[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

Search Results (243)

Search Parameters:
Keywords = high spatiotemporal retrieval

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 32897 KiB  
Article
On the Suitability of Different Satellite Land Surface Temperature Products to Study Surface Urban Heat Islands
by Alexandra Hurduc, Sofia L. Ermida and Carlos C. DaCamara
Remote Sens. 2024, 16(20), 3765; https://doi.org/10.3390/rs16203765 - 10 Oct 2024
Viewed by 546
Abstract
Remote sensing satellite data have been a crucial tool in understanding urban climates. The variety of sensors with different spatiotemporal characteristics and retrieval methodologies gave rise to a multitude of approaches when analyzing the surface urban heat island effect (SUHI). Although there are [...] Read more.
Remote sensing satellite data have been a crucial tool in understanding urban climates. The variety of sensors with different spatiotemporal characteristics and retrieval methodologies gave rise to a multitude of approaches when analyzing the surface urban heat island effect (SUHI). Although there are considerable advantages that arise from these different characteristics (spatiotemporal resolution, time of observation, etc.), it also means that there is a need for understanding the ability of sensors in capturing spatial and temporal SUHI patterns. For this, several land surface temperature products are compared for the cities of Madrid and Paris, retrieved from five sensors: the Spinning Enhanced Visible and InfraRed Imager onboard Meteosat Second Generation, the Advanced Very-High-Resolution Radiometer onboard Metop, the Moderate-resolution Imaging Spectroradiometer onboard both Aqua and Terra, and the Thermal Infrared Sensor onboard Landsat 8 and 9. These products span a wide range of LST algorithms, including split-window, single-channel, and temperature–emissivity separation methods. Results show that the diurnal amplitude of SUHI may not be well represented when considering daytime and nighttime polar orbiting platforms. Also, significant differences arise in SUHI intensity and spatial and temporal variability due to the different methods implemented for LST retrieval. Full article
(This article belongs to the Section AI Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Land cover resampled for the three projections of LST products (<b>a</b>–<b>f</b>) along with the percentage of urban pixels (<b>g</b>–<b>l</b>).</p>
Full article ">Figure 2
<p>Time of observation of each sensor: (<b>a</b>) for Madrid during daytime and time of minimum SUHI (SUHI<sub>min</sub>), (<b>b</b>) for Madrid during nighttime and time of maximum SUHI (SUHI<sub>max</sub>), (<b>c</b>) for Paris during daytime and SUHI<sub>max</sub>, (<b>d</b>) for Paris during nighttime and SUHI<sub>min</sub>. Colored bins are sampled every 15 min.</p>
Full article ">Figure 3
<p>Mean DJF (December, January, February) LST for all products considered. (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST maximum and SUHI minimum are shown; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST minimum and SUHI maximum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 4
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for JJA (June, July, and August). (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST maximum and SUHI minimum are shown; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST minimum and SUHI maximum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 5
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for Paris. (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST and SUHI maximum; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST and SUHI minimum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 6
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for Paris and DJF; (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST and SUHI maximum; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST and SUHI minimum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime, an extension of the histogram in (<b>d6</b>) is seen in (<b>d8</b>). Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 7
<p>Diurnal cycle of SUHI for Madrid: (<b>a</b>) DJF, (<b>b</b>) MAM, (<b>c</b>) JJA, (<b>d</b>) SON.</p>
Full article ">Figure 8
<p>As <a href="#remotesensing-16-03765-f007" class="html-fig">Figure 7</a> but for Paris.</p>
Full article ">Figure 9
<p>Correlation of monthly SUHI anomalies between all products considered: (<b>a</b>) daytime, (<b>b</b>) nighttime. Blank spaces correspond to pairs of products with no significant correlation (<span class="html-italic">p</span>-value &gt; 0.05).</p>
Full article ">Figure 10
<p>As <a href="#remotesensing-16-03765-f009" class="html-fig">Figure 9</a> but for Paris. (<b>a</b>) daytime, (<b>b</b>) nighttime. Blank spaces correspond to pairs of products with no significant correlation (<span class="html-italic">p</span>-value &gt; 0.05).</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 609
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 ">
17 pages, 6013 KiB  
Article
Remote Sensing Monitoring and Multidimensional Impact Factor Analysis of Urban Heat Island Effect in Zhengzhou City
by Xiangjun Zhang, Guoqing Li, Haikun Yu, Guangxu Gao and Zhengfang Lou
Atmosphere 2024, 15(9), 1097; https://doi.org/10.3390/atmos15091097 - 9 Sep 2024
Viewed by 509
Abstract
In the 21st century, the rapid urbanization process has led to increasingly severe urban heat island effects and other urban thermal environment issues, posing significant challenges to urban planning and environmental management. This study focuses on Zhengzhou, China, utilizing Landsat remote sensing imagery [...] Read more.
In the 21st century, the rapid urbanization process has led to increasingly severe urban heat island effects and other urban thermal environment issues, posing significant challenges to urban planning and environmental management. This study focuses on Zhengzhou, China, utilizing Landsat remote sensing imagery data from five key years between 2000 and 2020. By applying atmospheric correction methods, we accurately retrieved the land surface temperature (LST). The study employed a gravity center migration model to track the spatial changes of heat island patches and used the geographical detector method to quantitatively analyze the combined impact of surface characteristics, meteorological conditions, and socio-economic factors on the urban heat island effect. Results show that the LST in Zhengzhou exhibits a fluctuating growth trend, closely related to the expansion of built-up areas and urban planning. High-temperature zones are mainly concentrated in built-up areas, while low-temperature zones are primarily found in areas covered by water bodies and vegetation. Notably, the Normalized Difference Built-up Index (NDBI) and the Normalized Difference Vegetation Index (NDVI) are the two most significant factors influencing the spatial distribution of land surface temperature, with explanatory power reaching 42.7% and 41.3%, respectively. As urban development enters a stable stage, government environmental management measures have played a positive role in mitigating the urban heat island effect. This study not only provides a scientific basis for understanding the spatiotemporal changes in land surface temperature in Zhengzhou but also offers new technical support for urban planning and management, helping to alleviate the urban heat island effect and improve the living environment quality for urban residents. Full article
(This article belongs to the Special Issue UHI Analysis and Evaluation with Remote Sensing Data)
Show Figures

