Viewed by 1419
Abstract
In viticulture, downy mildew is one of the most common diseases that, if not adequately treated, can diminish production yield. However, the uncontrolled use of pesticides to alleviate its occurrence can pose significant risks for farmers, consumers, and the environment. This paper presents [...] Read more.
In viticulture, downy mildew is one of the most common diseases that, if not adequately treated, can diminish production yield. However, the uncontrolled use of pesticides to alleviate its occurrence can pose significant risks for farmers, consumers, and the environment. This paper presents a new framework for the early detection and estimation of the mildew’s appearance in viticulture fields. The framework utilizes a protocol for the real-time acquisition of drones’ high-resolution RGB images and a cloud-docker-based video or image inference process using object detection CNN models. The authors implemented their framework proposition using open-source tools and experimented with their proposed implementation on the debina grape variety in Zitsa, Greece, during downy mildew outbursts. The authors present evaluation results of deep learning Faster R-CNN object detection models trained on their downy mildew annotated dataset, using the different object classifiers of VGG16, ViTDet, MobileNetV3, EfficientNet, SqueezeNet, and ResNet. The authors compare Faster R-CNN and YOLO object detectors in terms of accuracy and speed. From their experimentation, the embedded device model ViTDet showed the worst accuracy results compared to the fast inferences of YOLOv8, while MobileNetV3 significantly outperformed YOLOv8 in terms of both accuracy and speed. Regarding cloud inferences, large ResNet models performed well in terms of accuracy, while YOLOv5 faster inferences presented significant object classification losses. Full article
(This article belongs to the Special Issue Deep Learning for Image, Video and Signal Processing)
Show Figures

