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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,289)

Search Parameters:
Keywords = enhanced vegetation index

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
22 pages, 7624 KiB  
Article
Quantitative Assessment of Urban Expansion Impact on Vegetation in the Lanzhou–Xining Urban Agglomeration
by Wensheng Wang, Wenfei Luan, Haitao Jing, Jingyao Zhu, Kaixiang Zhang, Qingqing Ma, Shiye Zhang and Xiujuan Liang
Appl. Sci. 2024, 14(19), 8615; https://doi.org/10.3390/app14198615 (registering DOI) - 24 Sep 2024
Abstract
The Rapid expansion of the Lanzhou–Xining (Lanxi) urban cluster in China during recent decades poses a threat to the fragile arid environment. Quantitatively assessing the impact of urban expansion on vegetation in the Lanxi urban cluster has profound implications for future sustainable urban [...] Read more.
The Rapid expansion of the Lanzhou–Xining (Lanxi) urban cluster in China during recent decades poses a threat to the fragile arid environment. Quantitatively assessing the impact of urban expansion on vegetation in the Lanxi urban cluster has profound implications for future sustainable urban planning. This study investigated the urban expansion dynamics of the Lanxi urban cluster and its impacts on regional vegetation between 2001 and 2021 based on time series land cover data and auxiliary remote sensing data, such as digital elevation model (DEM) data, nighttime light data, and administrative boundary data. Thereinto, urban expansion dynamics were evaluated using the annual China Land Cover Dataset (CLCD, 2001–2021). Urban expansion impacts on regional vegetation were assessed via the Vegetation Disturbance Index (VDI), an index capable of quantitatively assessing the positive and negative impacts of urban expansion at the pixel level, which can be obtained by overlaying the Enhanced Vegetation Index (EVI) and rainfall data. The major findings indicate that: (1) Over the past two decades, the Lanxi region has experienced rapid urban expansion, with the built-up area expanding from 183.50 km2 to 294.30 km2, which is an average annual expansion rate of 2.39%. Notably, Lanzhou, Baiyin, and Xining dominated the expansion. (2) Urban expansion negatively affected approximately 53.50 km2 of vegetation, while about 39.56 km2 saw positive impacts. The negative effects were mainly due to the loss of cropland and grassland. Therefore, cities in drylands should balance urban development and vegetation conservation by strictly controlling cropland and grassland occupancy and promoting intelligent urban growth. Full article
(This article belongs to the Section Ecology Science and Engineering)
Show Figures

Figure 1

Figure 1
<p>The Lanxi urban cluster. (<b>a</b>) Mean precipitation of the Lanxi urban cluster from 2000 to 2020. (<b>b</b>) The proportion of different land cover types in the Lanxi urban cluster in 2021.</p>
Full article ">Figure 2
<p>Workflow for assessing the impact of urban expansion on vegetation.</p>
Full article ">Figure 3
<p>Urban expansion dynamics in the Lanxi urban cluster from 2001 to 2021.</p>
Full article ">Figure 4
<p>Land cover conversion from 2001 to 2021 (km<sup>2</sup>).</p>
Full article ">Figure 5
<p>The disturbed area and distribution of vegetation in the Lansi urban cluster (km<sup>2</sup>). (<b>a</b>) Areas of each district/county positively affected, (<b>b</b>) Total area with positive impact in the region, (<b>c</b>) Areas of each district/county negatively affected, (<b>d</b>) Total area with negative impact in the region.</p>
Full article ">Figure 6
<p>Land cover changes between the expansion zones of the Lanxi urban cluster and other land types from 2001 to 2021.</p>
Full article ">Figure 7
<p>Urban expansion occupies various types of land area (km<sup>2</sup>).</p>
Full article ">
16 pages, 13129 KiB  
Article
Disentangling the Effects of Atmospheric and Soil Dryness on Autumn Phenology across the Northern Hemisphere
by Kangbo Dong and Xiaoyue Wang
Remote Sens. 2024, 16(19), 3552; https://doi.org/10.3390/rs16193552 - 24 Sep 2024
Abstract
In recent decades, drought has intensified along with continuous global warming, significantly impacting terrestrial vegetation. High atmospheric water demand, indicated by vapor pressure deficit (VPD), and insufficient soil moisture (SM) are considered the primary factors causing drought stress in vegetation. However, the influences [...] Read more.
In recent decades, drought has intensified along with continuous global warming, significantly impacting terrestrial vegetation. High atmospheric water demand, indicated by vapor pressure deficit (VPD), and insufficient soil moisture (SM) are considered the primary factors causing drought stress in vegetation. However, the influences of VPD and SM on the autumn phenology are still unknown. Using satellite observations and meteorological data, we examined the impacts of VPD and SM on the end of the growing season (EOS) across the Northern Hemisphere (>30°N) from 1982 to 2022. We found that VPD and SM were as important as temperature, precipitation, and radiation in controlling the variations in the EOS. Moreover, the EOS was predominantly influenced by VPD or SM in more than one-third (33.8%) of the study area. In particular, a ridge regression analysis indicated that the EOS was more sensitive to VPD than to SM and the other climatic factors, with 25% of the pixels showing the highest sensitivity to VPD. In addition, the effects of VPD and SM on the EOS varied among biome types and climate zones. VPD significantly advanced the EOS in 25.8% of temperate grasslands, while SM had the greatest impact on advancing the EOS in 17.7% of temperate coniferous forests. Additionally, 27.7% of the midlatitude steppe (BSk) exhibited a significant negative correlation between VPD and the EOS, while 19.4% of the marine west coast climate (Cfb) showed a positive correlation between SM and the EOS. We also demonstrated that the correlation between VPD and the EOS was linearly affected by VPD and the leaf area index, while the correlation between SM and the EOS was affected by SM, precipitation, and the leaf area index. Our study highlights the importance of VPD and SM in regulating autumn phenology and enhances our understanding of terrestrial ecosystem responses to climate change. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Biome type (<span class="html-italic">Terrestrial Ecoregions of the World</span>) (&gt;30°N). (<b>b</b>) Köppen–Geiger climate zones (&gt;30°N).</p>
Full article ">Figure 2
<p>An illustration of the method used to extract the end of the growing season (EOS) from NDVI data.</p>
Full article ">Figure 3
<p>Spatial distributions of the mean (<b>a</b>,<b>e</b>,<b>i</b>), standard deviation (<b>b</b>,<b>f</b>,<b>j</b>), slope (<b>c</b>,<b>g</b>,<b>k</b>), and significance level (<b>d</b>,<b>h</b>,<b>l</b>) of the EOS, SM, and VPD from 1982 to 2022, respectively.</p>
Full article ">Figure 4
<p>Spatial pattern of partial correlation coefficients between factors and EOS: (<b>a</b>) averaged temperature, (<b>b</b>) soil moisture, (<b>c</b>) precipitation, (<b>d</b>) vapor pressure deficit, (<b>e</b>) solar radiation, and (<b>f</b>) the most dominant factor. The gray areas represent non-significant pixels. The bars on the left express the proportions of the partial correlation coefficients between the EOS and corresponding factors.</p>
Full article ">Figure 5
<p>Spatial pattern of ridge regression coefficients between climatic factors and the EOS: (<b>a</b>) averaged temperature, (<b>b</b>) soil moisture, (<b>c</b>) precipitation, (<b>d</b>) vapor pressure deficit, (<b>e</b>) solar radiation, and (<b>f</b>) the most dominant factor. The bars on the left are the frequencies of the ridge regression coefficients between the EOS and the corresponding factors.</p>
Full article ">Figure 6
<p>Partial correlation coefficients between the EOS and factors of each biome type. The bars above 0 represent percentages of positive correlations, and the others show negative percentages. Colored areas indicate significant percentages (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Partial correlation coefficients between the EOS and factors of each climate zone type. The bars above 0 represent percentages of positive correlations, and the others show negative percentages. Colored areas indicate significant percentages (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Average EOS in every percentile bin of SM and VPD.</p>
Full article ">Figure 9
<p>The correlations between the coefficients of partial correlation between EOS and SM (<b>a</b>,<b>c</b>), VPD (<b>b</b>,<b>d</b>), and the variables (LAI, LAI, SM, and VPD), respectively.</p>
Full article ">
26 pages, 6130 KiB  
Article
Comprehensive Spatial-Temporal and Risk Factor Insights for Optimizing Livestock Anthrax Vaccination Strategies in Karnataka, India
by Jayashree Anandakumar, Kuralayanapalya Puttahonnappa Suresh, Archana Veeranagouda Patil, Chethan A. Jagadeesh, Sushma Bylaiah, Sharanagouda S. Patil and Divakar Hemadri
Vaccines 2024, 12(9), 1081; https://doi.org/10.3390/vaccines12091081 - 22 Sep 2024
Viewed by 412
Abstract
Anthrax, a zoonotic disease affecting both livestock and humans globally, is caused by Bacillus anthracis. The objectives of this study were the following: (1) to identify environmental risk factors for anthrax and use this information to develop an improved predictive risk map, and [...] Read more.