Figure 1

Figure 1
<p>Study area. Sub-figure (<b>a</b>) shows the geographical distribution of the study area and (<b>b</b>) shows the mean annual temperature and solar radiation profile of the study area.</p>
Full article ">Figure 2
<p>Land surface temperature maps and time series of LST and solar radiation (2000–2020).</p>
Full article ">Figure 2 Cont.
<p>Land surface temperature maps and time series of LST and solar radiation (2000–2020).</p>
Full article ">Figure 3
<p>Thermal landscape classification maps and time series of LST and solar radiation (2000–2020).</p>
Full article ">Figure 3 Cont.
<p>Thermal landscape classification maps and time series of LST and solar radiation (2000–2020).</p>
Full article ">Figure 4
<p>Maps of heat island level areas by district.</p>
Full article ">Figure 4 Cont.
<p>Maps of heat island level areas by district.</p>
Full article ">Figure 5
<p>Spatial distribution maps of the center of gravity of heat island patches and built-up land.</p>
Full article ">
18 pages, 9719 KiB  
Article
Detection and Retrieval of Supercooled Water in Stratocumulus Clouds over Northeastern China Using Millimeter-Wave Radar and Microwave Radiometer
by Hao Hu, Yan Yin, Jing Yang, Xinghua Bao, Bo Zhang and Wei Gao
Remote Sens. 2024, 16(17), 3232; https://doi.org/10.3390/rs16173232 - 31 Aug 2024
Viewed by 588
Abstract
Supercooled water in mixed-phase clouds plays a significant role in precipitation formation, atmospheric radiation, weather modification, and aircraft flight safety. Identifying supercooled water in mixed-phase clouds is a crucial-frontier scientific issue in atmospheric detection research. In this study, we propose a new algorithm [...] Read more.
Supercooled water in mixed-phase clouds plays a significant role in precipitation formation, atmospheric radiation, weather modification, and aircraft flight safety. Identifying supercooled water in mixed-phase clouds is a crucial-frontier scientific issue in atmospheric detection research. In this study, we propose a new algorithm for identifying supercooled water based on the multi-spectral peak characteristics in cloud radar power spectra, combined with radar reflectivity factor and mean Doppler velocity. Using microwave radiometer data, we conducted retrieval analyses on two stratocumulus cases in the spring over the northeastern Daxing’anling region, China. The retrieval results show that the supercooled water in the spring stratocumulus clouds over the region is widespread, with liquid water content (LWC) ranging around 0.1 ± 0.05 g/m3, and particle sizes not exceeding 10 μm. The influence of updrafts on supercooled water is evident, with both showing good consistency in spatiotemporal variation trends. Comparing the liquid water path (LWP) variations retrieved from cloud radar and microwave radiometer, both showed good consistency in variation trends and high LWC areas, indicating the reliability of the identification algorithm developed in this study. Full article
Show Figures

