[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

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (821)

Search Parameters:
Keywords = canopy reflectance

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 11081 KiB  
Article
Mapping Forest Carbon Stock Distribution in a Subtropical Region with the Integration of Airborne Lidar and Sentinel-2 Data
by Xiaoyu Sun, Guiying Li, Qinquan Wu, Jingyi Ruan, Dengqiu Li and Dengsheng Lu
Remote Sens. 2024, 16(20), 3847; https://doi.org/10.3390/rs16203847 - 16 Oct 2024
Viewed by 289
Abstract
Forest carbon stock is an important indicator reflecting a forest ecosystem’s structures and functions. Its spatial distribution is valuable for managing natural resources, protecting ecosystems and biodiversity, and further promoting sustainability, but accurately mapping the forest carbon stock distribution in a large area [...] Read more.
Forest carbon stock is an important indicator reflecting a forest ecosystem’s structures and functions. Its spatial distribution is valuable for managing natural resources, protecting ecosystems and biodiversity, and further promoting sustainability, but accurately mapping the forest carbon stock distribution in a large area is a challenging task. This study selected Changting County, Fujian Province, as a case study to explore a method to map the forest carbon stock distribution using the integration of airborne Lidar, Sentinel-2, and ancillary data in 2022. The Bayesian hierarchical modeling approach was used to estimate the local forest carbon stock based on airborne Lidar data and field measurements, and then the random forest approach was used to develop a regional forest carbon stock estimation model based on the Sentinel-2 and ancillary data. The results indicated that the Lidar-based carbon stock distribution effectively provided sample plots with good spatial representativeness for modeling regional carbon stock with a coefficient of determination (R2) of 0.7 and root mean square error (RMSE) of 12.94 t/ha. The average carbon stocks were 48.55 t/ha, 55.51 t/ha, and 57.04 t/ha for Masson pine, Chinese fir, and broadleaf forests, respectively. The carbon stock in non-conservation regions was 15.2–16.1 t/ha higher than that in conservation regions. This study provides a promising method through the use of airborne Lidar data as a linkage between sample plots and Sentinel-2 data to map the regional carbon stock distribution in those subtropical regions where serious soil erosion has led to a relatively sparse forest canopy density. The results are valuable for local government to make scientific decisions for promoting ecosystem restoration due to water and soil erosion. Full article
(This article belongs to the Special Issue Forest Biomass/Carbon Monitoring towards Carbon Neutrality)
Show Figures

Figure 1

Figure 1
<p>Study area—Changting County, in the western part of Fujian Province, China ((<b>a</b>) Location of Changting County; (<b>b</b>) Distribution of major land cover types, overlaid by the conservation area, Lidar data, and field samples).</p>
Full article ">Figure 2
<p>Framework for developing a forest carbon stock estimation model in Changting County based on the combination of field measurements, airborne Lidar, Sentinel-2, and ancillary data (CHM: canopy height model; DEM: digital elevation model).</p>
Full article ">Figure 3
<p>Scatter plots between predicted carbon stock from Lidar data and reference data from field measurements in typical areas using Bayesian hierarchical model ((<b>a</b>)—forest type and soil subgroup, (<b>b</b>)—forest type and conservation or not, (<b>c</b>)—forest type and slope group).</p>
Full article ">Figure 4
<p>Spatial distribution of predicted forest carbon stock in different typical areas which were labelled as (<b>a</b>–<b>g</b>).</p>
Full article ">Figure 5
<p>Scatter plot between predicted carbon stock and reference data at regional scale (The red line represents 1:1 line).</p>
Full article ">Figure 6
<p>Spatial distribution of predicted forest carbon stock in Changting County in 2022.</p>
Full article ">
25 pages, 3566 KiB  
Article
Excellent Canopy Structure in Soybeans Can Improve Their Photosynthetic Performance and Increase Yield
by Shuyuan He, Xiuni Li, Menggen Chen, Xiangyao Xu, Wenjing Zhang, Huiling Chi, Panxia Shao, Fenda Tang, Tao Gong, Ming Guo, Mei Xu, Wenyu Yang and Weiguo Liu
Agriculture 2024, 14(10), 1783; https://doi.org/10.3390/agriculture14101783 - 11 Oct 2024
Viewed by 419
Abstract
In the maize-soybean intercropping system, varying degrees of maize leaf shading are an important factor that reduces the uniformity of light penetration within the soybean canopy, altering the soybean canopy structure. Quantitative analysis of the relationship between the soybean canopy structure and canopy [...] Read more.
In the maize-soybean intercropping system, varying degrees of maize leaf shading are an important factor that reduces the uniformity of light penetration within the soybean canopy, altering the soybean canopy structure. Quantitative analysis of the relationship between the soybean canopy structure and canopy photosynthesis helps with breeding shade-tolerant soybean varieties for intercropping systems. This study examined the canopy structure and photosynthesis of intercropped soybeans during the shading stress period (28 days before the corn harvest), the high light adaptation period (15 days after the corn harvest), and the recovery period (35 and 55 days after the corn harvest), using a field high-throughput phenotyping platform and a plant gas exchange testing system (CAPTS). Additionally, indoor shading experiments were conducted for validation. The results indicate that shade-tolerant soybean varieties (STV varieties) have significantly higher yields than shade-sensitive soybean varieties (SSV varieties). This is attributable to the STV varieties having a larger top area, lateral width, and lateral external rectangular area. Compared to the SSV varieties, the four top areas of the STV varieties are, on average, 52.09%, 72.05%, and 61.37% higher during the shading stress, high light adaptation, and recovery periods, respectively. Furthermore, the average maximum growth rates (GRs) for the side mean width (SMW) and side rectangle area (SRA) of the STV varieties are 62.92% and 22.13% in the field, and 83.36% and 55.53% in the indoor environment, respectively. This results in a lower canopy overlap in STV varieties, leading to a more uniform light distribution within the canopy, which is reflected in higher photosynthetic rates (Pn), apparent quantum efficiency, and whole-leaf photosynthetic potential (WLPP) for the STV varieties, thereby enhancing their adaptability to shading stress. Above-ground dry matter accumulation was higher in STV varieties, with more assimilates stored in the source and sink, promoting assimilate accumulation in the grains. These results provide new insights into how the superior canopy structure and photosynthesis of shade-tolerant soybean varieties contribute to increased yield. Full article
(This article belongs to the Section Crop Production)
Show Figures

Figure 1

Figure 1
<p>Soybean test layout. Figure (<b>a</b>) shows the experimental layout of the soybeans in field conditions. Figure (<b>b</b>) represents the original image generated by an RGB camera; Figure (<b>c</b>) shows the layout of the field environmental test and the real state of canopy photosynthesis.</p>
Full article ">Figure 2
<p>Changes in dry matter content in leaves, main stems, and branches with emergence time. The red solid line in the figure indicates the day of the corn harvest. STV-1, STV-2, SSV-1, and SSV-2 represent ND12, NJQP, C103, and BYH soybean varieties, respectively. “M” represents soybean monoculture, and “I” represents soybean and corn intercropping.</p>
Full article ">Figure 3
<p>Figures (<b>a</b>–<b>f</b>) represent, in sequence, soybean yield per plant, number of grains per plant, 100-grain weight, number of branches, number of pods on branches, and number of pods on the main stem. The significance analysis in this study is conducted at the <span class="html-italic">p</span> = 0.05 level. In the figures, lowercase and uppercase letters are used to distinguish between sole cropping and intercropping levels. Lowercase letters indicate sole cropping, while uppercase letters indicate intercropping, and the same applies to the following figures.</p>
Full article ">Figure 4
<p>Diurnal variation of canopy photosynthetic rate.</p>
Full article ">Figure 5
<p>Light response diagram. Figures (<b>a</b>–<b>d</b>) represent the light response curves of soybeans at 28 days before corn harvest, and at 15, 35, and 55 days after corn harvest, respectively. The solid line in the figure represents the information concerning the canopy light response curve, and the dashed line represents the apparent quantum efficiency of the STV-1 and SSV-1 varieties.</p>
Full article ">Figure 6
<p>Whole-leaf photosynthetic potential map. The red solid line in the figure indicates the day of the corn harvest.</p>
Full article ">Figure 7
<p>Four canopy structure parameters related to the top area. The significance analysis in this study is conducted at the <span class="html-italic">p</span> = 0.05 level. In the figures, lowercase and uppercase letters are used to distinguish between sole cropping and intercropping levels. Lowercase letters indicate sole cropping, while uppercase letters indicate intercropping.</p>
Full article ">Figure 7 Cont.
<p>Four canopy structure parameters related to the top area. The significance analysis in this study is conducted at the <span class="html-italic">p</span> = 0.05 level. In the figures, lowercase and uppercase letters are used to distinguish between sole cropping and intercropping levels. Lowercase letters indicate sole cropping, while uppercase letters indicate intercropping.</p>
Full article ">Figure 8
<p>Two canopy structure parameters related to side width. The significance analysis in this study is conducted at the <span class="html-italic">p</span> = 0.05 level. In the figures, lowercase and uppercase letters are used to distinguish between sole cropping and intercropping levels. Lowercase letters indicate sole cropping, while uppercase letters indicate intercropping.</p>
Full article ">Figure 9
<p>1_5, 2_5, 3_5, 4_5, 5_5 side width diagram. In the figure, M represents the net planting mode, and I represents the intercropping planting mode; 1–4 represent STV-1, STV-2, SSV-1, and SSV-2, respectively.</p>
Full article ">Figure 10
<p>Illustrates the mechanism of STV variety yield prominence. The blue rectangles in the canopy structure represent the top and side bounding rectangle areas, and the yellow circles represent the top bounding circle area. The gray boxes highlight the advantages of STV varieties in terms of canopy structure, photosynthetic activity, and assimilate accumulation, which lead to their outstanding yield performance.</p>
Full article ">
19 pages, 11653 KiB  
Article
Influence of Vegetation Phenology on the Temporal Effect of Crop Fractional Vegetation Cover Derived from Moderate-Resolution Imaging Spectroradiometer Nadir Bidirectional Reflectance Distribution Function–Adjusted Reflectance
by Yinghao Lin, Tingshun Fan, Dong Wang, Kun Cai, Yang Liu, Yuye Wang, Tao Yu and Nianxu Xu
Agriculture 2024, 14(10), 1759; https://doi.org/10.3390/agriculture14101759 - 5 Oct 2024
Viewed by 406
Abstract
Moderate-Resolution Imaging Spectroradiometer (MODIS) Nadir Bidirectional Reflectance Distribution Function (BRDF)-Adjusted Reflectance (NBAR) products are being increasingly used for the quantitative remote sensing of vegetation. However, the assumption underlying the MODIS NBAR product’s inversion model—that surface anisotropy remains unchanged over the 16-day retrieval period—may [...] Read more.
Moderate-Resolution Imaging Spectroradiometer (MODIS) Nadir Bidirectional Reflectance Distribution Function (BRDF)-Adjusted Reflectance (NBAR) products are being increasingly used for the quantitative remote sensing of vegetation. However, the assumption underlying the MODIS NBAR product’s inversion model—that surface anisotropy remains unchanged over the 16-day retrieval period—may be unreliable, especially since the canopy structure of vegetation undergoes stark changes at the start of season (SOS) and the end of season (EOS). Therefore, to investigate the MODIS NBAR product’s temporal effect on the quantitative remote sensing of crops at different stages of the growing seasons, this study selected typical phenological parameters, namely SOS, EOS, and the intervening stable growth of season (SGOS). The PROBA-V bioGEOphysical product Version 3 (GEOV3) Fractional Vegetation Cover (FVC) served as verification data, and the Pearson correlation coefficient (PCC) was used to compare and analyze the retrieval accuracy of FVC derived from the MODIS NBAR product and MODIS Surface Reflectance product. The Anisotropic Flat Index (AFX) was further employed to explore the influence of vegetation type and mixed pixel distribution characteristics on the BRDF shape under different stages of the growing seasons and different FVC; that was then combined with an NDVI spatial distribution map to assess the feasibility of using the reflectance of other characteristic directions besides NBAR for FVC correction. The results revealed the following: (1) Generally, at the SOSs and EOSs, the differences in PCCs before vs. after the NBAR correction mainly ranged from 0 to 0.1. This implies that the accuracy of FVC derived from MODIS NBAR is lower than that derived from MODIS Surface Reflectance. Conversely, during the SGOSs, the differences in PCCs before vs. after the NBAR correction ranged between –0.2 and 0, suggesting the accuracy of FVC derived from MODIS NBAR surpasses that derived from MODIS Surface Reflectance. (2) As vegetation phenology shifts, the ensuing differences in NDVI patterning and AFX can offer auxiliary information for enhanced vegetation classification and interpretation of mixed pixel distribution characteristics, which, when combined with NDVI at characteristic directional reflectance, could enable the accurate retrieval of FVC. Our results provide data support for the BRDF correction timescale effect of various stages of the growing seasons, highlighting the potential importance of considering how they differentially influence the temporal effect of NBAR corrections prior to monitoring vegetation when using the MODIS NBAR product. Full article
Show Figures