Anthrax, a zoonotic disease affecting both livestock and humans globally, is caused by Bacillus anthracis. The objectives of this study were the following: (1) to identify environmental risk factors for anthrax and use this information to develop an improved predictive risk map, and (2) to estimate spatial variation in basic reproduction number (Ro) and herd immunity threshold at the village level, which can be used to optimize vaccination policies within high-risk regions. Based on the anthrax incidences from 2000–2023 and vaccine administration figures between 2008 and 2022 in Karnataka, this study depicted spatiotemporal pattern analysis to derive a risk map employing machine learning algorithms and estimate Ro and herd immunity threshold for better vaccination coverage. Risk factors considered were key meteorological, remote sensing, soil, and geographical parameters. Spatial autocorrelation and SaTScan analysis revealed the presence of hotspots and clusters predominantly in the southern, central, and uppermost northern districts of Karnataka and temporal cluster distribution between June and September. Factors significantly associated with anthrax were air temperature, surface pressure, land surface temperature (LST), enhanced vegetation index (EVI), potential evapotranspiration (PET), soil temperature, soil moisture, pH, available potassium, sulphur, and boron, elevation, and proximity to waterbodies and waterways. Ensemble technique with random forest and classification tree models were used to improve the prediction accuracy of anthrax. High-risk areas are expected in villages in the southern, central, and extreme northern districts of Karnataka. The estimated Ro revealed 11 high-risk districts with Ro > 1.50 and respective herd immunity thresholds ranging from 11.24% to 55.47%, and the assessment of vaccination coverage at the 70%, 80%, and 90% vaccine efficacy levels, all serving for need-based strategic vaccine allocation. A comparison analysis of vaccinations administered and vaccination coverage estimated in this study is used to illustrate difference in the supply and vaccine force. The findings from the present study may support in planning preventive interventions, resource allocation, especially of vaccines, and other control strategies against anthrax across Karnataka, specifically focusing on predicted high-risk regions. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Spatial map representing cumulative anthrax incidences recorded in Karnataka between 2000 and 2023, (<b>B</b>) Actual geolocations of villages where anthrax incidences were recorded between 2000 and 2023.</p>
Full article ">Figure 2
<p>(<b>A</b>) Annual number of livestock anthrax incidences recorded in Karnataka during 2000–2023; (<b>B</b>) Cumulative monthly anthrax incidences recorded in Karnataka during 2000–2023.</p>
Full article ">Figure 3
<p>Hotspot map of anthrax in Karnataka identified through spatial autocorrelation.</p>
Full article ">Figure 4
<p>The map depicting anthrax incidence points plotted over (<b>A</b>) Elevation, (<b>B</b>) Roadways, and (<b>C</b>) Waterways.</p>
Full article ">Figure 5
<p>Predicted village-level risk map of anthrax in Karnataka showing variations in color, ranging from yellow (low risk) to red (high risk).</p>
Full article ">Figure 6
<p>Assessment of vaccination coverage at different vaccine efficacy levels.</p>
Full article ">Figure 7
<p>Vaccination supply trend in Karnataka (2008–2022).</p>
Full article ">Figure 8
<p>Differences in vaccines supplied vs. vaccines required to achieve herd immunity thresholds in high-risk anthrax districts in Karnataka.</p>
Full article ">
21 pages, 8374 KiB  
Article
Response of Fuel Characteristics, Potential Fire Behavior, and Understory Vegetation Diversity to Thinning in Platycladus orientalis Forest in Beijing, China
by Min Gao, Sifan Chen, Aoli Suo, Feng Chen and Xiaodong Liu
Forests 2024, 15(9), 1667; https://doi.org/10.3390/f15091667 - 22 Sep 2024
Viewed by 246
Abstract
Objective: Active fuel management operations, such as thinning, can minimize extreme wildfire conditions while preserving ecosystem services, including maintaining understory vegetation diversity. However, the appropriate thinning intensity for balancing the above two objectives has not been sufficiently studied. Methods: This study was conducted [...] Read more.
Objective: Active fuel management operations, such as thinning, can minimize extreme wildfire conditions while preserving ecosystem services, including maintaining understory vegetation diversity. However, the appropriate thinning intensity for balancing the above two objectives has not been sufficiently studied. Methods: This study was conducted to assess the impact of various thinning intensities (light thinning, LT, 15%; moderate thinning, MT, 35%; heavy thinning, HT, 50%; and control treatment, CK) on fuel characteristics, potential fire behavior, and understory vegetation biodiversity in Platycladus orientalis forest in Beijing using a combination of field measurements and fire behavior simulations (BehavePlus 6.0.0). Results: A significant reduction in surface and canopy fuel loads with increasing thinning intensity, notably reducing CBD to below 0.1 kg/m3 under moderate thinning, effectively prevented the occurrence of active crown fires, even under extreme weather conditions. Additionally, moderate thinning enhanced understory species diversity, yielding the highest species diversity index compared to other treatments. Conclusions: These findings suggest that moderate thinning (35%) offers an optimal balance, substantially reducing the occurrence of active crown fires while promoting biodiversity. Therefore, it is recommended to carry out moderate thinning in the study area. Forest managers can leverage this information to devise technical strategies that simultaneously meet fire prevention objectives and enhance understory vegetation species diversity in areas suitable for thinning-only treatments. Full article
(This article belongs to the Special Issue Fire Ecology and Management in Forest—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Location of the study site in Haidian District, Beijing, China.</p>
Full article ">Figure 2
<p>Standard branch diagram of <span class="html-italic">P. orientalis</span> canopy. Note: H = tree height, CL = crown length, H<sub>base</sub> = canopy base height, d= branch basal diameter.</p>
Full article ">Figure 3
<p>The process and result of fire type generation in BehavePlus 6.0.0. Note: Surface = surface fire; Torching = passive crown fire; Conditional Crown = active crown fire possible if the fire transitions to the overstory; Crowning = active crown fire.</p>
Full article ">Figure 4
<p>Mean surface fuel loads and depth under the four thinning intensity treatments at different times after thinning. Note: (<b>a</b>–<b>c</b>): downed and dead woody fuel loads; (<b>d</b>,<b>e</b>): live fuel loads; (<b>f</b>,<b>g</b>): litter leaves fuel load; (<b>h</b>): litter depth; (<b>i</b>): fuel bed depth. Different uppercase letters represent significant differences among thinning intensities (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Mean canopy fuel load (<b>a</b>) and canopy bulk density (<b>b</b>) under different thinning intensities at three time points after thinning. Different uppercase letters represent significant differences among thinning intensities (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Potential surface fire behavior indicators of <span class="html-italic">P. orientalis</span> stand under four thinning intensities. Note: (<b>a</b>,<b>d</b>,<b>g</b>): surface fire spread rate under three thinning years; (<b>b</b>,<b>e</b>,<b>h</b>): surface flame length under three thinning years; (<b>c</b>,<b>f</b>,<b>i</b>): heat per unit area under three thinning years.</p>
Full article ">Figure 7
<p>Potential crown fire behavior indicators of <span class="html-italic">P. orientalis</span> stand under four thinning intensities. Note: (<b>a</b>,<b>d</b>,<b>g</b>): crown flame length under three thinning years; (<b>b</b>,<b>e</b>,<b>h</b>): crown fireline intensity under three thinning years; (<b>c</b>,<b>f</b>,<b>i</b>): heat per unit area under three thinning years.</p>
Full article ">Figure 8
<p>Potential surface fire behavior indicators of <span class="html-italic">P. orientalis</span> stand under four thinning intensities. Note: (<b>A</b>,<b>D</b>,<b>G</b>): surface fireline intensity; (<b>B</b>,<b>E</b>,<b>H</b>): critical surface fireline intensity; (<b>C</b>,<b>F</b>,<b>I</b>): heat per unit area.</p>
Full article ">Figure 9
<p>Critical crown fire spread rate, crown fire spread rate, and the active ratio of <span class="html-italic">P. orientalis</span>.</p>
Full article ">
47 pages, 19713 KiB  
Article
Enhancing Drought Resilience through Groundwater Engineering by Utilizing GIS and Remote Sensing in Southern Lebanon
by Nasser Farhat
Hydrology 2024, 11(9), 156; https://doi.org/10.3390/hydrology11090156 - 21 Sep 2024
Viewed by 707
Abstract
Countries face challenges of excess, scarcity, pollution, and uneven water distribution. This study highlights the benefits of advances in groundwater engineering that improve the understanding of utilizing local geological characteristics due to their crucial role in resisting drought in southern Lebanon. The type [...] Read more.
Countries face challenges of excess, scarcity, pollution, and uneven water distribution. This study highlights the benefits of advances in groundwater engineering that improve the understanding of utilizing local geological characteristics due to their crucial role in resisting drought in southern Lebanon. The type of drought in the region was determined using the Standardized Precipitation Index (SPI), Standardized Vegetation Index (NDVI), Vegetation Condition Index (VCI), and Soil Moisture Anomaly Index (SM). The dry aquifer and its characteristics were analyzed using mathematical equations and established hydrogeological principles, including Darcy’s law. Additionally, a morphometric assessment of the Litani River was performed to evaluate its suitability for artificial recharge, where the optimal placement of the water barrier and recharge tunnels was determined using Spearman’s rank correlation coefficient. This analysis involved excluding certain parameters based on the Shapiro–Wilk test for normality. Accordingly, using the Geographic Information System (GIS), we modeled and simulated the potential water table. The results showed the importance and validity of linking groundwater engineering and morphometric characteristics in combating the drought of groundwater layers. The Eocene layer showed a clearer trend for the possibility of being artificially recharged from the Litani River than any other layer. The results showed that the proposed method can enhance artificial recharge, raise the groundwater level to four levels, and transform it into a large, saturated thickness. On the other hand, it was noted that the groundwater levels near the surface will cover most of the area of the studied region and could potentially store more than one billion cubic meters of water, mitigating the effects of climate change for decades. Full article
(This article belongs to the Section Surface Waters and Groundwaters)
Show Figures

Figure 1

Figure 1
<p>The study area in Lebanon.</p>
Full article ">Figure 2
<p>Annual rainfall in the study area.</p>
Full article ">Figure 3
<p>Trend in temperature increase.</p>
Full article ">Figure 4
<p>The lithological map of the study area.</p>
Full article ">Figure 5
<p>A geological section from east to west showing the fractures and the inclination of the geological strata in addition to the Bint Jbeil syncline. E2: Eocene; C6: Senonian; C4: Cenomanian.</p>
Full article ">Figure 6
<p>Stratum with the lineaments and faults in the study area.</p>
Full article ">Figure 7
<p>The topography of the region.</p>
Full article ">Figure 8
<p>The Standardized Precipitation Index (SPI) from 1981 to 2024.</p>
Full article ">Figure 9
<p>Vegetation Condition Index (VCI).</p>
Full article ">Figure 10
<p>Soil Moisture Anomaly Index (SM).</p>
Full article ">Figure 11
<p>DEM with dimension of the Eocene aquifer.</p>
Full article ">Figure 12
<p>Geological section of the syncline of the Eocene aquifer. C6 is the Senonian isolate layer; C4 is the Cenomanian aquifer; E2 is the Eocene saturated thickness.</p>
Full article ">Figure 13
<p>Aquifers in the study area.</p>
Full article ">Figure 14
<p>Hydrological map of the study area.</p>
Full article ">Figure 15
<p>Litani watershed in the study area.</p>
Full article ">Figure 16
<p>The Litani River in its valley and the optimal site of the dam.</p>
Full article ">Figure 16 Cont.