Figure 1

Figure 1
<p>Proposed object detection framework inputs, outputs, and steps.</p>
Full article ">Figure 2
<p>Proposed high-level system architecture that supports the object detection framework process.</p>
Full article ">Figure 3
<p>IoT plant-level monitoring camera nodes, ThingsBoard dashboard, and normal and mildew-infected leaf inferences using different Faster R-CNN models. (<b>a</b>) IoT plant-level autonomous camera nodes and their corresponding parts (1)–(4). (<b>b</b>) IoT plant-level monitoring ThingsBoard dashboard. (<b>c</b>) IoT plant-level inferences using the MobileNetV3 model. (<b>d</b>) IoT plant-level inferences using the ResNet-50 model.</p>
Full article ">Figure 4
<p>IoT plant-level monitoring camera devices, ThingsBoard dashboard drone GPS locations acquired from image metadata, and normal and mildew-infected leaf inferences using object detection models on image streams. (<b>a</b>) IoT plant-level monitored viticulture field using drones. (<b>b</b>) Drone GPS locations from captured image EXIF metadata [<a href="#B53-information-15-00178" class="html-bibr">53</a>], as illustrated in ThingsBoard. (<b>c</b>) IoT plant-level video stream inferences using YOLOv5-small model.</p>
Full article ">Figure 4 Cont.
<p>IoT plant-level monitoring camera devices, ThingsBoard dashboard drone GPS locations acquired from image metadata, and normal and mildew-infected leaf inferences using object detection models on image streams. (<b>a</b>) IoT plant-level monitored viticulture field using drones. (<b>b</b>) Drone GPS locations from captured image EXIF metadata [<a href="#B53-information-15-00178" class="html-bibr">53</a>], as illustrated in ThingsBoard. (<b>c</b>) IoT plant-level video stream inferences using YOLOv5-small model.</p>
Full article ">Figure 5
<p>IoT plant-level video stream inferences using (<b>1</b>) ResNet-50 and (<b>2</b>) ResNet-152 models.</p>
Full article ">Figure 6
<p>Annotation process using LabelImg tool on (<b>a</b>) IoT camera nodes and (<b>b</b>) drone acquired images. Two distinct annotation classes were used for normal and downy mildew-infected leaves.</p>
Full article ">Figure 7
<p>Precision—recall mAP scores for threshold values 0.5–0.95 and a step of 0.05 (<math display="inline"><semantics> <mrow> <mi>m</mi> <mi>A</mi> <msub> <mi>P</mi> <mrow> <mn>0.5</mn> <mo>:</mo> <mn>0.95</mn> </mrow> </msub> </mrow> </semantics></math>) for cloud object detection models.</p>
Full article ">Figure 8
<p>Classification loss scores over epochs for cloud object detection models.</p>
Full article ">Figure 9
<p>Precision—recall mAP scores for threshold values 0.5–0.95 and step of 0.05 (<math display="inline"><semantics> <mrow> <mi>m</mi> <mi>A</mi> <msub> <mi>P</mi> <mrow> <mn>0.5</mn> <mo>:</mo> <mn>0.95</mn> </mrow> </msub> </mrow> </semantics></math>) for embedded and mobile device object detection models.</p>
Full article ">Figure 10
<p>Classification loss scores over epochs for embedded and mobile device object detection models.</p>
Full article ">
14 pages, 1243 KiB  
Article
Determination of Pesticide Residues in Vine Leaves Using the QuEChERS Method and Liquid Chromatography-Tandem Mass Spectrometry
by Mehmet Keklik, Ozgur Golge, Miguel Ángel González-Curbelo and Bulent Kabak
Foods 2024, 13(6), 909; https://doi.org/10.3390/foods13060909 - 17 Mar 2024
Cited by 1 | Viewed by 1413
Abstract
Commercial viticulture necessitates regular pesticide applications to manage diseases and pests, raising significant concerns regarding pesticide residues among stakeholders. Due to health risks associated with these residues in Turkish vine leaves, the European Commission has increased the frequency of official control from 20% [...] Read more.
Commercial viticulture necessitates regular pesticide applications to manage diseases and pests, raising significant concerns regarding pesticide residues among stakeholders. Due to health risks associated with these residues in Turkish vine leaves, the European Commission has increased the frequency of official control from 20% to 50%. Thus, the aim of this study was to determine multi-class pesticide residues in brined vine leaves from Turkey. A total of 766 samples of vine leaves were collected between May 2022 and June 2023. More than 500 residues were analyzed using the quick, easy, cheap, effective, rugged, and safe (QuEChERS) method, followed by liquid chromatography-tandem mass spectrometry. In-house validation data demonstrated that the analytical method exhibits fit-for-purpose performance in terms of linearity, accuracy, precision, and measurement uncertainty. Out of 766 samples analyzed, 180 samples (23.5%) contained one (131, 17.1%) or multiple (49, 6.4%) pesticides. Both the frequencies of occurrence and the rate of maximum residue level (MRL) exceedance increased in 2023 compared to 2022, with the MRL exceedance rate rising from 9.5% to 25.2%. Forty-three different residues were found in quantifiable concentrations and eight of them were non-approved. Among the residues, the non-systemic pyrethroid insecticides, lambda-cyhalothrin (8.0%) and cypermethrin (7.2%), were the two most frequently detected, with concentrations ranging from 0.010 to 0.248 mg kg−1 and from 0.011 to 0.533 mg kg−1, respectively. Turkey is a major exporter of vine leaves and these results provide crucial information regarding pesticide occurrence and quality assessment of vine leaves. The significant increase in both pesticide occurrence and MRL exceedance rates between 2022 and 2023 underscores the urgency for regulatory bodies to reassess current pesticide usage and monitoring practices. The findings emphasize the importance of implementing more stringent rules and improving enforcement methods in order to reduce the spread of unapproved pesticides and ensure adherence to global food safety standards. Full article
(This article belongs to the Section Food Analytical Methods)
Show Figures

Figure 1

Figure 1
<p>The procedure of the QuEChERS sample preparation method.</p>
Full article ">Figure 2
<p>Percentage of vine leaf samples from 2022 without any residues or with residues.</p>
Full article ">Figure 3
<p>Percentage of vine leaf samples from 2023 without any residues or with residues.</p>
Full article ">
Back to TopTop