Figure 1

Figure 1
<p>Spatial extent of the Wancheng District study area (in Henan Province, China). (<b>a</b>) Map of land cover types showing the location of sampling points across the study area. This map came from MCD12Q1 (v061). (<b>b</b>–<b>d</b>) True-color images of the three mixed pixels, obtained from Sentinel-2. The distribution characteristics are as follows: crops above with buildings below (<b>b</b>); crops below with buildings above (<b>c</b>); and buildings in the upper-left corner, crops in the remainder (<b>d</b>).</p>
Full article ">Figure 2
<p>Monthly average temperature and monthly total precipitation in the study area, from 2017 to 2021.</p>
Full article ">Figure 3
<p>Data processing flow chart. The green rectangles from top to the bottom represent three steps: crop phenological parameters extraction with TIMESAT; Fractional Vegetation Cover (FVC) derived from MOD09GA and MCD43A4; and accuracy evaluation, respectively. Blue solid rectangles refer to a used product or derived results, while blue dashed rectangles refer to the software or model used in this study. NDVI<sub>MOD09GA</sub>: NDVI derived from MOD09GA, NDVI<sub>MCD43A4</sub>: NDVI derived from MCD43A4, FVC<sub>MOD09GA</sub>: FVC derived from MOD09GA, FVC<sub>MCD43A4</sub>: FVC derived from MCD43A4. PCC<sub>MOD09GA</sub>: Pearson correlation coefficient (PCC) calculated for FVC<sub>MOD09GA</sub> and GEOV3 FVC, PCC<sub>MCD43A4</sub>: PCC calculated for FVC<sub>MCD43A4</sub> and GEOV3 FVC.</p>
Full article ">Figure 4
<p>NDVI and EVI time series fitted curves and phenological parameters of crops. SOS: start of season; EOS: end of season; SGOS: stable growth of season.</p>
Full article ">Figure 5
<p>Spatial distribution of Fractional Vegetation Cover (FVC) derived from MOD09GA and MCD43A4, and the difference images of FVC. FVC<sub>MOD09GA</sub>: FVC derived from MOD09GA, FVC<sub>MCD43A4</sub>: FVC derived from MCD43A4. (<b>a</b>–<b>c</b>) FVC derived from MOD09GA, MCD43A4, and the difference between FVC<sub>MOD09GA</sub> and FVC<sub>MCD43A4</sub> on 15 November 2020, respectively; (<b>d</b>–<b>f</b>) FVC derived from MOD09GA, MCD43A4, and the difference between FVC<sub>MOD09GA</sub> and FVC<sub>MCD43A4</sub> on 10 February 2021, respectively; (<b>g</b>–<b>i</b>) FVC derived from MOD09GA, MCD43A4, and the difference between FVC<sub>MOD09GA</sub> and FVC<sub>MCD43A4</sub> on 30 September 2021, respectively.</p>
Full article ">Figure 6
<p>Pearson correlation coefficients (PCCs) of Fractional Vegetation Cover (FVC) derived before and after the NBAR correction with GEOV3 FVC at different stages of the growing seasons. FVC<sub>MOD09GA</sub>: FVC derived from MOD09GA. FVC<sub>MCD43A4</sub>: FVC derived from MCD43A4. PCC<sub>MOD09GA</sub>: PCC calculated for FVC<sub>MOD09GA</sub> and GEOV3 FVC, PCC<sub>MCD43A4</sub>: PCC calculated for FVC<sub>MCD43A4</sub> and GEOV3 FVC. (<b>a</b>) PCC<sub>MOD09GA</sub> and PCC<sub>MCD43A4</sub> in 2018–2021; (<b>b</b>) Scatterplot of numerical differences between PCC<sub>MOD09GA</sub> and PCC<sub>MCD43A4</sub>. SOS: start of season; EOS: end of season; SGOS: stable growth of season.</p>
Full article ">Figure 7
<p>NDVI spatial distribution maps of crop pixel, savanna pixel, and grassland pixel in different stages of the growing seasons. (<b>a</b>–<b>d</b>) Crop. (<b>e</b>–<b>h</b>) Savanna. (<b>i</b>–<b>l</b>) Grassland. SZA: Solar Zenith Angle, FVC: Fractional Vegetation Cover, AFX_RED: Anisotropic Flat Index (AFX) in the red band, AFX_NIR: AFX in the near-infrared band.</p>
Full article ">Figure 8
<p>NDVI spatial distribution maps of mixed pixels in different stages of the growing seasons. (<b>a</b>–<b>d</b>) Crops above and buildings below. (<b>e</b>–<b>h</b>) Crops below and buildings above. (<b>i</b>–<b>l</b>) Buildings in the upper-left corner and crops in the remainder. SZA: Solar Zenith Angle, FVC: Fractional Vegetation Cover, AFX_RED: Anisotropic Flat Index (AFX) in the red band, AFX_NIR: AFX in the near-infrared band.</p>
Full article ">
29 pages, 12094 KiB  
Article
Bitemporal Radiative Transfer Modeling Using Bitemporal 3D-Explicit Forest Reconstruction from Terrestrial Laser Scanning
by Chang Liu, Kim Calders, Niall Origo, Louise Terryn, Jennifer Adams, Jean-Philippe Gastellu-Etchegorry, Yingjie Wang, Félicien Meunier, John Armston, Mathias Disney, William Woodgate, Joanne Nightingale and Hans Verbeeck
Remote Sens. 2024, 16(19), 3639; https://doi.org/10.3390/rs16193639 - 29 Sep 2024
Viewed by 752
Abstract
Radiative transfer models (RTMs) are often used to retrieve biophysical parameters from earth observation data. RTMs with multi-temporal and realistic forest representations enable radiative transfer (RT) modeling for real-world dynamic processes. To achieve more realistic RT modeling for dynamic forest processes, this study [...] Read more.
Radiative transfer models (RTMs) are often used to retrieve biophysical parameters from earth observation data. RTMs with multi-temporal and realistic forest representations enable radiative transfer (RT) modeling for real-world dynamic processes. To achieve more realistic RT modeling for dynamic forest processes, this study presents the 3D-explicit reconstruction of a typical temperate deciduous forest in 2015 and 2022. We demonstrate for the first time the potential use of bitemporal 3D-explicit RT modeling from terrestrial laser scanning on the forward modeling and quantitative interpretation of: (1) remote sensing (RS) observations of leaf area index (LAI), fraction of absorbed photosynthetically active radiation (FAPAR), and canopy light extinction, and (2) the impact of canopy gap dynamics on light availability of explicit locations. Results showed that, compared to the 2015 scene, the hemispherical-directional reflectance factor (HDRF) of the 2022 forest scene relatively decreased by 3.8% and the leaf FAPAR relatively increased by 5.4%. At explicit locations where canopy gaps significantly changed between the 2015 scene and the 2022 scene, only under diffuse light did the branch damage and closing gap significantly impact ground light availability. This study provides the first bitemporal RT comparison based on the 3D RT modeling, which uses one of the most realistic bitemporal forest scenes as the structural input. This bitemporal 3D-explicit forest RT modeling allows spatially explicit modeling over time under fully controlled experimental conditions in one of the most realistic virtual environments, thus delivering a powerful tool for studying canopy light regimes as impacted by dynamics in forest structure and developing RS inversion schemes on forest structural changes. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Geographic location and map of Wytham Woods with plot indicated by ‘X’ [<a href="#B60-remotesensing-16-03639" class="html-bibr">60</a>].</p>
Full article ">Figure 2
<p>Spectral properties of different tree species in the plot [<a href="#B2-remotesensing-16-03639" class="html-bibr">2</a>,<a href="#B51-remotesensing-16-03639" class="html-bibr">51</a>]. (<b>a</b>) Reflectance and transmittance of leaves; (<b>b</b>) reflectance of bark.</p>
Full article ">Figure 3
<p>Locations of canopy gap dynamics and photosynthetically active radiation (PAR) sensors simulated, shown in the TLS point cloud (top view): (<b>a</b>) 2015; (<b>b</b>) 2022.</p>
Full article ">Figure 4
<p>Vertical profiles of different types of canopy gap dynamics observed by terrestrial laser scanning, and the position of simulated PAR sensors.</p>
Full article ">Figure 5
<p>Flowchart of research methodology. QSMs of woody structure were reconstructed using leaf-off TLS data.</p>
Full article ">Figure 6
<p>Segmented TLS leaf-off point cloud of 1-ha Wytham Woods forest stand (top view): (<b>a</b>) 2015; (<b>b</b>) and 2022. Each color represents an individual tree.</p>
Full article ">Figure 7
<p>The dynamic change of wood structure of a Common ash (<span class="html-italic">Fraxinus excelsior</span>) tree from 2015 to 2022. (<b>a</b>) 2015 leaf-off point cloud; (<b>b</b>) 2022 leaf-off point cloud.</p>
Full article ">Figure 8
<p>3D-explicit reconstruction of a Sycamore (<span class="html-italic">Acer pseudoplatanus</span>) tree. (<b>a</b>) TLS point cloud colored by height (leaf-off); (<b>b</b>) QSM overlaid with TLS leaf-off point cloud; (<b>c</b>) QSM, the modeled branch length was 3863.3 m; (<b>d</b>) Fully reconstructed tree: QSM + leaves, the leaf area assigned to this tree was 888.2 m<sup>2</sup>.</p>
Full article ">Figure 9
<p>The 3D-explicit models of the complete 1-ha Wytham Woods forest stand in (<b>a</b>) 2015 and (<b>b</b>) 2022. The different leaf colors represent the different tree species present in Wytham Woods. The stems and branches of all trees are shown in brown.</p>
Full article ">Figure 10
<p>The vertical profiles of simulated (<b>a</b>) light extinction, (<b>b</b>) light absorption, and (<b>c</b>) leaf area per meter of height in 2015 and 2022 forest scenes. The results of light extinction and absorption were based on the PAR band. The illumination zenith angle (IZA) was 38.4° and the illumination azimuth angle (IAA) was 125.2°.</p>
Full article ">Figure 11
<p>The vertical profiles of simulated (<b>a</b>) light extinction and (<b>b</b>) light absorption in the blue, green, red, and NIR bands for the 2015 and 2022 forest scenes. Illumination zenith angle (IZA) 38.4°, illumination azimuth angle (IAA) 125.2°.</p>
Full article ">Figure 12
<p>Simulated top of canopy images of Wytham Woods forest scenes in 2015 and 2022. The images were simulated under nadir viewing directions and Sentinel-2 RGB bands. IZA 38.4°, IAA 125.2°. (<b>a</b>,<b>b</b>) Ultra-high resolution images in 2015 and 2022 (spatial resolution: 1 cm); (<b>d</b>,<b>e</b>) 25 cm resolution images in 2015 and 2022; (<b>g</b>,<b>h</b>) 10 m resolution images in 2015 and 2022; (<b>c</b>,<b>f</b>,<b>i</b>) Spatial pattern of HDRF variation from 2015 to 2022 (red band).</p>
Full article ">Figure 13
<p>Light extinction profiles of downward PAR at location 1: (<b>a</b>) diffuse light; (<b>b</b>) midday direct light (IZA 28.4°, IAA 180°); (<b>c</b>) morning direct light (IZA 81.3°, IAA 27.3°). The X axis is the local light availability represented as the percentage of incident solar irradiance. The Y axis is the height from the simulated sensors to the ground. (<b>d</b>) The canopy gap dynamic at this location.</p>
Full article ">Figure 14
<p>Light extinction profiles of downward PAR at location 2: (<b>a</b>) diffuse light; (<b>b</b>) midday direct light (IZA 28.4°, IAA 180°); (<b>c</b>) morning direct light (IZA 81.3°, IAA 27.3°). The X axis is the local light availability represented as the percentage of incident solar irradiance. The Y axis is the height from the simulated sensors to the ground. (<b>d</b>) The canopy gap dynamic at this location.</p>
Full article ">Figure 15
<p>Light extinction profiles of downward PAR at location 3: (<b>a</b>) diffuse light; (<b>b</b>) midday direct light (IZA 28.4°, IAA 180°); (<b>c</b>) morning direct light (IZA 81.3°, IAA 27.3°). The X axis is the local light availability represented as the percentage of incident solar irradiance. The Y axis is the height from the simulated sensors to the ground. (<b>d</b>) The canopy gap dynamic at this location.</p>
Full article ">Figure 16
<p>Light extinction profiles of downward PAR at location 4: (<b>a</b>) diffuse light; (<b>b</b>) midday direct light (IZA 28.4°, IAA 180°); (<b>c</b>) morning direct light (IZA 81.3°, IAA 27.3°). The X axis is the local light availability represented as the percentage of incident solar irradiance. The Y axis is the height from the simulated sensors to the ground. (<b>d</b>) The canopy gap dynamic at this location.</p>
Full article ">
29 pages, 1326 KiB  
Review
Site-Specific Nitrogen Fertilizer Management Using Canopy Reflectance Sensors, Chlorophyll Meters and Leaf Color Charts: A Review
by Ali M. Ali, Haytham M. Salem and Bijay-Singh
Nitrogen 2024, 5(4), 828-856; https://doi.org/10.3390/nitrogen5040054 - 27 Sep 2024
Viewed by 693
Abstract
The efficient management of nitrogen (N) on a site-specific basis is critical for the improvement of crop yield and the reduction of environmental impacts. This review examines the application of three primary technologies—canopy reflectance sensors, chlorophyll meters, and leaf color charts—in the context [...] Read more.
The efficient management of nitrogen (N) on a site-specific basis is critical for the improvement of crop yield and the reduction of environmental impacts. This review examines the application of three primary technologies—canopy reflectance sensors, chlorophyll meters, and leaf color charts—in the context of site-specific N fertilizer management. It delves into the development and effectiveness of these tools in assessing and managing crop N status. Reflectance sensors, which measure the reflection of light at specific wavelengths, provide valuable data on plant N stress and variability. The advent of innovative sensor technology, exemplified by the GreenSeeker, Crop Circle sensors, and Yara N-Sensor, has facilitated real-time monitoring and precise adjustments in fertilizer N application. Chlorophyll meters, including the SPAD meter and the atLeaf meter, quantify chlorophyll content and thereby estimate leaf N levels. This indirect yet effective method of managing N fertilization is based on the principle that the concentration of chlorophyll in leaves is proportional to the N content. These meters have become an indispensable component of precision agriculture due to their accuracy and ease of use. Leaf color charts, while less sophisticated, offer a cost-effective and straightforward approach to visual N assessment, particularly in developing regions. This review synthesizes research on the implementation of these technologies, emphasizing their benefits, constraints, and practical implications. Additionally, it explores integration strategies for combining these tools to enhance N use efficiency and sustainability in agriculture. The review culminates with recommendations for future research and development to further refine the precision and efficacy of N management practices. Full article
Show Figures