<p>The Litani River in its valley and the optimal site of the dam.</p>
Full article ">Figure 17
<p>The water table in the Eocene aquifer showing its relationship with the proposed tunnels.</p>
Full article ">Figure 18
<p>Topographic section from the dam to the southern part of the study area showing the tunnels and the water table.</p>
Full article ">Figure 19
<p>Regional depth of the water table following filling of the Litani River reservoir.</p>
Full article ">Figure 20
<p>This figure shows the potential groundwater level in a very small area within deep valleys where the water table is theoretically supposed to approach the ground surface.</p>
Full article ">Figure 21
<p>Areas where the water table will be at depths ranging from 1 to 150 m.</p>
Full article ">Figure 22
<p>Areas where the water table will be at depths ranging from 150 to 250 m.</p>
Full article ">Figure 23
<p>Areas where the water table will be at depths greater than 250 m.</p>
Full article ">Figure 24
<p>Calculation of the groundwater volume balance residual and seepage velocity vector (direction and magnitude) for steady flow in an aquifer by GIS. (<b>A</b>) Head elevation: the head elevation raster comes from various sources. It has been interpolated from borehole data using the surface interpolation tool Kriging. This head is consistent with the transmissivity raster and reflects its flow through its field. (<b>B</b>) Porosity is defined as the volume of void space that contributes to fluid flow divided by the entire volume. The effective porosity field, a physical property of the aquifer, is estimated from geological data. It was expressed as a value of around 35 to 41 percent of the volume of the porous medium contributing to fluid flow. (<b>C</b>) Transmissivity, measured in area square per day units, was interpreted from geological information. Transmissivity is the rate at which water is transmitted through a unit width of an aquifer under a unit hydraulic gradient. It is expressed as the product of the average hydraulic conductivity and thickness of the saturated portion of an aquifer. (<b>D</b>) The saturated thickness, measured in length units, was interpreted from geological information. The saturated thickness of an unconfined Eocene aquifer is the distance between the water table and the lower confining layer. (<b>E</b>) The magnitude is in units of length over time and is an optional output raster where each cell value represents the magnitude of the seepage velocity vector (average linear velocity) at the center of the cell. (<b>F</b>) The velocity vector’s direction is expressed in compass coordinates (degrees clockwise from north). Each cell value corresponds to the direction of the seepage velocity vector (average linear velocity) at the cell’s center. This is calculated as the average value of the seepage velocity through the four faces of the cell. (<b>G</b>) Darcy flow is the volume balance residual raster. Each cell value represents the groundwater volume balance residual for steady flow in an aquifer, as determined by Darcy’s Law.</p>
Full article ">Figure 24 Cont.
<p>Calculation of the groundwater volume balance residual and seepage velocity vector (direction and magnitude) for steady flow in an aquifer by GIS. (<b>A</b>) Head elevation: the head elevation raster comes from various sources. It has been interpolated from borehole data using the surface interpolation tool Kriging. This head is consistent with the transmissivity raster and reflects its flow through its field. (<b>B</b>) Porosity is defined as the volume of void space that contributes to fluid flow divided by the entire volume. The effective porosity field, a physical property of the aquifer, is estimated from geological data. It was expressed as a value of around 35 to 41 percent of the volume of the porous medium contributing to fluid flow. (<b>C</b>) Transmissivity, measured in area square per day units, was interpreted from geological information. Transmissivity is the rate at which water is transmitted through a unit width of an aquifer under a unit hydraulic gradient. It is expressed as the product of the average hydraulic conductivity and thickness of the saturated portion of an aquifer. (<b>D</b>) The saturated thickness, measured in length units, was interpreted from geological information. The saturated thickness of an unconfined Eocene aquifer is the distance between the water table and the lower confining layer. (<b>E</b>) The magnitude is in units of length over time and is an optional output raster where each cell value represents the magnitude of the seepage velocity vector (average linear velocity) at the center of the cell. (<b>F</b>) The velocity vector’s direction is expressed in compass coordinates (degrees clockwise from north). Each cell value corresponds to the direction of the seepage velocity vector (average linear velocity) at the cell’s center. This is calculated as the average value of the seepage velocity through the four faces of the cell. (<b>G</b>) Darcy flow is the volume balance residual raster. Each cell value represents the groundwater volume balance residual for steady flow in an aquifer, as determined by Darcy’s Law.</p>
Full article ">
24 pages, 20022 KiB  
Article
Variation in Vegetation Composition and Structure across Mudflat Areas in the Yellow River Delta, China
by He Li, Qingsheng Liu, Chong Huang, Xin Zhang, Shuxuan Wang, Wei Wu and Lei Shi
Remote Sens. 2024, 16(18), 3495; https://doi.org/10.3390/rs16183495 - 20 Sep 2024
Viewed by 235
Abstract
Variations in vegetation composition and structure are significant signals of the succession of mudflat ecosystems and have important indicative roles in mudflat ecosystem degradation. Due to poor accessibility of vast even mudflat areas, variation in vegetation composition and structure across mudflat areas remains [...] Read more.
Variations in vegetation composition and structure are significant signals of the succession of mudflat ecosystems and have important indicative roles in mudflat ecosystem degradation. Due to poor accessibility of vast even mudflat areas, variation in vegetation composition and structure across mudflat areas remains unclear in the Yellow River Delta (YRD), China. We provided an UAV multispectral orthomosaic with 10 cm ground sample distance to classify and compare the vegetation composition and structure across mudflat areas in the YRD. The vegetation classification overall accuracy achieved 95.0%. We found that although a significant difference (p < 0.05) was checked out in terms of the Shannon–Wiener diversity index (from 1.33 to 0.92) and evenness index (from 0.96 to 0.66) among the eight subareas from land to sea, all four dominant vegetation communities (S. salsa, L. bicolor, T. chinensis, and P. australis) were discovered at all eight subareas. Our findings support the idea that the regional environment and local microtopography are the predominant forces for variation in vegetation composition and structure across mudflat areas. From the perspective of vegetation restoration and conservation, changing the local microtopography will be an interesting way to enhance the vegetation diversity of the mudflat ecosystems in the YRD. Full article
Show Figures

Figure 1

Figure 1
<p>Study area. (<b>a</b>) Geographic location of the study area; (<b>b</b>) Sentinel-2 imagery of the study area acquired on 24 June 2024 and the eight subareas surrounded by the yellow line (from A to H) from land to sea.</p>
Full article ">Figure 2
<p>Photo of community <span class="html-italic">P. australis</span>, <span class="html-italic">T. chinensis</span>, and <span class="html-italic">S. salsa</span>. Taken on 1 November 2023.</p>
Full article ">Figure 3
<p>The roads, tidal creeks, or ditches in the study area.</p>
Full article ">Figure 4
<p>The vegetation composition and the UAV multispectral orthomosaic imagery of Subarea A. (<b>a</b>) The vegetation composition; (<b>b</b>) the UAV multispectral orthomosaic RGB true color composite imagery.</p>
Full article ">Figure 5
<p>Statistics on vegetation composition and structure across mudflat areas from Subarea A to Subarea H. (<b>a</b>) Number of patches (N), number of vegetation patches (Nv), number of community <span class="html-italic">S. Salsa</span> (Ns)<span class="html-italic">,</span> number of community <span class="html-italic">T. chinensis</span> (Nt), number of community <span class="html-italic">P. australis</span> (Nph), number of community <span class="html-italic">S. Salsa</span> + community <span class="html-italic">L. bicolor</span> (Nsl), and number of vegetation patches per 100 hectares (PD); (<b>b</b>) Total area of patches (A), total area of vegetation patches (Av), total area of community <span class="html-italic">S. Salsa</span> (As), total area of community <span class="html-italic">T. chinensis</span> (At), total area of community <span class="html-italic">P. australis</span> (Aph), and total area of community <span class="html-italic">S. Salsa</span> + community <span class="html-italic">L. bicolor</span> (Asl); (<b>c</b>) Mean area of vegetation patches (MAv),mean area of community <span class="html-italic">S. Salsa</span> (MAs), mean area of community <span class="html-italic">T. chinensis</span> (MAt), mean area of community <span class="html-italic">P. australis</span> (MAph), and mean area of community <span class="html-italic">S. Salsa</span> + community <span class="html-italic">L. bicolor</span> (MAsl); (<b>d</b>) Vegetation cover (VC), vegetation cover of community <span class="html-italic">S. Salsa</span> (VCs), vegetation cover of community <span class="html-italic">T. chinensis</span> (VCt), vegetation cover of community <span class="html-italic">P. australis</span> (VCph), vegetation cover of community <span class="html-italic">S. Salsa</span> + community <span class="html-italic">L. bicolor</span> (VCsl), relative dominance of community <span class="html-italic">S. Salsa</span> (RDOs), relative dominance of community <span class="html-italic">T. chinensis</span> (RDOt), relative dominance of community <span class="html-italic">P. australis</span> (RDOph), and relative dominance of community <span class="html-italic">S. Salsa</span> + community <span class="html-italic">L. bicolor</span> (RDOsl); (<b>e</b>) The proportional abundance of community <span class="html-italic">S. Salsa</span> (Ps), proportional abundance of community <span class="html-italic">T. chinensis</span> (Pt), proportional abundance of community <span class="html-italic">P. australis</span> (Pph), proportional abundance of community <span class="html-italic">S. Salsa</span> + community <span class="html-italic">L. bicolor</span> Psl), Shannon-Wiener diversity index (H), Evenness index (E), an ecological variable (<math display="inline"><semantics> <mrow> <mi mathvariant="normal">X</mi> </mrow> </semantics></math>), and mean shape index (MSI).</p>
Full article ">Figure 6
<p>The compositions and ring structures of individual vegetation patches in Subarea A. (<b>a</b>) The UAV multispectral orthomosaic RGB true color composite imagery; (<b>b</b>) the vegetation composition.</p>
Full article ">Figure A1
<p>The vegetation composition of Subarea B.</p>
Full article ">Figure A2
<p>The vegetation composition of Subarea C.</p>
Full article ">Figure A3
<p>The vegetation composition of Subarea D.</p>
Full article ">Figure A4
<p>The vegetation composition of Subarea E.</p>
Full article ">Figure A5
<p>The vegetation composition of Subarea F.</p>
Full article ">Figure A6
<p>The vegetation composition of Subarea G.</p>
Full article ">Figure A7
<p>The vegetation composition of Subarea H.</p>
Full article ">
23 pages, 22713 KiB  
Article
Evaluation of Ecological Environment Quality Using an Improved Remote Sensing Ecological Index Model
by Yanan Liu, Wanlin Xiang, Pingbo Hu, Peng Gao and Ai Zhang
Remote Sens. 2024, 16(18), 3485; https://doi.org/10.3390/rs16183485 - 20 Sep 2024
Viewed by 319
Abstract
The Remote Sensing Ecological Index (RSEI) model is widely used for large-scale, rapid Ecological Environment Quality (EEQ) assessment. However, both the RSEI and its improved models have limitations in explaining the EEQ with only two-dimensional (2D) factors, resulting in [...] Read more.