Figure 1

Figure 1
<p>Global consumption of agricultural fertilizer from 1965 to 2021, disaggregated by nutrient. Data source: <a href="https://www.statista.com" target="_blank">https://www.statista.com</a> (accessed 26 July 2024).</p>
Full article ">Figure 2
<p>Illustrative example of normalized difference vegetation index (NDVI) measurement and analysis using a hand-held optical sensor.</p>
Full article ">Figure 3
<p>Example of chlorophyll index measurement and analysis using a SPAD meter.</p>
Full article ">Figure 4
<p>Illustration of an LCC showing a series of color panels representing varying levels of leaf greenness. The leaf in the example corresponds to panel number 3 on the LCC.</p>
Full article ">
20 pages, 6068 KiB  
Article
Design and Testing of a Fruit Tree Variable Spray System Based on ExG-AABB
by Daozong Sun, Zhiwei Quan, Peiran Wu, Weikang Liu, Xiuyun Xue, Shuran Song, Jiaxing Xie and Sheng Jiang
Agronomy 2024, 14(10), 2199; https://doi.org/10.3390/agronomy14102199 - 25 Sep 2024
Viewed by 389
Abstract
This paper addresses the issue of pesticide waste and low utilization rates resulting from traditional plant protection via spraying operations, which apply equal dosages to different targets or to different parts of the same target. To tackle this problem, we designed a variable [...] Read more.
This paper addresses the issue of pesticide waste and low utilization rates resulting from traditional plant protection via spraying operations, which apply equal dosages to different targets or to different parts of the same target. To tackle this problem, we designed a variable fruit tree spraying system based on the ExG-AABB (excess green and axis-aligned bounding box) algorithm. We used a Kinect depth camera to capture information about the fruit tree canopy and constructed a spray flow model using pulse width modulation and variable spray control technology. Variable multi-nozzle spraying was guided by combining this canopy data. We evaluated the accuracy of each model in calculating canopy volume by comparing the coefficient of determination (R2) and root mean square error (RMSE) of the ExG-AABB with the slice convex hull method, voxel method, three-dimensional alpha-shape method, and QuickHull method. The ExG-AABB algorithm had the highest R2 value (0.9334) and the lowest RMSE value (0.0353 m3) among the five models, indicating that it most accurately reflects the true volume of the fruit tree canopy. This validates the effectiveness of the ExG-AABB algorithm in calculating canopy volume. We established a correlation model between canopy volume and spray volume, designed a canopy-adaptive layering method based on point cloud processing, and achieved precise calculation of nozzle flow. Comparative field experiments were conducted to analyze the spray coverage rate and observed flow, thereby evaluating the spraying effect of this variable spraying system. The experimental results showed that compared to conventional continuous spraying, this variable spraying system not only achieves more uniform spray coverage but also significantly reduces pesticide usage by 48.1%. Furthermore, through system optimization, the average coverage rate of the middle layer of the canopy decreased by 17.53%, effectively reducing the phenomenon of overlapping spraying from multiple nozzles and improving spraying efficiency. Full article
(This article belongs to the Section Precision and Digital Agriculture)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of the system and (<b>b</b>) physical diagram of the system. Note: (1) sprayer; (2) depth camera; (3) tracked vehicle; (4) flow meter; (5) water tank.</p>
Full article ">Figure 2
<p>Manual measurement schematic of canopy volume.</p>
Full article ">Figure 3
<p>Simulated trees.</p>
Full article ">Figure 4
<p>Canopy profile.</p>
Full article ">Figure 5
<p>(<b>a</b>) Three-dimensional point cloud scene of fruit trees before filtering and (<b>b</b>) after filtering.</p>
Full article ">Figure 6
<p>Sample distribution diagram.</p>
Full article ">Figure 7
<p>Canopy and background classification results.</p>
Full article ">Figure 8
<p>Schematic diagram of the canopy point cloud after using the AABB algorithm.</p>
Full article ">Figure 9
<p>Correlation between the canopy volume logarithm and leaf wall area of the fruit trees.</p>
Full article ">Figure 10
<p>Schematic diagram of the results of three volume algorithms: (<b>a</b>) convex hull calculated in slices; (<b>b</b>) voxel-based method; (<b>c</b>) QuickHull (<b>d</b>) three-dimensional alpha shape.</p>
Full article ">Figure 11
<p>Spray flow rate versus a PWM duty cycle at 0.4 MPa.</p>
Full article ">Figure 12
<p>Water-sensitive paper droplet coverage.</p>
Full article ">Figure 13
<p>(<b>a</b>) Distribution structure and (<b>b</b>) structural geometric plane diagram. Note: a, c: distance in the vertical direction of spraying, b: distance in the vertical direction of overlapping spraying area.</p>
Full article ">Figure 14
<p>(<b>a</b>) Distribution of sampling points in the fruit tree canopy and (<b>b</b>) distribution of water-sensitive paper.</p>
Full article ">Figure 15
<p>Scatter plots using five volume algorithms with manually measured volumes: (<b>a</b>) ExG-AABB method; (<b>b</b>) slicing convex hull method; (<b>c</b>) the volumetric method; (<b>d</b>) three-dimensional alpha shape; (<b>e</b>) QuickHull.</p>
Full article ">Figure 16
<p>Variable spray coverage compared with continuous spray coverage.</p>
Full article ">
19 pages, 12139 KiB  
Article
Inversion Modeling of Chlorophyll Fluorescence Parameters in Cotton Canopy via Moisture Data and Spectral Analysis
by Fuqing Li, Caiyun Yin, Zhen Li, Jiaqiang Wang, Long Jiang, Buping Hou and Jing Shi
Agronomy 2024, 14(10), 2190; https://doi.org/10.3390/agronomy14102190 - 24 Sep 2024
Viewed by 384
Abstract
The study of chlorophyll fluorescence parameters is very important for understanding plant photosynthesis. Monitoring cotton chlorophyll fluorescence parameters via spectral technology can aid in understanding the photosynthesis, growth, and stress of cotton fields in real time and provide support for cotton growth regulation [...] Read more.
The study of chlorophyll fluorescence parameters is very important for understanding plant photosynthesis. Monitoring cotton chlorophyll fluorescence parameters via spectral technology can aid in understanding the photosynthesis, growth, and stress of cotton fields in real time and provide support for cotton growth regulation and planting management. In this study, cotton plot experiments with different water treatments were set up to obtain the spectral reflectance of the cotton canopy, the maximum photochemical quantum yield (Fv/Fm), and the photochemical quenching coefficient (qP) of leaves at different growth stages. Support vector machine regression (SVR), random forest regression (RFR), and artificial neural network regression (ANNR) were used to establish a fluorescence parameter inversion model of the cotton canopy leaves. The results show that the original spectrum was transformed by multivariate scattering correction (MSC), the standard normal variable (SNV), and continuous wavelet transform (CWT), and the model constructed with Fv/Fm passed accuracy verification. The SNV-SVR model at the budding stage, the MSC-SVR model at the early flowering stage, the SNV-SVR model at the full flowering stage, the MSC-SVR model at the flowering stage, and the CWT-SVR model at the full boll stage had the highest estimation accuracy. The accuracies of the three spectral preprocessing and qP models were verified, and the MSC-SVR model at the budding stage, SNV-SVR model at the early flowering stage, MSC-SVR model at the full flowering stage, SNV-SVR model at the flowering stage, and CWT-SVR model at the full boll stage presented the highest estimation accuracies. Full article
Show Figures

Figure 1

Figure 1
<p>Changes in the chlorophyll fluorescence parameters of cotton leaves at different growth stages. (<b>a</b>) Fv/Fm changes at different growth stages, (<b>b</b>) qP changes at different growth stages. (In the figure, a, b, c, ab and bc are the letter marking methods for significance analysis. If they do not contain the same letters, there is a significant difference).</p>
Full article ">Figure 2
<p>Correlation heatmap of chlorophyll fluorescence parameters with moisture data. (<b>a</b>) Heat map of correlation between BS growth period and water data, (<b>b</b>) Heat map of correlation between EF growth period and water data, (<b>c</b>) Heat map of correlation between FB growth period and water data, (<b>d</b>) Heat map of correlation between FBS growth period and water data, (<b>e</b>) Heat map of correlation between PBP growth period and water data.</p>
Full article ">Figure 3
<p>Heatmap of the correlation between characteristic bands and chlorophyll fluorescence characteristic parameters. (<b>a</b>) Heat map of correlation between Fv/Fm and characteristic bands during BS growth period, (<b>b</b>) Heat map of correlation between qP and characteristic bands during BS growth period, (<b>c</b>) Heat map of correlation between Fv/Fm and characteristic bands during EF growth period, (<b>d</b>) Heat map of correlation between qP and characteristic bands during EF growth period, (<b>e</b>) Heat map of correlation between Fv/Fm and characteristic bands during FB growth period, (<b>f</b>) Heat map of correlation between qP and characteristic bands during FB growth period, (<b>g</b>) Heat map of correlation between qP and characteristic bands during FBS growth period, (<b>h</b>) Heat map of correlation between qP and characteristic bands during FBS growth period, (<b>i</b>) Heat map of correlation between Fv/Fm and characteristic bands during PBP growth period, (<b>j</b>) Heat map of correlation between qP and characteristic bands during PBP growth period.</p>
Full article ">Figure 4
<p>Scatter plots of the optimal simulated and measured values of Fv/Fm (<b>a</b>) and qP (<b>b</b>) during the BS period.</p>
Full article ">Figure 5
<p>Scatter plots of the optimal simulated and measured values of Fv/Fm (<b>a</b>) and qP (<b>b</b>) in the EF period.</p>
Full article ">Figure 6
<p>Scatter plots of the optimal simulated and measured values of Fv/Fm (<b>a</b>) and qP (<b>b</b>) during the FB period linearly fitted to the measured values.</p>
Full article ">Figure 7
<p>Scatter plots of the optimal simulated and measured values of Fv/Fm (<b>a</b>) and qP (<b>b</b>) during the FBS period.</p>
Full article ">Figure 8
<p>Scatter plots of the optimal simulated and measured values of Fv/Fm (<b>a</b>) and qP (<b>b</b>) during the PBP period.</p>
Full article ">
19 pages, 7626 KiB  
Article
Measuring Biophysical Parameters of Wheat Canopy with MHz- and GHz-Frequency Range Impulses Employing Contactless GPR
by Konstantin Muzalevskiy, Sergey Fomin, Andrey Karavayskiy, Julia Leskova, Alexey Lipshin and Vasily Romanov
Remote Sens. 2024, 16(19), 3547; https://doi.org/10.3390/rs16193547 - 24 Sep 2024
Viewed by 412
Abstract
In this paper, the advantages of the joint use of MHz- and GHz-frequency band impulses when employing contactless ground penetration radar (GPR) for the remote sensing of biomass, the height of the wheat canopy, and underlying soil moisture were experimentally investigated. A MHz-frequency [...] Read more.
In this paper, the advantages of the joint use of MHz- and GHz-frequency band impulses when employing contactless ground penetration radar (GPR) for the remote sensing of biomass, the height of the wheat canopy, and underlying soil moisture were experimentally investigated. A MHz-frequency band nanosecond impulse with a duration of 1.2 ns (average frequency of 750 MHz and spectrum bandwidth of 580 MHz, at a level of –6 dB) was emitted and received by a GPR OKO-3 equipped with an AB-900 M3 antenna unit. A GHz-frequency band sub-nanosecond impulse with a duration of 0.5 ns (average frequency of 3.2 GHz and spectral bandwidth of 1.36 GHz, at a level of −6 dB) was generated using a horn antenna and a Keysight FieldFox N9917B 18 GHz vector network analyzer. It has been shown that changes in the relative amplitudes and time delays of nanosecond impulses, reflected from a soil surface covered with wheat at a height from 0 to 87 cm and fresh above-ground biomass (AGB) from 0 to 1.5 kg/m2, do not exceed 6% and 0.09 ns, respectively. GPR nanosecond impulses reflected/scattered by the wheat canopy have not been detected. In this research, sub-nanosecond impulses reflected/scattered by the wheat canopy have been confidently identified and make it possible to measure the wheat height (fresh AGB up to 2.3 kg/m2 and height up to 104 cm) with a determination coefficient (R2) of ~0.99 and a bias of ~−7 cm, as well as fresh AGB where R2 = 0.97, with a bias = −0.09 kg/m2, and a root-mean-square error of 0.1 kg/m2. The joint use of impulses in two different MHz- and GHz-frequency bands will, in the future, make it possible to create UAV-based reflectometers for simultaneously mapping the soil moisture, height, and biomass of vegetation for precision farming systems. Full article
Show Figures