The Remote Sensing Ecological Index (RSEI) model is widely used for large-scale, rapid Ecological Environment Quality (EEQ) assessment. However, both the RSEI and its improved models have limitations in explaining the EEQ with only two-dimensional (2D) factors, resulting in inaccurate evaluation results. Incorporating more comprehensive, three-dimensional (3D) ecological information poses challenges for maintaining stability in large-scale monitoring, using traditional weighting methods like the Principal Component Analysis (PCA). This study introduces an Improved Remote Sensing Ecological Index (IRSEI) model that integrates 2D (normalized difference vegetation factor, normalized difference built-up and soil factor, heat factor, wetness, difference factor for air quality) and 3D (comprehensive vegetation factor) ecological factors for enhanced EEQ monitoring. The model employs a combined subjective–objective weighting approach, utilizing principal components and hierarchical analysis under minimum entropy theory. A comparative analysis of IRSEI and RSEI in Miyun, a representative study area, reveals a strong correlation and consistent monitoring trends. By incorporating air quality and 3D ecological factors, IRSEI provides a more accurate and detailed EEQ assessment, better aligning with ground truth observations from Google Earth satellite imagery. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study area with different ecological function zones. A1: core zone for comprehensive development; A2: ecological connotation zone; A3: ecological conservation zone; A4: water conservation zone; A5: green development zone.</p>
Full article ">Figure 2
<p>The pipeline of the proposed method.</p>
Full article ">Figure 3
<p>Map of the calculated ecological factors.</p>
Full article ">Figure 4
<p>The distribution of the randomly selected field measurements using <b><span class="html-italic">GE</span></b> images. The symbols A1–A5 represent the five ecological function zones outlined in <a href="#sec2dot2-remotesensing-16-03485" class="html-sec">Section 2.2</a>.</p>
Full article ">Figure 5
<p>Visual interpretation on model validation using <b><span class="html-italic">GE</span></b> images.</p>
Full article ">Figure 6
<p>The correlation coefficients between the various factors.</p>
Full article ">Figure 7
<p>The coefficients between each EEQ factor and the <b><span class="html-italic">IRSEI</span></b> model.</p>
Full article ">Figure 8
<p>Map of the <b><span class="html-italic">EEQ</span></b> results using <b><span class="html-italic">IRSEI</span></b> and <b><span class="html-italic">RSEI</span></b> models.</p>
Full article ">Figure 9
<p>Visual and statistical comparison of EEQ results.</p>
Full article ">Figure 10
<p>A map of the EEQ change transfer matrix.</p>
Full article ">Figure 11
<p><b><span class="html-italic">EEQ</span></b> results were obtained by <b><span class="html-italic">IRSEI</span></b> and <b><span class="html-italic">RSEI</span></b> in different functional zones.</p>
Full article ">Figure 12
<p>Comparison of <b><span class="html-italic">EEQ</span></b> results between the 3D factor (<b><span class="html-italic">CVI</span></b>) and 2D factor (<b><span class="html-italic">NDVI</span></b>).</p>
Full article ">Figure 13
<p>The visualization of the relationship between each model and its ecological factors.</p>
Full article ">Figure 13 Cont.
<p>The visualization of the relationship between each model and its ecological factors.</p>
Full article ">Figure 14
<p>A visual summary of ecological factor distribution.</p>
Full article ">Figure 15
<p>Comparison of <b><span class="html-italic">EEQ</span></b> results between the <b><span class="html-italic">RSEI</span></b> and <b><span class="html-italic">IRSEI</span></b> models.</p>
Full article ">
24 pages, 8315 KiB  
Article
Spatiotemporal Changes in Vegetation Cover during the Growing Season and Its Implications for Chinese Grain for Green Program in the Luo River Basin
by Xuning Qiao, Jing Zhang, Liang Liu, Jinyuan Zhang and Tongqian Zhao
Forests 2024, 15(9), 1649; https://doi.org/10.3390/f15091649 - 19 Sep 2024
Viewed by 341
Abstract
The Grain for Green Program (GFGP) plays a critical role in enhancing watershed vegetation cover. Analyzing changes in vegetation cover provides significant practical value in guiding ecological conservation and restoration in vulnerable regions. This study utilizes MOD13Q1 NDVI data to construct the Kernel [...] Read more.
The Grain for Green Program (GFGP) plays a critical role in enhancing watershed vegetation cover. Analyzing changes in vegetation cover provides significant practical value in guiding ecological conservation and restoration in vulnerable regions. This study utilizes MOD13Q1 NDVI data to construct the Kernel Normalized Difference Vegetation Index (kNDVI) and analyzes the spatiotemporal evolution and future trends of vegetation cover from 2000 to 2020, covering key periods of the GFGP. The study innovatively combines the optimal parameter geographic detector with constraint lines to comprehensively reveal the nonlinear constraints, intensities, and critical thresholds imposed by various driving factors on the kNDVI. The results indicate that the following: (1) The vegetation cover of the Luo River Basin increased significantly between 2000 and 2020, with a noticeable increase in the percentage of high-quality vegetation. Spatially, the vegetation cover followed a pattern of being “high in the southwest and low in the northeast”, with 73.69% of the region displaying improved vegetation conditions. Future vegetation degradation is predicted to threaten 59.40% of the region, showing a continuous or future declining trend. (2) The primary driving factors for changes in the vegetation cover are evapotranspiration, elevation, population density, and geomorphology type, with temperatures and GDP being secondary factors. Dual-factor enhancement or nonlinear enhancement was observed in interactions among the factors, with evapotranspiration and population density having the largest interaction (q = 0.76). (3) The effects of driving factors on vegetation exhibited various patterns, with thresholds existing for the hump-shaped and concave-waved types. The stability of the kNDVI in 40.23% of the areas showed moderate to high fluctuations, with the most significant fluctuations observed in low-altitude and high-temperature areas, as well as those impacted by dense human activities. (4) By overlaying the kNDVI classifications on the GFGP areas, priority reforestation areas totaling 68.27 km2 were identified. The findings can help decisionmakers optimize the next phase of the GFGP and in effective regional ecological management. Full article
(This article belongs to the Special Issue Application of Remote Sensing in Vegetation Dynamic and Ecology)
Show Figures

Figure 1

Figure 1
<p>The study area in the Luo River Basin.</p>
Full article ">Figure 2
<p>The flow chart of the study.</p>
Full article ">Figure 3
<p>Interannual variation; (<b>a</b>) Annual annual mean variation trend in the kNDVI; (<b>b</b>) the results of the Mann–Kendall change–point test.</p>
Full article ">Figure 4
<p>Statistics on the multiyear average kNDVI in the Luo River Basin: (<b>a</b>) average kNDVI spatial distribution; (<b>b</b>) proportion of different kNDVI levels under various geomorphological types.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>d</b>) Characteristics of the changes in the kNDVI during the critical period; (<b>e</b>) variations in the percentage of the various classes of the kNDVI from 2000 to 2020.</p>
Full article ">Figure 6
<p>Significance analysis of the kNDVI trends and future sustainability. (<b>a</b>) the trend of Sen–MK changed significantly; (<b>b</b>) Hurst index; (<b>c</b>) future change trend.</p>
Full article ">Figure 7
<p>The explanatory power of various driving factors on the spatial pattern of the kNDVI in 2020: (<b>a</b>) results of the detection of the driving factors of the kNDVI; (<b>b</b>) results of the detection of interactions among the driving factors. Note * indicates nonlinear enhancement, and other dual-factor enhancement.</p>
Full article ">Figure 8
<p>The constraint relationship between the continuous driving factor indicators and kNDVI. (<b>a</b>) Constraint line of DEM; (<b>b</b>) Constraint line of Slope; (<b>c</b>) Constraint line of PRE; (<b>d</b>) Constraint line of TEM; (<b>e</b>) Constraint line of PET; (<b>f</b>) Constraint line of POP; (<b>g</b>) Constraint line of GDP; (<b>h</b>) Constraint line of NL; (<b>i</b>) Constraint line of NHD; (<b>j</b>) Constraint line of PHD; (<b>k</b>) Constraint line of RD; (<b>l</b>) Constraint line of WD.</p>
Full article ">Figure 9
<p>Spatial distribution pattern of the CV of the annual mean kNDVI in the Luo River Basin.</p>
Full article ">Figure 10
<p>Proportion of the kNDVI stability classes in key impact factor subdivisions.</p>
Full article ">Figure 11
<p>The kNDVI identification of prioritized GFGP areas.</p>
Full article ">
20 pages, 16133 KiB  
Article
Changes in Vegetation Cover and the Relationship with Surface Temperature in the Cananéia–Iguape Coastal System, São Paulo, Brazil
by Jakeline Baratto, Paulo Miguel de Bodas Terassi and Emerson Galvani
Remote Sens. 2024, 16(18), 3460; https://doi.org/10.3390/rs16183460 - 18 Sep 2024
Viewed by 492
Abstract
The objective of this article is to investigate the possible correlations between vegetation indices and surface temperature in the Cananéia–Iguape Coastal System (CICS), in São Paulo (Brazil). Vegetation index data from MODIS orbital products were used to carry out this work. The Normalized [...] Read more.