Figure 1

Figure 1
<p>Geographic location of the test field (red rectangle).</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>d</b>) The process used in the experiment on the remote sensing of the wheat canopy in the MHz-frequency range with an OKO-3 GPR (23 August 2023); (<b>d</b>) measurement over a metal screen. Free space calibration is not shown.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>d</b>) The process used in the experiment on the remote sensing of the wheat canopy in the GHz-frequency range with a horn antenna (3 September 2023); (<b>d</b>) measurement over a metal screen. Free space calibration is not shown.</p>
Full article ">Figure 4
<p>Lambda dependency of the <span class="html-italic">b</span>-factors: 1—corn [<a href="#B64-remotesensing-16-03547" class="html-bibr">64</a>,<a href="#B65-remotesensing-16-03547" class="html-bibr">65</a>], 2—soybean [<a href="#B64-remotesensing-16-03547" class="html-bibr">64</a>,<a href="#B65-remotesensing-16-03547" class="html-bibr">65</a>,<a href="#B66-remotesensing-16-03547" class="html-bibr">66</a>,<a href="#B67-remotesensing-16-03547" class="html-bibr">67</a>], 3—wheat [<a href="#B64-remotesensing-16-03547" class="html-bibr">64</a>,<a href="#B65-remotesensing-16-03547" class="html-bibr">65</a>], 4—alfalfa [<a href="#B64-remotesensing-16-03547" class="html-bibr">64</a>,<a href="#B65-remotesensing-16-03547" class="html-bibr">65</a>], 5—wheat grains [<a href="#B60-remotesensing-16-03547" class="html-bibr">60</a>], 6—cereals, and sorghum [<a href="#B66-remotesensing-16-03547" class="html-bibr">66</a>] are <span class="html-italic">b<sub>rad</sub></span>-factors, estimated based on radiometric measurements, and the 7–<span class="html-italic">b<sub>refr</sub></span>-factor, calculated based on the refractive mixing model (4).</p>
Full article ">Figure 5
<p>Flowchart of algorithms used for soil geophysical and canopy biometric parameters, based on the measured OKO-3 GPR data in the time domain (TD) and VNA data in the frequency domain (FD).</p>
Full article ">Figure 6
<p>(<b>a</b>) Time shapes of the GPR impulse MHz-frequency range, reflected from the vegetation–soil cover <span class="html-italic">s<sub>sv</sub></span>(<span class="html-italic">t</span>) (color lines) and metal reflector <span class="html-italic">s<sub>ref</sub></span>(<span class="html-italic">t</span>) (gray solid line), with the normalized envelope of impulse, reflected from the metal screen |<span class="html-italic">R<sub>ref</sub></span>(<span class="html-italic">t</span>)|=<math display="inline"><semantics> <mrow> <mtext> </mtext> <mfenced open="|" close="|" separators="|"> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>s</mi> </mrow> <mo>˙</mo> </mover> </mrow> <mrow> <mi>r</mi> <mi>e</mi> <mi>f</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>t</mi> </mrow> </mfenced> </mrow> </mfenced> <mo>/</mo> <msub> <mrow> <mfenced open="|" close="|" separators="|"> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>s</mi> </mrow> <mo>˙</mo> </mover> </mrow> <mrow> <mi>r</mi> <mi>e</mi> <mi>f</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>t</mi> </mrow> </mfenced> </mrow> </mfenced> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> (gray dashed line); (<b>b</b>) the corresponding normalized module of the impulse spectrum |<span class="html-italic">S<sub>sv,ref</sub></span>(<span class="html-italic">f</span>)|; (<b>c</b>) the module of the reflection coefficient |<span class="html-italic">R<sub>sv</sub></span>(<span class="html-italic">f</span><sub>0</sub>)| and the delay time Δ<span class="html-italic">t</span> (calculated from the maximum envelopes) of GPR impulses (see <a href="#remotesensing-16-03547-f006" class="html-fig">Figure 6</a>a), depending on the measured fresh AGB value calculated by the thermostat-weight method.</p>
Full article ">Figure 7
<p>(<b>a</b>) The measured (solid lines) and retrieved (dash lines) modules <math display="inline"><semantics> <mrow> <mo>|</mo> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>s</mi> <mi>v</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>f</mi> <mo>,</mo> <mi>B</mi> </mrow> </mfenced> <mo>|</mo> <mtext> </mtext> </mrow> </semantics></math> and arguments <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> <mi>r</mi> <mi>g</mi> <mtext> </mtext> <mi>R</mi> </mrow> <mrow> <mi>s</mi> <mi>v</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>f</mi> <mo>,</mo> <mi>B</mi> </mrow> </mfenced> </mrow> </semantics></math> of the spectrum of reflection coefficients; (<b>b</b>) the correlation between measured and retrieved fresh AGB values. The optimally found parameters while solving the inverse problem (6) were <span class="html-italic">W<sub>retr</sub> </span>= 23.4 ± 2.6% and σ<span class="html-italic"><sub>r,retr</sub></span>= 1.4 ± 0.3 cm. The color scheme of solid and dashed color lines is the same (various colors of lines correspond to the different values for fresh AGB in [kg/m<sup>2</sup>]).</p>
Full article ">Figure 8
<p>(<b>a</b>) Time shapes of the GPR impulses GHz-frequency range, reflected from the vegetation–soil cover <span class="html-italic">s<sub>sv</sub></span>(<span class="html-italic">t</span>) (color lines) and metal reflector <span class="html-italic">s<sub>ref</sub></span>(<span class="html-italic">t</span>) (gray solid line), normalized envelope of impulse, reflected from the metal screen |<span class="html-italic">R<sub>ref</sub></span>(<span class="html-italic">t</span>)| (gray dashed line); (<b>b</b>) the corresponding normalized module of the impulse spectrum |<span class="html-italic">S<sub>sv,ref</sub></span>(<span class="html-italic">f</span>)|. Various colors of lines correspond to the different values of fresh AGB in [kg/m<sup>2</sup>]. The color scheme is the same as in the pictures.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) Normalized time shapes of impulse <span class="html-italic">s<sub>sv</sub></span>(<span class="html-italic">t</span>), reflected from the vegetation–soil cover and (<b>c</b>,<b>d</b>) their envelopes <math display="inline"><semantics> <mrow> <mfenced open="|" close="|" separators="|"> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>s</mi> <mi>v</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>t</mi> </mrow> </mfenced> </mrow> </mfenced> <mo>=</mo> <mfenced open="|" close="|" separators="|"> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>s</mi> </mrow> <mo>˙</mo> </mover> </mrow> <mrow> <mi>s</mi> <mi>v</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>t</mi> </mrow> </mfenced> </mrow> </mfenced> <mo>/</mo> <msub> <mrow> <mfenced open="|" close="|" separators="|"> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>s</mi> </mrow> <mo>˙</mo> </mover> </mrow> <mrow> <mi>r</mi> <mi>e</mi> <mi>f</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>t</mi> </mrow> </mfenced> </mrow> </mfenced> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> as wheat is cut down from the top to the soil surface. In the legend, the numbers indicate fresh AGB in [kg/m<sup>2</sup>], measured in situ by the thermostat–weight method. The key for the black circles will be made clear further on.</p>
Full article ">Figure 10
<p>Retrieved local values of the volumetric contents of vegetation elements Δ<span class="html-italic">v<sub>v,retr</sub></span>(<span class="html-italic">z</span>) and the refractive index <span class="html-italic">n<sub>can,retr</sub></span>(<span class="html-italic">z</span>) in the canopy. The numbers in the legend indicate the values of fresh AGB in [kg/m<sup>2</sup>], measured by the thermostat-weight method.</p>
Full article ">Figure 11
<p>Dependence of the retrieved data on the measured canopy heights.</p>
Full article ">Figure 12
<p>(<b>a</b>) The retrieved <span class="html-italic">B<sub>retr</sub></span> (color lines) and measured <span class="html-italic">B<sub>meas</sub></span> (line with black circles) for fresh AGB, depending on the retrieved <span class="html-italic">h<sub>can,retr</sub></span> and measured <span class="html-italic">h<sub>can,meas</sub></span> canopy heights, respectively; (<b>b</b>) correlation between the retrieved and measured fresh AGB values.</p>
Full article ">Figure 13
<p>Relief of the norm of the difference between measured and retrieved fresh AGB values in total for all wheat-height cuts (see <a href="#remotesensing-16-03547-f012" class="html-fig">Figure 12</a>a), depending on the half-bandwidth Δ<span class="html-italic">f</span> and the central frequency <span class="html-italic">f</span><sub>0</sub> of the Gaussian window function (see <a href="#sec2dot2-remotesensing-16-03547" class="html-sec">Section 2.2</a> and the text for Formulas (1)–(2)).</p>
Full article ">
17 pages, 4648 KiB  
Article
Multitemporal Hyperspectral Characterization of Wheat Infested by Wheat Stem Sawfly, Cephus cinctus Norton
by Lochlin S. Ermatinger, Scott L. Powell, Robert K. D. Peterson and David K. Weaver
Remote Sens. 2024, 16(18), 3505; https://doi.org/10.3390/rs16183505 - 21 Sep 2024
Viewed by 419
Abstract
Wheat (Triticum aestivum L.) production in the Northern Great Plains of North America has been challenged by wheat stem sawfly (WSS), Cephus cinctus Norton, for a century. Damaging WSS populations have increased, highlighting the need for reliable surveys. Remote sensing (RS) can [...] Read more.
Wheat (Triticum aestivum L.) production in the Northern Great Plains of North America has been challenged by wheat stem sawfly (WSS), Cephus cinctus Norton, for a century. Damaging WSS populations have increased, highlighting the need for reliable surveys. Remote sensing (RS) can be used to correlate reflectance measurements with nuanced phenomena like cryptic insect infestations within plants, yet little has been done with WSS. To evaluate interactions between WSS-infested wheat and spectral reflectance, we grew wheat plants in a controlled environment, experimentally infested them with WSS and recorded weekly hyperspectral measurements (350–2500 nm) of the canopies from prior to the introduction of WSS to full senescence. To assess the relationships between WSS infestation and wheat reflectance, we employed sparse multiway partial least squares regression (N-PLS), which models multidimensional covariance structures inherent in multitemporal hyperspectral datasets. Multitemporal hyperspectral measurements of wheat canopies modeled with sparse N-PLS accurately estimated the proportion of WSS-infested stems (R2 = 0.683, RMSE = 13.5%). The shortwave-infrared (1289–1380 nm) and near-infrared (942–979 nm) spectral regions were the most important in estimating infestation, likely due to internal feeding that decreases plant-water content. Measurements from all time points were important, suggesting aerial RS of WSS in the field should incorporate the visible through shortwave spectra collected from the beginning of WSS emergence at least weekly until the crop reaches senescence. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>(top left): Interior view of lights; (bottom left) view of the plant from where the sensor is situated; (right) outside view of the entire sampling environment. Sampling environment telescopes vertically to adjust sampling area of the ASD FieldSpec Pro to the canopy of each pot.</p>
Full article ">Figure 2
<p>Mean reflectance at 14 days after infestation: (<b>a</b>) red vertical lines indicate cut-off points of noisy spectra; (<b>b</b>) black lines are the VIF values calculated from the IBRA to the reflectance spectrum of all control samples from 14 days after infestation.</p>
Full article ">Figure 3
<p>Flowchart illustrating the experimental, data preprocessing, and analysis steps followed to link multitemporal hyperspectral reflectance to the proportion of adequately infested wheat stem sawfly (WSS) stems within a pot.</p>
Full article ">Figure 4
<p>Opposing viewpoints of the multitemporal reflectance of all sampled wheat canopies averaged at each point in time. Early time points are characterized by strong absorption features while these regions increase in reflectance as plant phenology and wheat stem sawfly infestation progresses.</p>
Full article ">Figure 5
<p>RMSE of calibration and validation dataset based on number of latent variables. The vertical black dashed line indicates the sparse N-PLS model with 13 latent variables produced the optimal fit based on RMSE with the calibration and validation datasets.</p>
Full article ">Figure 6
<p>Predicted vs. fitted values amongst the calibration and validation datasets. Spread of points along the 1:1 line suggests no issues with heteroscedasticity. The green dashed line represents the line of best fit, where the gray area is the 95% confidence interval.</p>
Full article ">Figure 7
<p>Distribution of spectral–temporal features retained by the final sparse N-PLS model. Imposing sparsity via LASSO significantly reduced the number of spectral–temporal features from 12,159 to 400. The signs of the beta coefficients indicate the features’ relationship with estimating the proportion of adequate WSS infestation.</p>
Full article ">Figure 8
<p>Mean spectral reflectance and beta coefficients given days after infestation. The average spectral reflectance increased over time, especially in the VIS and SWIR absorption features, as phenology advanced and plants senesced. The signs and magnitudes of the beta coefficients represent their correlations and weights in estimating the proportion of adequately WSS-infested stems within a pot.</p>
Full article ">Figure A1
<p>Picture of wheat stem sawfly (WSS) larva feeding inside a wheat stem. The frass is excrement produced during and after feeding by WSS larvae. Photo by Jackson R. Strand.</p>
Full article ">Figure A2
<p>Examples of possible wheat stem sawfly (WSS) infestation categories recorded at the conclusion of the experiment. A first instar larva that lived and died within a single internode is represented on the far left (<b>a</b>), denoted as <span class="html-italic">dead neonate</span>. A larva that burrowed through <span class="html-italic">two or more nodes</span> but ultimately died before cutting the stem is depicted in (<b>b</b>). The most significant possible form of WSS damage, <span class="html-italic">WSS cut</span> (<b>c</b>), occurs when a WSS larva burrows through all, or almost all nodes, and eventually severs the stem near the soil surface, where it creates a hibernaculum to prepare for diapause. An <span class="html-italic">uninfested</span> stem that did not experience WSS infestation will be devoid of frass and all nodes remain intact (<b>d</b>).</p>
Full article ">Figure A3
<p>Proportion of adequately wheat stem sawfly infested stems within a pot by planting replicate. Replicates were planted weekly with A being the first and D being the last. Infestation varied by planting replicate and did not display a normal distribution even when considering all replicates together.</p>
Full article ">
30 pages, 3394 KiB  
Article
Integrating Hyperspectral Reflectance-Based Phenotyping and SSR Marker-Based Genotyping for Assessing the Salt Tolerance of Wheat Genotypes under Real Field Conditions
by Salah El-Hendawy, Muhammad Bilawal Junaid, Nasser Al-Suhaibani, Ibrahim Al-Ashkar and Abdullah Al-Doss
Plants 2024, 13(18), 2610; https://doi.org/10.3390/plants13182610 - 19 Sep 2024
Viewed by 537
Abstract
Wheat breeding programs are currently focusing on using non-destructive and cost-effective hyperspectral sensing tools to expeditiously and accurately phenotype large collections of genotypes. This approach is expected to accelerate the development of the abiotic stress tolerance of genotypes in breeding programs. This study [...] Read more.
Wheat breeding programs are currently focusing on using non-destructive and cost-effective hyperspectral sensing tools to expeditiously and accurately phenotype large collections of genotypes. This approach is expected to accelerate the development of the abiotic stress tolerance of genotypes in breeding programs. This study aimed to assess salt tolerance in wheat genotypes using non-destructive canopy spectral reflectance measurements as an alternative to direct laborious and time-consuming phenological selection criteria. Eight wheat genotypes and sixteen F8 RILs were tested under 150 mM NaCl in real field conditions for two years. Fourteen spectral reflectance indices (SRIs) were calculated from the spectral data, including vegetation SRIs and water SRIs. The effectiveness of these indices in assessing salt tolerance was compared with four morpho-physiological traits using genetic parameters, SSR markers, the Mantel test, hierarchical clustering heatmaps, stepwise multiple linear regression, and principal component analysis (PCA). The results showed significant differences (p ≤ 0.001) among RILs/cultivars for both traits and SRIs. The heritability, genetic gain, and genotypic and phenotypic coefficients of variability for most SRIs were comparable to those of measured traits. The SRIs effectively differentiated between salt-tolerant and sensitive genotypes and exhibited strong correlations with SSR markers (R2 = 0.56–0.89), similar to the measured traits and allelic data of 34 SSRs. A strong correlation (r = 0.27, p < 0.0001) was found between the similarity coefficients of SRIs and SSR data, which was higher than that between measured traits and SSR data (r = 0.20, p < 0.0003) based on the Mantel test. The PCA indicated that all vegetation SRIs and most water SRIs were grouped with measured traits in a positive direction and effectively identified the salt-tolerant RILs/cultivars. The PLSR models, which were based on all SRIs, accurately and robustly estimated the various morpho-physiological traits compared to using individual SRIs. The study suggests that various SRIs can be integrated with PLSR in wheat breeding programs as a cost-effective and non-destructive tool for phenotyping and screening large wheat populations for salt tolerance in a short time frame. This approach can replace the need for traditional morpho-physiological traits and accelerate the development of salt-tolerant wheat genotypes. Full article
(This article belongs to the Section Crop Physiology and Crop Production)
Show Figures