The objective of this article is to investigate the possible correlations between vegetation indices and surface temperature in the Cananéia–Iguape Coastal System (CICS), in São Paulo (Brazil). Vegetation index data from MODIS orbital products were used to carry out this work. The Normalized Difference Vegetation Index (NDVI) and the Enhanced Vegetation Index (EVI) were acquired from the MODIS/Aqua sensor (MYD13Q1) and the leaf area index (LAI) from the MODIS/Terra (MOD15A2H). Surface temperature data were acquired from MODIS/Aqua (MYD11A2). The data were processed using Google Earth Engine and Google Colab. The data were collected, and spatial and temporal correlations were applied. Correlations were applied in the annual and seasonal period. The annual temporal correlation between vegetation indices and surface temperature was positive, but statistically significant for the LAI, with r = 0.43 (90% significance). In the seasonal period, positive correlations occurred in JFM for all indices (95% significance). Spatially, the results of this research indicate that the largest area showed a positive correlation between VI and LST. The hottest and rainiest periods (OND and JFM) had clearer and more significant correlations. In some regions, significant and clear correlations were observed, such as in some areas in the north, south and close to the city of Iguape. This highlights the complexity of the interactions between vegetation indices and climatic attributes, and highlights the importance of considering other environmental variables and processes when interpreting changes in vegetation. However, this research has significantly progressed the field, by establishing new correlations and demonstrating the importance of considering climate variability, for a more accurate understanding of the impacts on vegetation indices. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area (<b>A</b>,<b>B</b>) and land use mapping (<b>C</b>).</p>
Full article ">Figure 2
<p>Variation in surface temperature and monthly (<b>A</b>) and annual (<b>B</b>) rainfall for the Cananéia-Iguape Coastal System for the 20032022 period. Source: MODIS/Aqua and CHIRPS, 2024.</p>
Full article ">Figure 3
<p>Annual variation in vegetation indices for the 2003–2022 period in the Cananéia–Iguape Coastal System.</p>
Full article ">Figure 4
<p>Scatter plot of annual NDVI (<b>a</b>), EVI (<b>b</b>) and LAI (<b>c</b>) values and surface temperature from 2003 to 2022.</p>
Full article ">Figure 5
<p>Scatter plot of seasonal values of VI–NDVI (<b>a</b>,<b>d</b>), EVI (<b>b</b>,<b>e</b>) and LAI (<b>c</b>,<b>f</b>)—and surface temperature for the JFM (<b>a</b>–<b>c</b>) and AMJ (<b>d</b>–<b>f</b>) quarter.</p>
Full article ">Figure 6
<p>Scatter plot of seasonal values of VI–NDVI (<b>a</b>,<b>d</b>), EVI (<b>b</b>,<b>e</b>) and LAI (<b>c</b>,<b>f</b>)—and climate variables for the JAS (<b>a</b>–<b>c</b>) and OND (<b>d</b>–<b>f</b>) quarter.</p>
Full article ">Figure 7
<p>Annual linear correlation between surface temperature and NDVI (<b>A</b>), EVI (<b>B</b>) and LAI (<b>C</b>) between 2003 and 2022.</p>
Full article ">Figure 8
<p>Seasonal linear correlation between surface temperature and VI between 2004 and 2022 for the JFM (<b>A</b>–<b>C</b>) and AMJ (<b>D</b>–<b>F</b>) periods.</p>
Full article ">Figure 9
<p>Seasonal linear correlation between surface temperature and VI between 2004 and 2022 for the JAS (<b>A</b>–<b>C</b>) and OND (<b>D</b>–<b>F</b>) periods.</p>
Full article ">
21 pages, 5562 KiB  
Article
Interdecadal Variations in Agricultural Drought Monitoring Using Land Surface Temperature and Vegetation Indices: A Case of the Amahlathi Local Municipality in South Africa
by Phumelelani Mbuqwa, Hezekiel Bheki Magagula, Ahmed Mukalazi Kalumba and Gbenga Abayomi Afuye
Sustainability 2024, 16(18), 8125; https://doi.org/10.3390/su16188125 - 18 Sep 2024
Viewed by 1027
Abstract
Agricultural droughts in South Africa, particularly in the Amahlathi Local Municipality (ALM), significantly impact socioeconomic activities, sustainable livelihoods, and ecosystem services, necessitating urgent attention to improved resilience and food security. The study assessed the interdecadal drought severity and duration in Amahlathi’s agricultural potential [...] Read more.
Agricultural droughts in South Africa, particularly in the Amahlathi Local Municipality (ALM), significantly impact socioeconomic activities, sustainable livelihoods, and ecosystem services, necessitating urgent attention to improved resilience and food security. The study assessed the interdecadal drought severity and duration in Amahlathi’s agricultural potential zone from 1989 to 2019 using various vegetation indicators. Landsat time series data were used to analyse the land surface temperature (LST), soil-adjusted vegetation index (SAVI), normalized difference vegetation index (NDVI), and standardized precipitation index (SPI). The study utilised GIS-based weighted overlay, multiple linear regression models, and Pearson’s correlation analysis to assess the correlations between LST, NDVI, SAVI, and SPI in response to the agricultural drought extent. The results reveal a consistent negative correlation between LST and NDVI in the ALM, with an increase in vegetation (R2 = 0.9889) and surface temperature. LST accuracy in dry areas increased to 55.8% in 2019, despite dense vegetation and a high average temperature of 40.12 °C, impacting water availability, agricultural land, and local ecosystems. The regression analysis shows a consistent negative correlation between LST and NDVI in the ALM from 1989 to 2019, with the correlation between vegetation and surface temperature increasing since 2019. The SAVI indicates a slight improvement in overall average vegetation health from 0.18 in 1989 to 0.25 in 2009, but a slight decrease to 0.21 in 2019. The SPI at 12 and 24 months indicates that drought severely impacted vegetation cover from 2014 to 2019, with notable recovery during improved wet periods in 1993, 2000, 2003, 2006, 2008, and 2013, possibly due to temporary drought relief. The findings can guide provincial drought monitoring and early warning programs, enhancing drought resilience, productivity, and sustainable livelihoods, especially in farming communities. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Map of the Amahlathi Local Municipality and (<b>b</b>) the distribution of land cover in agricultural potential zones in the Eastern Cape Province, South Africa.</p>
Full article ">Figure 2
<p>Spatiotemporal analysis of NDVI from 1989 to 2019.</p>
Full article ">Figure 3
<p>Spatiotemporal analysis of SAVI from 1989 to 2019.</p>
Full article ">Figure 4
<p>Spatiotemporal analysis of LST from 1989 to 2019.</p>
Full article ">Figure 5
<p>Drought classification for the ALM from 1989 to 2019.</p>
Full article ">Figure 6
<p>Correlation analysis between LST and NDVI from 1989 to 2019 in the ALM.</p>
Full article ">Figure 7
<p>Correlation analysis between LST and SAVI from 1989 to 2019 in the ALM.</p>
Full article ">Figure 8
<p>SPI drought patterns for 12 and 24 months in the ALM (1989–2019).</p>
Full article ">
19 pages, 9732 KiB  
Article
Improved Methods for Retrieval of Chlorophyll Fluorescence from Satellite Observation in the Far-Red Band Using Singular Value Decomposition Algorithm
by Kewei Zhu, Mingmin Zou, Shuli Sheng, Xuwen Wang, Tianqi Liu, Yongping Cheng and Hui Wang
Remote Sens. 2024, 16(18), 3441; https://doi.org/10.3390/rs16183441 - 17 Sep 2024
Viewed by 497
Abstract
Solar-induced chlorophyll fluorescence (SIF) is highly correlated with photosynthesis and can be used for estimating gross primary productivity (GPP) and monitoring vegetation stress. The far-red band of the solar Fraunhofer lines (FLs) is close to the strongest SIF emission peak and is unaffected [...] Read more.
Solar-induced chlorophyll fluorescence (SIF) is highly correlated with photosynthesis and can be used for estimating gross primary productivity (GPP) and monitoring vegetation stress. The far-red band of the solar Fraunhofer lines (FLs) is close to the strongest SIF emission peak and is unaffected by chlorophyll absorption, making it suitable for SIF intensity retrieval. In this study, we propose a retrieval window for far-red SIF, significantly enhancing the sensitivity of data-driven methods to SIF signals near 757 nm. This window introduces a weak O2 absorption band based on the FLs window, allowing for better separation of SIF signals from satellite spectra by altering the shape of specific singular vectors. Additionally, a frequency shift correction algorithm based on standard non-shifted reference spectra is proposed to discuss and eliminate the influence of the Doppler effect. SIF intensity retrieval was achieved using data from the GOSAT satellite, and the retrieved SIF was validated using GPP, enhanced vegetation index (EVI) from the MODIS platform, and published GOSAT SIF products. The validation results indicate that the SIF products provided in this study exhibit higher fitting goodness with GPP and EVI at high spatiotemporal resolutions, with improvements ranging from 55% to 129%. At low spatiotemporal resolutions, the SIF product provided in this study shows higher consistency with EVI and GPP spatially. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>Singular vectors in the forward model and the state vector of Fs within two retrieval windows: (<b>a</b>) FLs band, (<b>b</b>) joint retrieval for FLs-O<sub>2</sub> absorption bands.</p>
Full article ">Figure 2
<p>The identical set of spectra before and after wavenumber correction. For the sake of clarity, only a limited portion of the FTS-Band1 spectrum is displayed.</p>
Full article ">Figure 3
<p>Scatter plot and goodness-of-fit (R<sup>2</sup>), Pearson correlation coefficient (P) of monthly SIF products with GPP and VI for January 2019 at 0.1° spatial resolution.</p>
Full article ">Figure 4
<p>Goodness-of-fit (GOF) of the two SIF products with GPP and EVI at 0.1° spatial resolution and monthly scale from January 2018 to June 2020. The shaded part corresponds to December 2018 when SIF retrieves too few results.</p>
Full article ">Figure 5
<p>Pearson correlation coefficients (<span class="html-italic">p</span>-values) of the two SIF products with GPP and EVI at 0.1° spatial resolution and monthly scale from January 2018 to June 2020. The shaded part corresponds to December 2018 when SIF retrieves too few results.</p>
Full article ">Figure 6
<p>Scatter plot and goodness-of-fit (R), Pearson correlation coefficient (P) of 2019 annual SIF products with GPP and VI at 0.1° spatial resolution.</p>
Full article ">Figure 7
<p>Scatter plot and goodness-of-fit (R), Pearson correlation coefficient (P) of 2019 annual SIF products with GPP and VI at 2° spatial resolution.</p>
Full article ">Figure 8
<p>Intensity distribution of the two 2019 annual mean SIF products and 2019 annual mean EVI, GPP in 0.1° grid units. To facilitate observation, morphological dilation was applied to the SIF intensity distribution images.</p>
Full article ">Figure 9
<p>Intensity distribution of the two 2019 annual solar-normalized SIF products and annual mean EVI, GPP in 2° grid units.</p>
Full article ">Figure 10
<p>Maps of the mean 2019 annual intensity distribution for the two SIF products at 2° grid cells and of the mean annual results for EVI at 0.1° grid cells. The color–intensity relationship is the same for each row of subplots. The maps contain four regions: Northern South America, the United States and southern Canada, Western Europe, and southern Africa.</p>
Full article ">Figure 11
<p>Slope and GOF of the linear fit between the SIF retrieval results from the combined retrieval window (O<sub>2</sub> absorption and FLs bands) and FLs band alone with GPP/EVI.</p>
Full article ">Figure 12
<p>The first six singular vectors and the seventh to eighth singular vectors obtained from the spectra of the training set before and after frequency shift correction.</p>
Full article ">Figure 13
<p>Scatter plots and linear fitting results of retrieval outcomes with GPP and EVI before and after satellite spectral frequency shift correction in January 2019.</p>
Full article ">Figure 14
<p>The intensity distribution of the existing daily average SIF (<b>a</b>) and the proposed daily average SIF (<b>b</b>) is presented for the entire year of 2019 under 10.5 Km × 10.5 Km spatial resolution. Additionally, the intensity distribution of monthly MODIS Enhanced Vegetation Index (EVI) (<b>c</b>) under 0.5° grid cells and annual GPP (<b>d</b>) under 500 m SIN grid is displayed for the entire year of 2019.</p>
Full article ">
18 pages, 4238 KiB  
Article
Combining Vegetation Indices to Identify the Maize Phenological Information Based on the Shape Model
by Huizhu Wu, Bing Liu, Bingxue Zhu, Zhijun Zhen, Kaishan Song and Jingquan Ren
Agriculture 2024, 14(9), 1608; https://doi.org/10.3390/agriculture14091608 - 14 Sep 2024
Viewed by 333
Abstract
Maize is the world’s largest food crop and plays a critical role in global food security. Accurate phenology information is essential for improving yield estimation and enabling timely field management. Yet, much of the research has concentrated on general crop growth periods rather [...] Read more.
Maize is the world’s largest food crop and plays a critical role in global food security. Accurate phenology information is essential for improving yield estimation and enabling timely field management. Yet, much of the research has concentrated on general crop growth periods rather than on pinpointing key phenological stages. This gap in understanding presents a challenge in determining how different vegetation indices (VIs) might accurately extract phenological information across these stages. To address this, we employed the shape model fitting (SMF) method to assess whether a multi-index approach could enhance the precision of identifying key phenological stages. By analyzing time-series data from various VIs, we identified five phenological stages (emergence, seven-leaf, jointing, flowering, and maturity stages) in maize cultivated in Jilin Province. The findings revealed that each VI had distinct advantages depending on the phenological stage, with the land surface water index (LSWI) being particularly effective for jointing and flowering stages due to its correlation with vegetation water content, achieving a root mean square error (RMSE) of three to four days. In contrast, the normalized difference vegetation index (NDVI) was more effective for identifying the emergence and seven-leaf stages, with an RMSE of four days. Overall, combining multiple VIs significantly improved the accuracy of phenological stage identification. This approach offers a novel perspective for utilizing diverse VIs in crop phenology, thereby enhancing the precision of agricultural monitoring and management practices. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the study area and sites.</p>
Full article ">Figure 2
<p>Division of vegetative and reproductive growth stages in maize phenology.</p>
Full article ">Figure 3
<p>Flowchart of phenological period identification.</p>
Full article ">Figure 4
<p>Relationship between field-observed and estimated phenological stages obtained from different indicators and the RMSE. (<b>a</b>,<b>c</b>,<b>e</b>) The data from 2003 to 2014 and (<b>b</b>,<b>d</b>,<b>f</b>) the data from 2015 to 2019. Both the x and y axes represent the DOY.</p>
Full article ">Figure 5
<p>Box plots of the errors in identifying phenological phases in different regions: (<b>a</b>) emergence, (<b>b</b>) seven-leaf stage, (<b>c</b>) jointing, (<b>d</b>) flowering, and (<b>e</b>) maturity. The errors are represented by residuals, with positive values indicating delayed predictions and negative values indicating earlier predictions.</p>
Full article ">Figure 6
<p>The target curve fitting reference curve in Baicheng City, Jilin Province, in 2019, where the orange curve is the reference curve, the blue curve is the target curve, and the green curve is the reference curve after deformation by the fitting function. In the subplots, (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>) represent the fitting cases where NDVI is used as the reference curve for the shape model, (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>) represent the cases with NDPI as the reference curve, and (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>) represent the cases with LSWI as the reference curve, with each subplot labeled accordingly.</p>
Full article ">Figure 7
<p>Spatial distribution of the five phenological stages of maize obtained from the LSWI as a reference curve.</p>
Full article ">Figure 8
<p>Reference curves of different VIs in Jilin City.</p>
Full article ">Figure 9
<p>Reflectance values and reference curves at different phenological stages of different stations. (<b>a</b>) represents the case where the phenological stages of the stations correspond to NDVI values falling on the reference curve, (<b>b</b>) represents the case for NDPI values, and (<b>c</b>) represents the case for LSWI values.</p>
Full article ">
25 pages, 5793 KiB  
Article
Prolonged Post-Harvest Preservation in Lettuce (Lactuca sativa L.) by Reducing Water Loss Rate and Chlorophyll Degradation Regulated through Lighting Direction-Induced Morphophysiological Improvements
by Jingli Yang, Jinnan Song, Jie Liu, Xinxiu Dong, Haijun Zhang and Byoung Ryong Jeong
Plants 2024, 13(18), 2564; https://doi.org/10.3390/plants13182564 - 12 Sep 2024
Viewed by 542
Abstract
To investigate the relationship between the lighting direction-induced morphophysiological traits and post-harvest storage of lettuce, the effects of different lighting directions (top, T; top + side, TS; top + bottom, TB; side + bottom, SB; and top + side + bottom, TSB; the [...] Read more.