Figure 1

Figure 1
<p>Genotypic variability in four destructively measured traits among 24 RILs/cultivars under salinity conditions. PDW, LRWC, Chlt, and GY indicate plant dry weight (g plant<sup>−1</sup>), leaf relative water content (%), total chlorophyll content (mg g<sup>−1</sup> FW), and grain yield (ton ha<sup>−1</sup>), respectively. Bars with the same letters are not significantly different at the 0.05 level based on Tukey’s test. Data for each trait are the average of two seasons and three replicates, with bars representing the standard error (n = 6). The dark green color represents salt-tolerant genotypes; the red color represents salt-sensitive genotypes; the aqua color indicates moderately tolerant genotypes; the light green color represents 11 RILs resulting from the cross between salt-tolerant (Sakha 93) and salt-sensitive (Sakha 61) genotypes; and the blue color represents 5 RILs resulting from the cross between salt-tolerant and moderately salt-tolerant (Sids 1) genotypes.</p>
Full article ">Figure 2
<p>Genotypic variability in seven vegetation spectral reflectance indices among 24 RILs/cultivars under salinity conditions. The full names of indices are mentioned in Table 6 in the Materials and Methods Section. Bars with the same letters are not significantly different at the 0.05 level based on Tukey’s test. Data for each trait are the average of two seasons and three replicates, with bars representing the standard error (n = 6). The dark green color represents salt-tolerant genotypes; the red color represents salt-sensitive genotypes; the aqua color indicates moderately tolerant genotypes; the light green color represents 11 RILs resulting from the cross between salt-tolerant (Sakha 93) and salt-sensitive (Sakha 61) genotypes; and the blue color represents 5 RILs resulting from the cross between salt-tolerant and moderately salt-tolerant (Sids 1) genotypes.</p>
Full article ">Figure 3
<p>Genotypic variability in seven water spectral reflectance indices among 24 RILs/cultivars under salinity conditions. The full names of indices are mentioned in Table 6 in the Materials and Methods Section. Bars with the same letters are not significantly different at the 0.05 level based on Tukey’s test. Data for each trait are the average of two seasons and three replicates, with bars representing the standard error (n = 6). The dark green color represents salt-tolerant genotypes; the red color represents salt-sensitive genotypes; the aqua color indicates moderately tolerant genotypes; the light green color represents 11 RILs resulting from the cross between salt-tolerant (Sakha 93) and salt-sensitive (Sakha 61) genotypes; and the blue color represents 5 RILs resulting from the cross between salt-tolerant and moderately salt-tolerant (Sids 1) genotypes.</p>
Full article ">Figure 4
<p>Heatmap cluster analysis displaying the associations among 24 wheat RILs/cultivars based on four destructively measured traits (<b>A</b>) and different spectral reflectance indices (SRIs) (<b>B</b>). The different color schemes and densities were adjusted based on associations between RILs/cultivars, morpho-physiological traits, and SRIs. The darker red indicates lower values, while the darker blue indicates higher values. PDW, LRWC, Chlt, and GY indicate plant dry weight (g plant<sup>−1</sup>), leaf relative water content (%), total chlorophyll content (mg g<sup>−1</sup> FW), and grain yield (ton ha<sup>−1</sup>), respectively. The full names of the abbreviations for the different SRIs are listed in Table 6 in the Materials and Methods Section.</p>
Full article ">Figure 5
<p>Unweighted neighbor-joining clustering trees of 24 RILs/cultivars based on 34 simple sequence repeat (SSR) markers linked to salt-tolerant genes. The blue lines represent the salt tolerant group, red lines represent moderately salt-tolerant group, and green lines showed the salt sensitive group.</p>
Full article ">Figure 6
<p>Principal component analysis biplot for 24 RILs/cultivars, morpho-physiological traits, vegetation SRIs, and water SRIs. PDW, LRWC, Chlt, and GY indicate plant dry weight (g plant<sup>−1</sup>), leaf relative water content (%), total chlorophyll content (mg g<sup>−1</sup> FW), and grain yield (ton ha<sup>−1</sup>), respectively. The full names of the abbreviations for the different SRIs are listed in Table 6 in the Materials and Methods Section.</p>
Full article ">
19 pages, 5717 KiB  
Article
Remote Prediction of Soybean Yield Using UAV-Based Hyperspectral Imaging and Machine Learning Models
by Adilson Berveglieri, Nilton Nobuhiro Imai, Fernanda Sayuri Yoshino Watanabe, Antonio Maria Garcia Tommaselli, Glória Maria Padovani Ederli, Fábio Fernandes de Araújo, Gelci Carlos Lupatini and Eija Honkavaara
AgriEngineering 2024, 6(3), 3242-3260; https://doi.org/10.3390/agriengineering6030185 - 9 Sep 2024
Viewed by 736
Abstract
Early soybean yield estimation has become a fundamental tool for market policy and food security. Considering a heterogeneous crop, this study investigates the spatial and spectral variability in soybean canopy reflectance to achieve grain yield estimation. Besides allowing crop mapping, remote sensing data [...] Read more.
Early soybean yield estimation has become a fundamental tool for market policy and food security. Considering a heterogeneous crop, this study investigates the spatial and spectral variability in soybean canopy reflectance to achieve grain yield estimation. Besides allowing crop mapping, remote sensing data also provide spectral evidence that can be used as a priori knowledge to guide sample collection for prediction models. In this context, this study proposes a sampling design method that distributes sample plots based on the spatial and spectral variability in vegetation spectral indices observed in the field. Random forest (RF) and multiple linear regression (MLR) approaches were applied to a set of spectral bands and six vegetation indices to assess their contributions to the soybean yield estimates. Experiments were conducted with a hyperspectral sensor of 25 contiguous spectral bands, ranging from 500 to 900 nm, carried by an unmanned aerial vehicle (UAV) to collect images during the R5 soybean growth stage. The tests showed that spectral indices specially designed from some bands could be adopted instead of using multiple bands with MLR. However, the best result was obtained with RF using spectral bands and the height attribute extracted from the photogrammetric height model. In this case, Pearson’s correlation coefficient was 0.91. The difference between the grain yield productivity estimated with the RF model and the weight collected at harvest was 1.5%, indicating high accuracy for yield prediction. Full article
Show Figures