To investigate the relationship between the lighting direction-induced morphophysiological traits and post-harvest storage of lettuce, the effects of different lighting directions (top, T; top + side, TS; top + bottom, TB; side + bottom, SB; and top + side + bottom, TSB; the light from different directions for a sum of light intensity of 600 μmol·m−2·s−1 photosynthetic photon flux density (PPFD)) on the growth morphology, root development, leaf thickness, stomatal density, chlorophyll concentration, photosynthesis, and chlorophyll fluorescence, as well as the content of nutrition such as carbohydrates and soluble proteins in lettuce were analyzed. Subsequently, the changes in water loss rate, membrane permeability (measured as relative conductivity and malondialdehyde (MDA) content), brittleness (assessed by both brittleness index and β-galactosidase (β-GAL) activity), and yellowing degree (evaluated based on chlorophyll content, and activities of chlorophyllase (CLH) and pheophytinase (PPH)) were investigated during the storage after harvest. The findings indicate that the TS treatment can effectively reduce shoot height, increase crown width, enhance leaves’ length, width, number, and thickness, and improve chlorophyll fluorescence characteristics, photosynthetic capacity, and nutrient content in lettuce before harvest. Specifically, lettuce’s leaf thickness and stomatal density showed a significant increase. Reasonable regulation of water loss in post-harvested lettuce is essential for delaying chlorophyll degradation. It was utilized to mitigate the increase in conductivity and hinder the accumulation of MDA in lettuce. The softening speed of leafy vegetables was delayed by effectively regulating the activity of the β-GAL. Chlorophyll degradation was alleviated by affecting CLH and PPH activities. This provides a theoretical basis for investigating the relationship between creating a favorable light environment and enhancing the post-harvest preservation of leafy vegetables, thus prolonging their post-harvest storage period through optimization of their morphophysiological phenotypes. Full article
(This article belongs to the Section Crop Physiology and Crop Production)
Show Figures

Figure 1

Figure 1
<p>The morphology of lettuce (<span class="html-italic">Lactuca sativa</span> L.) ‘Caesar Green’ plants under various lighting directions for 45 days. The bar indicates 5 cm.</p>
Full article ">Figure 2
<p>The thickness of lettuce leaves under various lighting directions during 45 days of cultivation. (<b>A</b>) The micrograph of the cross-section of the middle section of the mature fourth leaf from the top; the bars indicate 0.2 mm. (<b>B</b>) The leaf thickness at the same site under different treatments. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 3
<p>The growth and development of lettuce roots under various lighting directions during 45 days of cultivation. (<b>A</b>) The raw state of the lettuce root after removing the pot. (<b>B</b>) The state of the cleaned lettuce roots; the bar indicates 5 cm. (<b>C</b>) Root length (average most extended root length). (<b>D</b>) Root fresh weight. (<b>E</b>) Root dry weight. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 4
<p>The trait of lettuce leaf epidermal cells under various lighting directions during 45 days of cultivation. Micrographs of epidermal cells (20×). (<b>A</b>) The micrograph of the upper and lower epidermal cells in the mature fourth leaf from the top; the bars indicate 10 μm. (<b>B</b>) The average length and width of upper and lower epidermal cells. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 5
<p>The trait of stomatal cells in the lower surface of lettuce leaves under various lighting directions during 45 days of cultivation. Micrographs of stomatal density and morphology (20×). (<b>A</b>) The micrograph of stomatal cells in the lower surface of the mature fourth leaf from the top; the bars indicate 10 μm. (<b>B</b>) The stomatal density per 100 mm<sup>2</sup>. (<b>C</b>) The length and width of guard cell pairs. (<b>D</b>) The length and width of stomatal pores. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 6
<p>The photosynthetic and chlorophyll fluorescence characteristics of lettuce plants under various lighting directions during 45 days of cultivation. (<b>A</b>) Net photosynthetic rate. (<b>B</b>) Transpiration rate. (<b>C</b>) Stomatal conductance. (<b>D</b>) Intercellular CO<sub>2</sub> concentration. (<b>E</b>) The maximal PSII quantum yield. (<b>F</b>) The photochemical efficiency of PSII. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 7
<p>The content of carbohydrates and soluble proteins of lettuce plants under various lighting directions during 45 days of cultivation. (<b>A</b>) Total soluble sugar content. (<b>B</b>) Starch content. (<b>C</b>) Soluble protein content. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 8
<p>Cultured under various lighting directions for 45 days, after picking, the (<b>A</b>) chlorophyll content, (<b>B</b>) chlorophyllase (CLH), and (<b>C</b>) pheophytinase (PPH) contents in lettuce leaves during storage at room temperature (22 °C, 60% RH (Relative Humidity)) for 0, 3, 6, 9, and 12 days, respectively. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9).</p>
Full article ">Figure 9
<p>Cultured under various lighting directions for 45 days, after picking, the water loss rate in lettuce leaves during storage at room temperature (22 °C, 60% RH (Relative Humidity)) for 3, 6, 9, and 12 days, respectively. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9).</p>
Full article ">Figure 10
<p>Cultured under various lighting directions for 45 days, after picking, the (<b>A</b>) relative conductivity and (<b>B</b>) malondialdehyde (MDA) content in lettuce leaves during storage at room temperature (22 °C, 60% RH (Relative Humidity)) for 3, 6, 9, and 12 days, respectively. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9).</p>
Full article ">Figure 11
<p>Cultured under various lighting directions for 45 days, after picking, the (<b>A</b>) β-Galactosidase (β-GAL) activity and (<b>B</b>) brittleness in lettuce leaves during storage at room temperature (22 °C, 60% RH (Relative Humidity)) for 0, 3, 6, 9, and 12 days, respectively. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9).</p>
Full article ">Figure 12
<p>The experimental layout and design of the lighting direction combinations. (<b>A</b>) One of the plant culture shelves in the closed-type plant factory. The T, TS, TB, SB, and TSB refer to the top (1/1), top (1/2) + side (1/2), top (1/2) + bottom (1/2), side (1/2) + bottom (1/2), and top (1/3) + side (1/3) + bottom (1/3) lighting, respectively; please see the detailed information in <a href="#plants-13-02564-t002" class="html-table">Table 2</a>. (<b>B</b>) The experimental light treatments utilized white LEDs with a spectral distribution of ~400–750 nm, peaking at 452 nm. The light period followed a 12 h day/night cycle starting at 8:00 a.m. daily. (<b>C</b>) Shading treatments between layers treated with different light directions.</p>
Full article ">Figure 13
<p>Freehand slice cutting range (the middle section of the mature fourth leaf from the top of the treated plants).</p>
Full article ">
16 pages, 5287 KiB  
Article
Nano ZnO and Bioinoculants Mitigate Effects of Deficit Irrigation on Nutritional Quality of Green Peppers
by Bruna Lorrane Rosendo Martins, Kaikí Nogueira Ferreira, Josinaldo Lopes Araujo Rocha, Railene Hérica Carlos Rocha Araujo, Guilherme Lopes, Leônidas Canuto dos Santos, Francisco Bezerra Neto, Francisco Vaniés da Silva Sá, Toshik Iarley da Silva, Whashington Idalino da Silva, Geovani Soares de Lima, Francisco Jean da Silva Paiva and José Zilton Lopes Santos
Horticulturae 2024, 10(9), 969; https://doi.org/10.3390/horticulturae10090969 - 12 Sep 2024
Viewed by 321
Abstract
Green peppers (Capsicum annuum L.) are a fruit vegetable with great culinary versatility and present important nutritional properties for human health. Water deficit negatively affects the nutritional quality of green peppers’ fruits. This study aimed to investigate the influence of zinc oxide [...] Read more.
Green peppers (Capsicum annuum L.) are a fruit vegetable with great culinary versatility and present important nutritional properties for human health. Water deficit negatively affects the nutritional quality of green peppers’ fruits. This study aimed to investigate the influence of zinc oxide nanoparticles (ZnONPs), associated with plant growth-promoting bacteria (PGPB), on the post-harvest nutritional quality of green peppers subjected to water deficit. In an open-field experiment, two irrigation levels (50 and 100% of crop evapotranspiration (Etc)), four treatments composed of a combination of ZnONPs, zinc sulfate (ZnSO4), and PGPB (T1 = ZnSO4 via leaves, T2 = ZnONPs via leaves, T3 = ZnONPs via leaves + PGPB via soil, T4 = ZnSO4 via soil + PGPB via soil), and a control treatment (Control) were tested. Water deficit or water deficit mitigation treatments did not interfere with the physical–chemical parameters (except vitamin C content) and physical color parameters (except the lightness) of green peppers. On average, the water deficit reduced the levels of Ca (−13.2%), Mg (−8.5%), P (−8.5%), K (−8.6%), Mn (−10.5%), Fe (−12.2%), B (−12.0%), and Zn (−11.5%) in the fruits. Under the water deficit condition, ZnONPs or ZnSO4 via foliar, associated or not with PGPB, increased the levels of Ca (+57% in the T2 and +69.0% in the T2), P, Mg, and Fe in the fruits. At 50% Etc, the foliar application of ZnONPs in association with PGPB increases vitamin C and mineral nutrients’ contents and nutritional quality index (+12.0%) of green peppers. Applying Zn via foliar as ZnONPs or ZnSO4 mitigated the negative effects of water deficit on the quality of pepper fruits that were enhanced by the Bacillus subtilis and B. amyloliquefaciens inoculation. The ZnONPs source was more efficient than the ZnSO4 source. The water deficit alleviating effect of both zinc sources was enhanced by the PGPB. Full article
(This article belongs to the Special Issue Advances in Sustainable Cultivation of Horticultural Crops)
Show Figures

Figure 1

Figure 1
<p>Climatological data on maximum (T max) and minimum (T min) air, maximum (RHmax) and minimum (RHmin) air relative humidity, and rainfall during the experimental period in the experimental area.</p>
Full article ">Figure 2
<p>Ascorbic acid content (<b>a</b>) and luminosity (*L) of color (<b>b</b>) in green pepper fruits, as a function of irrigation depths and water deficit alleviating treatments. Control—no product application, T1 = ZnSO<sub>4</sub> via leaves, T2 = ZnONPs via leaves, T3 = ZnONP via leaves + PGPB via soil, T4 = ZnSO<sub>4</sub> via soil + PGPB via soil. Bars with the same lowercase letters do not differ for water stress, and bars with the same uppercase letters do not differ for DHA treatments by Tukey’s test (<span class="html-italic">p</span> ≤ 0.05). Vertical bars represent the standard error.</p>
Full article ">Figure 3
<p>Contents of calcium (<b>a</b>), potassium (<b>b</b>), magnesium (<b>c</b>), phosphorus (<b>d</b>), boron (<b>e</b>), iron (<b>f</b>), manganese (<b>g</b>), and zinc (<b>h</b>) in green pepper fruits, as a function of irrigation depths and treatments to alleviate water deficit. Control—no product application, T1 = ZnSO<sub>4</sub> via leaves, T2 = ZnONPs via leaves, T3 = ZnONPs via leaves + PGPB via soil, T4 = ZnSO<sub>4</sub> via soil + PGPB via soil. Bars with the same lowercase letters do not differ for water stress, and bars with the same uppercase letters do not differ for DHA treatments by Tukey’s test (<span class="html-italic">p</span> ≤ 0.05). Vertical bars represent the standard error.</p>
Full article ">Figure 4
<p>General fruit quality index (FQI) of green pepper, as a function of irrigation depths and water deficit alleviating treatments. Control—no product application, T1 = ZnSO<sub>4</sub> via leaves, T2 = ZnONPs via leaves, T3 = ZnONPs via leaves + PGPB via soil, T4 = ZnSO<sub>4</sub> via soil + PGPB via soil. Columns with the same lowercase letters do not differ for water stress, and bars with the same uppercase letters do not differ for combinations involving zinc oxide nanoparticles (ZnONPs) or bioinoculants (PGPB) by Tukey’s test (<span class="html-italic">p</span> ≤ 0.05). Vertical bars represent the standard error.</p>
Full article ">Figure A1
<p>Experimental design used in the research showing the distribution of blocks, combination of irrigation levels with treatments from subplots, delimitation of experimental plots and useful plot, as well as spacing between blocks, plots, planting rows, and plants.</p>
Full article ">Figure A2
<p>Partial view of the experimental area showing the installation of the drip irrigation system (<b>a</b>) and the plants in the fruit production phase (<b>b</b>).</p>
Full article ">Figure A3
<p>Harvest (<b>a</b>), sanitization process and classification (<b>b</b>) of green pepper fruits before carrying out physical–chemical analyses.</p>
Full article ">
29 pages, 6780 KiB  
Article
Phenological and Biophysical Mediterranean Orchard Assessment Using Ground-Based Methods and Sentinel 2 Data
by Pierre Rouault, Dominique Courault, Guillaume Pouget, Fabrice Flamain, Papa-Khaly Diop, Véronique Desfonds, Claude Doussan, André Chanzy, Marta Debolini, Matthew McCabe and Raul Lopez-Lozano
Remote Sens. 2024, 16(18), 3393; https://doi.org/10.3390/rs16183393 - 12 Sep 2024
Viewed by 579
Abstract
A range of remote sensing platforms provide high spatial and temporal resolution insights which are useful for monitoring vegetation growth. Very few studies have focused on fruit orchards, largely due to the inherent complexity of their structure. Fruit trees are mixed with inter-rows [...] Read more.
A range of remote sensing platforms provide high spatial and temporal resolution insights which are useful for monitoring vegetation growth. Very few studies have focused on fruit orchards, largely due to the inherent complexity of their structure. Fruit trees are mixed with inter-rows that can be grassed or non-grassed, and there are no standard protocols for ground measurements suitable for the range of crops. The assessment of biophysical variables (BVs) for fruit orchards from optical satellites remains a significant challenge. The objectives of this study are as follows: (1) to address the challenges of extracting and better interpreting biophysical variables from optical data by proposing new ground measurements protocols tailored to various orchards with differing inter-row management practices, (2) to quantify the impact of the inter-row at the Sentinel pixel scale, and (3) to evaluate the potential of Sentinel 2 data on BVs for orchard development monitoring and the detection of key phenological stages, such as the flowering and fruit set stages. Several orchards in two pedo-climatic zones in southeast France were monitored for three years: four apricot and nectarine orchards under different management systems and nine cherry orchards with differing tree densities and inter-row surfaces. We provide the first comparison of three established ground-based methods of assessing BVs in orchards: (1) hemispherical photographs, (2) a ceptometer, and (3) the Viticanopy smartphone app. The major phenological stages, from budburst to fruit growth, were also determined by in situ annotations on the same fields monitored using Viticanopy. In parallel, Sentinel 2 images from the two study sites were processed using a Biophysical Variable Neural Network (BVNET) model to extract the main BVs, including the leaf area index (LAI), fraction of absorbed photosynthetically active radiation (FAPAR), and fraction of green vegetation cover (FCOVER). The temporal dynamics of the normalised FAPAR were analysed, enabling the detection of the fruit set stage. A new aggregative model was applied to data from hemispherical photographs taken under trees and within inter-rows, enabling us to quantify the impact of the inter-row at the Sentinel 2 pixel scale. The resulting value compared to BVs computed from Sentinel 2 gave statistically significant correlations (0.57 for FCOVER and 0.45 for FAPAR, with respective RMSE values of 0.12 and 0.11). Viticanopy appears promising for assessing the PAI (plant area index) and FCOVER for orchards with grassed inter-rows, showing significant correlations with the Sentinel 2 LAI (R2 of 0.72, RMSE 0.41) and FCOVER (R2 0.66 and RMSE 0.08). Overall, our results suggest that Sentinel 2 imagery can support orchard monitoring via indicators of development and inter-row management, offering data that are useful to quantify production and enhance resource management. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>Schematic of the three approaches used to monitor orchard development at different spatial scales throughout the year (from tree level for phenological observations to watershed level using Sentinel 2 data).</p>
Full article ">Figure 2
<p>(<b>a</b>) Locations of the monitored orchards in the Ouvèze–Ventoux watershed (green points at right) and in the La Crau area (yellow points at left), (<b>b</b>) pictures of 2 cherry orchards (13 September and 22 July 2022): top, non-grassed orchard drip-irrigated by two rows of drippers and bottom, grassed orchard drip-irrigated in summer, (<b>c</b>) pictures of 2 orchards in La Crau (top, nectarine tree in spring 22 March 2023 and bottom, in summer 26 June 2022).</p>
Full article ">Figure 3
<p>(<b>a</b>) Main steps in processing the hemispherical photographs. (<b>b</b>) The three methods of data acquisition around the central tree. (<b>c</b>) Protocol used with hemispherical photographs. (<b>d</b>) Protocol used with the Viticanopy application, with 3 trees monitored in the four directions (blue arrows). (<b>e</b>) Protocols used with the ceptometer: P1 measured in the shadow of the trees and (blue) P2 in the inter-rows (black).</p>
Full article ">Figure 4
<p>Protocol for the monitoring of the phenological stages of cherry trees. (<b>a</b>) Phenology of cherry trees according to BBCH; (<b>b</b>) at plot scale, in an orchard, three trees in red monitored by observations (BBCH scale); (<b>c</b>) at tree scale, two locations are selected to classify flowering stage in the tree; and (<b>d</b>) flowering stage of a cherry tree in April 2022.</p>
Full article ">Figure 5
<p>Comparison of temporal profiles of Sentinel 2 LAI interpolated profile (black line) and PAI obtained from the ceptometer (blue line, P2 protocol) and Viticanopy (green line) for three orchards: (<b>a</b>) 3099 (cherry—grassed—Ouvèze), (<b>b</b>) 183 (cherry—non-grassed—Ouvèze), and (<b>c</b>) 4 (nectarine—La Crau) at the beginning of 2023.</p>
Full article ">Figure 6
<p>Comparison between Sentinel 2 LAI and PAI from (<b>a</b>) ceptometer measurements taken at all orchards of the two areas (La Crau and Ouvèze), (<b>b</b>) Viticanopy measurements at all orchards, and (<b>c</b>) Viticanopy measurements excluding 2 non-grassed orchards (183, 259). The black line represents the optimal correlation 1:1; the red line represents the results from linear regression.</p>
Full article ">Figure 7
<p>(<b>a</b>)—(top graphs) Proportion of tree (orange <span class="html-italic">100*FCOVER<sub>t</sub>/FCOVER<sub>c</sub></span>, see Equation (1)) and of inter-row (green <span class="html-italic">100*((1-FCOVER<sub>t</sub>)*FCOVER<sub>g</sub>)/FCOVER<sub>c</sub></span>) components computed from hemispherical photographs used to estimate FCOVER for two dates, 22 March 2022 (doy:81) and 21 June 2022 (doy 172), for all the monitored fields. (<b>b</b>)—(bottom graphs) For two plots, left, field 183.2 and right, field 3099.1, temporal variations in proportion of tree and inter-row components for the different observation dates in 2022.</p>
Full article ">Figure 8
<p>(<b>a</b>) Averaged percentage of grass contribution on FAPAR computed from hemispherical photographs according to Equation (1) for all grassed orchard plots in 2022. Examples of Sentinel 2 FAPAR dynamics (black lines) for plots at (<b>b</b>) non-grassed site 183 and (<b>c</b>) grassed site 1418. Initial values of FAPAR, as computed from BVNET, are provided in black. The green line represents adjusted FAPAR after subtracting the grass contribution (percentage obtained from hemispherical photographs). It corresponds to FAPAR only for the trees. The percentage of grass contribution is in red.</p>
Full article ">Figure 9
<p>Correlation between (<b>a</b>) FCOVER obtained from hemispherical photographs (from Equation (1)) for all orchards of the two studied areas and FCOVER from Sentinel 2 computed with BVNET (<b>b</b>) FAPAR from hemispherical photographs and FAPAR from Sentinel 2 for all orchards and for the 3 years. (<b>c</b>) Correlation between FCOVER from Viticanopy and Sentinel 2 for all orchards for the two areas, except 183 and 259. (<b>d</b>) Correlation between FCOVER from upward-aimed hemispherical photographs and from Viticanopy for all plots.</p>
Full article ">Figure 10
<p>(<b>a</b>) LAI temporal profiles obtained from BVNET applied to Sentinel 2 data averaged at plot and field scales (field 3099) for the year 2022 and (<b>b</b>) soil water stock (in mm in blue) computed at 0–50 cm using capacitive sensors (described in <a href="#sec2dot1-remotesensing-16-03393" class="html-sec">Section 2.1</a>), with rainfall recorded at the Carpentras station (see <a href="#app1-remotesensing-16-03393" class="html-app">Supplementary Part S1 and Table S1</a>).</p>
Full article ">Figure 11
<p>Time series of FCOVER (mean value at field scale) for the cherry trees in field 3099 in Ouvèze area from 2016 to 2023.</p>
Full article ">Figure 12
<p>Sentinel 2 FAPAR evolution in 2022 for two cherry tree fields, with the date of flowering observation (in green) and the date of fruit set observation (in red) for (<b>a</b>) plot 183 (non-grassed cherry trees) and (<b>b</b>) plot 3099 (grassed cherry trees).</p>
Full article ">Figure 13
<p>Variability in dates for the phenological stages of a cherry tree orchard (plot 3099) observed in 2022.</p>
Full article ">Figure 14
<p>(<b>a</b>) Normalised FAPAR computed for all observed cherry trees relative to observation dates for BBCH stages in the Ouvèze area in 2021 for five plots. (<b>b</b>) Map of dates distinguishing between flowering and fruit set stages for 2021 obtained by thresholding FAPAR images.</p>
Full article ">
Back to TopTop