Figure 1

Figure 1
<p>Study area with soybean planting delimited by the red line—located in the western region of the state of São Paulo, Brazil. Image sources: Google Earth (on the <b>left</b>) and RGB image collected on 25 January 2020, with a UAV Phantom 3—DJI (on the <b>right</b>).</p>
Full article ">Figure 2
<p>Timeline presenting the progress of data collection in the study area.</p>
Full article ">Figure 3
<p>(<b>a</b>) FPI camera; (<b>b</b>) UAV in operation to acquire hyperspectral images.</p>
Full article ">Figure 4
<p>Image block adjustment of a hyperspectral band, showing two flying strips and six GCPs.</p>
Full article ">Figure 5
<p>(<b>a</b>) DTM. (<b>b</b>) Example of a cross-section in the CHM.</p>
Full article ">Figure 6
<p>Positions selected for 30 sample plots following the variability evidenced by the spectral indices: (<b>a</b>) NDVI; (<b>b</b>) SR; and (<b>c</b>) TCARI.</p>
Full article ">Figure 7
<p>(<b>a</b>) Descriptive statistics; (<b>b</b>) confidence interval; (<b>c</b>) normality test chart (red dots are the observed values).</p>
Full article ">Figure 8
<p>Grain yield production in kg/ha estimated from each sample plot, following the spatial distribution of <a href="#agriengineering-06-00185-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 9
<p>Comparison of results with the RF model when the height attribute is used with each vegetation index.</p>
Full article ">Figure 10
<p>Maps of grain productivity generated by the predictor models: (<b>a</b>) RF and (<b>b</b>) MLR.</p>
Full article ">
19 pages, 8433 KiB  
Article
Validation of In-House Imaging System via Code Verification on Petunia Images Collected at Increasing Fertilizer Rates and pHs
by Kahlin Wacker, Changhyeon Kim, Marc W. van Iersel, Mark Haidekker, Lynne Seymour and Rhuanito Soranz Ferrarezi
Sensors 2024, 24(17), 5809; https://doi.org/10.3390/s24175809 - 6 Sep 2024
Viewed by 629
Abstract
In a production environment, delayed stress recognition can impact yield. Imaging can rapidly and effectively quantify stress symptoms using indexes such as normalized difference vegetation index (NDVI). Commercial systems are effective but cannot be easily customized for specific applications, particularly post-processing. We developed [...] Read more.
In a production environment, delayed stress recognition can impact yield. Imaging can rapidly and effectively quantify stress symptoms using indexes such as normalized difference vegetation index (NDVI). Commercial systems are effective but cannot be easily customized for specific applications, particularly post-processing. We developed a low-cost customizable imaging system and validated the code to analyze images. Our objective was to verify the image analysis code and custom system could successfully quantify the changes in plant canopy reflectance. ‘Supercascade Red’, ‘Wave© Purple’, and ‘Carpet Blue’ Petunias (Petunia × hybridia) were transplanted individually and subjected to increasing fertilizer treatments and increasing substrate pH in a greenhouse. Treatments for the first trial were the addition of a controlled release fertilizer at six different rates (0, 0.5, 1, 2, 4, and 8 g/pot), and for the second trial, fertilizer solution with four pHs (4, 5.5, 7, and 8.5), with eight replications with one plant each. Plants were imaged twice a week using a commercial imaging system for fertilizer and thrice a week with the custom system for pH. The collected images were analyzed using an in-house program that calculated the indices for each pixel of the plant area. All cultivars showed a significant effect of fertilizer on the projected canopy size and dry weight of the above-substrate biomass and the fertilizer rate treatments (p < 0.01). Plant tissue nitrogen concentration as a function of the applied fertilizer rate showed a significant positive response for all three cultivars (p < 0.001). We verified that the image analysis code successfully quantified the changes in plant canopy reflectance as induced by increasing fertilizer application rate. There was no relationship between the pH and NDVI values for the cultivars tested (p > 0.05). Manganese and phosphorus had no significance with chlorophyll fluorescence for ‘Carpet Blue’ and ‘Wave© Purple’ (p > 0.05), though ‘Supercascade Red’ was found to have significance (p < 0.01). pH did not affect plant canopy size. Chlorophyll fluorescence pixel intensity against the projected canopy size had no significance except in ‘Wave© Purple’ (p = 0.005). NDVI as a function of the projected canopy size had no statistical significance. We verified the ability of the imaging system with integrated analysis to quantify nutrient deficiency-induced variability in plant canopies by increasing pH levels. Full article
(This article belongs to the Section Smart Agriculture)
Show Figures

Figure 1

Figure 1
<p>Flowchart diagram of the in-house imaging system to capture and analyze plant images under different light-emitting diodes (LEDs) wavelengths using chlorophyll fluorescence imaging to calculate spatial NDVI and canopy size per pixel for detailed plant analysis.</p>
Full article ">Figure 2
<p>Details of each image obtained by the imaging system, histogram representation, and normalized difference vegetation (NDVI) and anthocyanin content index (ACI) false color images.</p>
Full article ">Figure 3
<p>Projected canopy size of three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown under increasing fertilizer rates. The fertilizer rate applied has a significant effect on the two-dimensional area of the plant, as measured by a commercial imaging system and analyzed by our in-house software. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 4
<p>Dry mass of three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown under increasing fertilizer rates. All cultivars show significance in the treatments. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 5
<p>Nitrogen concentration as a function of increasing fertilizer rate on three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia). The nitrogen concentration was shown to be significantly related to the fertilizer rate. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 6
<p>Normalized difference vegetation index (NDVI) from the imaging system for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) at increasing fertilizer application rates. The normalized difference vegetation index (NDVI) responses are shown to be significantly related to fertilizer application. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 7
<p>Plant tissue nitrogen concentration as a function of the average pixel normalized difference vegetation index (NDVI) of the plant area for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) subjected to increasing fertilizer rates.</p>
Full article ">Figure 8
<p>Projected canopy size in pixels against the tissue nitrogen concentration for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown at increasing fertilizer rates. Primarily, this shows the effect of nitrogen concentration on the plant growth size.</p>
Full article ">Figure 9
<p>Dry biomass as a function of the projected canopy size for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown at increasing fertilizer rates. This shows the correlation between the imaged plant size and the dry mass.</p>
Full article ">Figure 10
<p>Projected canopy as a function of normalized difference vegetation index (NDVI), both from imaging system for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown at increasing fertilizer rates.</p>
Full article ">Figure 11
<p>(<b>A</b>) Phosphorus and (<b>B</b>) Manganese concentrations of ‘Supercascade Red’ in response to increasing pH. These nutrient decreases in the plant tissue were the desired effect in the experiment to display deficiencies or other visible symptoms. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 12
<p>Normalized difference vegetation index (NDVI) response to pH for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. Normalized difference vegetation index (NDVI) did not show a meaningful response to pH, except for ‘Supercascade Red’, which could be considered significant due to several extreme outliers. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 13
<p>Manganese content against chlorophyll fluorescence for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. There was no significant effect of Manganese on image-measured parameters on ‘Carpet Blue’ and ‘Wave© Purple’ cultivars.</p>
Full article ">Figure 14
<p>Phosphorus concentration against chlorophyll fluorescence for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. Stronger effect with phosphorus, explained by phosphorus being a macronutrient rather than a micronutrient.</p>
Full article ">Figure 15
<p>Average chlorophyll fluorescence pixel intensity as a function of pH for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 16
<p>Projected canopy size as a function of the pH treatments for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 17
<p>Chlorophyll fluorescence as a function of the projected canopy size for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions.</p>
Full article ">Figure 18
<p>Normalized difference vegetation index (NDVI) plotted against the projected canopy size for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions.</p>
Full article ">
25 pages, 5178 KiB  
Article
Sugarcane Mosaic Virus Detection in Maize Using UAS Multispectral Imagery
by Noah Bevers, Erik W. Ohlson, Kushal KC, Mark W. Jones and Sami Khanal
Remote Sens. 2024, 16(17), 3296; https://doi.org/10.3390/rs16173296 - 5 Sep 2024
Viewed by 600
Abstract
One of the most important and widespread corn/maize virus diseases is maize dwarf mosaic (MDM), which can be induced by sugarcane mosaic virus (SCMV). This study explores a machine learning analysis of five-band multispectral imagery collected via an unmanned aerial system (UAS) during [...] Read more.
One of the most important and widespread corn/maize virus diseases is maize dwarf mosaic (MDM), which can be induced by sugarcane mosaic virus (SCMV). This study explores a machine learning analysis of five-band multispectral imagery collected via an unmanned aerial system (UAS) during the 2021 and 2022 seasons for SCMV disease detection in corn fields. The three primary objectives are to (i) determine the spectral bands and vegetation indices that are most important or correlated with SCMV infection in corn, (ii) compare spectral signatures of mock-inoculated and SCMV-inoculated plants, and (iii) compare the performance of four machine learning algorithms, including ridge regression, support vector machine (SVM), random forest, and XGBoost, in predicting SCMV during early and late stages in corn. On average, SCMV-inoculated plants had higher reflectance values for blue, green, red, and red-edge bands and lower reflectance for near-infrared as compared to mock-inoculated samples. Across both years, the XGBoost regression model performed best for predicting disease incidence percentage (R2 = 0.29, RMSE = 29.26), and SVM classification performed best for the binary prediction of SCMV-inoculated vs. mock-inoculated samples (72.9% accuracy). Generally, model performances appeared to increase as the season progressed into August and September. According to Shapley additive explanations (SHAP analysis) of the top performing models, the simplified canopy chlorophyll content index (SCCCI) and saturation index (SI) were the vegetation indices that consistently had the strongest impacts on model behavior for SCMV disease regression and classification prediction. The findings of this study demonstrate the potential for the development of UAS image-based tools for farmers, aiming to facilitate the precise identification and mapping of SCMV infection in corn. Full article
(This article belongs to the Special Issue Crops and Vegetation Monitoring with Remote/Proximal Sensing II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Experimental layout in Synder field in 2021, overlaid on an aerial image captured on 14 July 2021. Red and blue boxes represent the randomized block design containing four single-row replicates of inoculated (red) and noninfected/mock-inoculated (blue) for each of the 51 hybrid varieties grown. (<b>b</b>) A corn crop that does not have SCMV infection or visual SCMV symptoms. (<b>c</b>) A corn crop that displays visual symptoms of SCMV infection.</p>
Full article ">Figure 2
<p>Experimental layout in Schaffter (<b>a</b>) and Synder (<b>b</b>) fields in 2022, overlaid on an aerial image captured on 1 September 2022, demonstrating randomized block designs with four 4-row replicates for each of the five hybrid varieties (including one control hybrid). Each box contains a unique plot number represented by a three-digit number, while the single-digit number corresponds to the hybrid group of that plot.</p>
Full article ">Figure 3
<p>Cumulative rainfall (mm), average daily air temperature (Celsius), and average daily relative humidity (%) for 2021 and 2022 corn field seasons (March–November). Rainfall is displayed as bar plots that accumulate value as season progresses. Data collected from Ohio State University CFAES Weather System at the Wooster Station.</p>
Full article ">Figure 4
<p>A flow diagram displaying the general methodological process implemented going from left to right. The best-performing regression models are identified for prediction of disease incidence percentage and binary classification prediction of disease presence. The best-performing models were further explored to identify the most important and most impactful features on the model’s performance. The number in parentheses for independent variables defines the number of variables included as a model input.</p>
Full article ">Figure 5
<p>Boxplots displaying (<b>A</b>) the disease incidence percentage values and (<b>B</b>) average yield values (bushels/acre), for each of the fields and years involved in the study for the SCMV-inoculated field plots. The solid black lines inside the plots represent medians, and the X marks represent the mean for each plot.</p>
Full article ">Figure 6
<p>Summary plot ((<b>a</b>), left) and bar plot ((<b>b</b>), right) visualize SHAP analysis for the XGBoost regression model for the prediction of disease incidence.</p>
Full article ">Figure 7
<p>Bar plot (<b>a</b>) depicts SHAP analysis for SVM classification based on 41 features, using ‘SCMV-inoculated’ and ‘mock-inoculated’ as categories. The confusion matrix (<b>b</b>) displays model performance metrics on a total of 267 test data samples. Numbers in green and red-orange boxes were correctly and incorrectly classified, respectively.</p>
Full article ">Figure 8
<p>Boxplots displaying the average values of the (<b>a</b>) simplified canopy chlorophyll content index (SCCCI) and (<b>b</b>) saturation index (SI) for SCMV-inoculated (red bars) and mock-inoculated (blue bars) samples across the growing season dates for both 2021 and 2022.</p>
Full article ">
16 pages, 2087 KiB  
Article
Dynamics of Mulatto Grass Regrowth Depending on Rotational Cattle Grazing Management
by Carlindo Santos Rodrigues, Márcia Cristina Teixeira da Silveira, Leandro Martins Barbero, Salim Jacaúna Sousa Júnior, Veridiana Aparecida Limão, Guilherme Pontes Silva, Sila Carneiro da Silva and Domicio do Nascimento Júnior
Grasses 2024, 3(3), 174-189; https://doi.org/10.3390/grasses3030013 - 3 Sep 2024
Viewed by 541
Abstract
This study was carried out to characterize the dynamics of forage accumulation during the regrowth of Mulatto grass submitted to rotational grazing strategies. The treatments corresponded to combinations between two pre-grazing conditions (95% and a maximum light interception during regrowth—LI95% and LI [...] Read more.
This study was carried out to characterize the dynamics of forage accumulation during the regrowth of Mulatto grass submitted to rotational grazing strategies. The treatments corresponded to combinations between two pre-grazing conditions (95% and a maximum light interception during regrowth—LI95% and LIMax) and two post-grazing conditions (post-grazing heights of 15 and 20 cm), according to a 2 × 2 factorial arrangement and randomized complete block design, with four replications. Rates of leaf growth (LGR), stems growth (SGR), total growth (TGR), leaf senescence (LSR), grass accumulation (GAR) (kg·ha−1·day−1), and the senescence/canopy growth ratio during different stages of regrowth. There was no difference between the management strategies for TGR. However, a higher GAR was reported for pastures managed with LI95% relative to LIMax, of 161.7 and 120.2 kg DM ha−1·day−1, respectively. Pastures managed with LI95% have a lower SGR in the intermediate and final regrowth period, reflecting the efficient control in the stalks production. On the other hand, in pastures managed, the LIMax showed higher SGR and LSR in the final regrowth phase. Thus, the LAI was higher in pastures managed at LI95% compared to those managed at LIMax, of 163.9 and 112.7 kg DM ha−1·day−1, respectively. Mulatto grass pastures, which were managed at LI95% pre-grazing, corresponded to approximately 30 cm in height, showed higher LAI, and ensured a low SGR throughout the regrowth period, constituting a more efficient management strategy. Full article
Show Figures

Figure 1

Figure 1
<p>Monthly averages of precipitation, maximum temperature, average temperature, and minimum temperature, January to April 2009, in the experimental area.</p>
Full article ">Figure 2
<p>Monthly water balance extract in the experimental area from January to April 2009.</p>
Full article ">Figure 3
<p>Relationship between light interception and leaf area index during regrowth depending on rotational cattle grazing management.</p>
Full article ">Figure 4
<p>Pre-grazing height during regrowth of Mulatto grass depending on rotational cattle grazing management characterized by pre-grazing targets LI<sub>95%</sub> and LI<sub>Max</sub> from January to April 2009. Averages followed by the same letter do not differ from each other (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 5
<p>Total growth rate (TGR) during regrowth of Mulatto depending on rotational cattle grazing management. Averages followed by the same letter do not differ from each other (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 6
<p>Leaf growth rate (LGR) of Mulatto grass managed with pre-grazing targets LI<sub>95%</sub> and LI<sub>Max</sub> (<b>A</b>) during the regrowth period, (<b>B</b>) depending on rotational cattle grazing management. Averages followed by the same letter do not differ from each other (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 7
<p>Mulatto grass accumulation rate (GAR) managed with pre-grazing targets LI<sub>95%</sub> and LI<sub>Max</sub> (<b>A</b>) during the regrowth period (<b>B</b>) depending on rotational cattle grazing management. Averages followed by the same letter did not differ (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">
24 pages, 6269 KiB  
Article
Aboveground Biomass Mapping in SemiArid Forests by Integrating Airborne LiDAR with Sentinel-1 and Sentinel-2 Time-Series Data
by Linjing Zhang, Xinran Yin, Yaru Wang and Jing Chen
Remote Sens. 2024, 16(17), 3241; https://doi.org/10.3390/rs16173241 - 1 Sep 2024
Viewed by 834
Abstract
Aboveground biomass (AGB) is a vital indicator for studying carbon sinks in forest ecosystems. Semiarid forests harbor substantial carbon storage but received little attention due to the high spatial–temporal heterogeneity that complicates the modeling of AGB in this environment. This study assessed the [...] Read more.
Aboveground biomass (AGB) is a vital indicator for studying carbon sinks in forest ecosystems. Semiarid forests harbor substantial carbon storage but received little attention due to the high spatial–temporal heterogeneity that complicates the modeling of AGB in this environment. This study assessed the performance of different data sources (annual monthly time-series radar was Sentinel-1 [S1]; annual monthly time series optical was Sentinel-2 [S2]; and single-temporal airborne light detection and ranging [LiDAR]) and seven prediction approaches to map AGB in the semiarid forests on the border between Gansu and Qinghai Provinces in China. Five experiments were conducted using different data configurations from synthetic aperture radar backscatter, multispectral reflectance, LiDAR point cloud, and their derivatives (polarimetric combination indices, texture information, vegetation indices, biophysical features, and tree height- and canopy-related indices). The results showed that S2 acquired better prediction (coefficient of determination [R2]: 0.62–0.75; root mean square error [RMSE]: 30.08–38.83 Mg/ha) than S1 (R2: 0.24–0.45; RMSE: 47.36–56.51 Mg/ha). However, their integration further improved the results (R2: 0.65–0.78; RMSE: 28.68–35.92 Mg/ha). The addition of single-temporal LiDAR highlighted its structural importance in semiarid forests. The best mapping accuracy was achieved by XGBoost, with the metrics from the S2 and S1 time series and the LiDAR-based canopy height information being combined (R2: 0.87; RMSE: 21.63 Mg/ha; relative RMSE: 14.45%). Images obtained during the dry season were effective for AGB prediction. Tree-based models generally outperformed other models in semiarid forests. Sequential variable importance analysis indicated that the most important S1 metric to estimate AGB was the polarimetric combination indices sum, and the S2 metrics were associated with red-edge spectral regions. Meanwhile, the most important LiDAR metrics were related to height percentiles. Our methodology advocates for an economical, extensive, and precise AGB retrieval tailored for semiarid forests. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Study area: (<b>a</b>) a geographical position map of Qinghai and Gansu Provinces in China; (<b>b</b>) a geographical position map of the study area at the border between Qinghai and Gansu Provinces; and (<b>c</b>) a true color image of the study area formed by clipping an S2 image acquired on 11 August 2019.</p>
Full article ">Figure 2
<p>Data-acquisition timeline: (<b>a</b>) time of acquisition for the S1 and S2 single image; (<b>b</b>) time of acquisition for the synthetic cloud-free S2 images from May to Jul 2019; (<b>c</b>) time of acquisition for the synthetic cloud-free S2 images from May to Jul 2020; and (<b>d</b>) time of acquisition for the synthetic cloud-free S2 images from May to Jul 2021.</p>
Full article ">Figure 3
<p>Flowchart of the study.</p>
Full article ">Figure 4
<p>Accuracy evaluation for different models and data sources. The bar represents R<sup>2</sup>, and the line represents RMSE<sub>r</sub>.</p>
Full article ">Figure 5
<p>Scatter plots of the measured AGB and predicted AGB from the best performance combination (experiment E-S1S2LiDAR) for different models (<b>a</b>) RF; (<b>b</b>) XGBoost; (<b>c</b>) SGB; (<b>d</b>) CNN; (<b>e</b>) GPR; (<b>f</b>) MLP; (<b>g</b>) LASSO.</p>
Full article ">Figure 6
<p>Flow chart of the sequential forward selection.</p>
Full article ">Figure 7
<p>Variable importance plots of the top 15 predictors for the XGBoost model with three combined datasets: (<b>a</b>) experiment C; (<b>b</b>) experiment D; (<b>c</b>) experiment E.</p>
Full article ">Figure 8
<p>AGB map in the study area from the XGBoost model with the optical, SAR, and LiDAR metrics.</p>
Full article ">
Back to TopTop