Figure 1

Figure 1
<p>Topographic distribution of northeastern China and the location of the observation station (marked by the red triangle).</p>
Full article ">Figure 2
<p>Flow chart of cloud supercooled water recognition algorithm based on Ka-MMCR data.</p>
Full article ">Figure 3
<p>Radar power spectra detected at 4.2 km by the Tulihe radar at 13:45 on 18 May 2023. (<b>a</b>,<b>b</b>) are the original and corrected spectra, respectively.</p>
Full article ">Figure 4
<p>Radar power spectra detected at 3.75 km and 3.85 km by the Tulihe radar at 13:10 on 18 May 2023 (Beijing time, the same below). (<b>a</b>,<b>b</b>) represent the cases of multimodal peaks and multiple modes, respectively.</p>
Full article ">Figure 5
<p>Distribution of <span class="html-italic">Z</span><sub>e</sub> (<b>a</b>) and <span class="html-italic">V</span><sub>D</sub> (<b>b</b>) for 7075 separated single-peak spectra within 2 km above the freezing layer height; when the velocity is greater than 0, the direction points upwards.</p>
Full article ">Figure 6
<p>Radar detection and retrieval results of stratocumulus clouds in Tulihe on 18 May 2023, from 13:00 to 14:00. (<b>a</b>–<b>h</b>) represent the reflectivity factor <span class="html-italic">Z</span><sub>e</sub> (dBZ), <span class="html-italic">σ</span><sub>v</sub> (m/s), mean Doppler velocity <span class="html-italic">V</span><sub>D</sub> (m/s), vertical air velocity <span class="html-italic">V</span><sub>air</sub> (m/s), particle mean fall velocity <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> (m/s), supercooled water identification, liquid water content LWC (g/m<sup>3</sup>), and effective radius of supercooled cloud droplets <span class="html-italic">R</span><sub>e</sub> (μm), respectively; when the velocity is greater than 0, the direction points upwards.</p>
Full article ">Figure 6 Cont.
<p>Radar detection and retrieval results of stratocumulus clouds in Tulihe on 18 May 2023, from 13:00 to 14:00. (<b>a</b>–<b>h</b>) represent the reflectivity factor <span class="html-italic">Z</span><sub>e</sub> (dBZ), <span class="html-italic">σ</span><sub>v</sub> (m/s), mean Doppler velocity <span class="html-italic">V</span><sub>D</sub> (m/s), vertical air velocity <span class="html-italic">V</span><sub>air</sub> (m/s), particle mean fall velocity <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> (m/s), supercooled water identification, liquid water content LWC (g/m<sup>3</sup>), and effective radius of supercooled cloud droplets <span class="html-italic">R</span><sub>e</sub> (μm), respectively; when the velocity is greater than 0, the direction points upwards.</p>
Full article ">Figure 7
<p>Radar detection and retrieval results of stratocumulus clouds in Tulihe on 30 May 2023, from 17:00 to 17:45. (<b>a</b>–<b>h</b>) represent the reflectivity factor <span class="html-italic">Z</span><sub>e</sub> (dBZ), <span class="html-italic">σ</span><sub>v</sub> (m/s), mean Doppler velocity <span class="html-italic">V</span><sub>D</sub> (m/s), vertical air velocity <span class="html-italic">V</span><sub>air</sub> (m/s), particle mean fall velocity <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> (m/s), supercooled water identification, liquid water content LWC (g/m<sup>3</sup>), and effective radius of supercooled cloud droplets <span class="html-italic">R</span><sub>e</sub> (μm), respectively.</p>
Full article ">Figure 7 Cont.
<p>Radar detection and retrieval results of stratocumulus clouds in Tulihe on 30 May 2023, from 17:00 to 17:45. (<b>a</b>–<b>h</b>) represent the reflectivity factor <span class="html-italic">Z</span><sub>e</sub> (dBZ), <span class="html-italic">σ</span><sub>v</sub> (m/s), mean Doppler velocity <span class="html-italic">V</span><sub>D</sub> (m/s), vertical air velocity <span class="html-italic">V</span><sub>air</sub> (m/s), particle mean fall velocity <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> (m/s), supercooled water identification, liquid water content LWC (g/m<sup>3</sup>), and effective radius of supercooled cloud droplets <span class="html-italic">R</span><sub>e</sub> (μm), respectively.</p>
Full article ">Figure 8
<p>Radar detection and inversion results of stratocumulus clouds in Tulihe on 30 May 2023, from 20:20 to 21:00. (<b>a</b>–<b>h</b>) are reflectivity factor <span class="html-italic">Z</span><sub>e</sub> (dBZ), spectral width <span class="html-italic">σ</span><sub>v</sub> (m/s), mean Doppler velocity <span class="html-italic">V</span><sub>D</sub> (m/s), vertical air velocity <span class="html-italic">V</span><sub>air</sub> (m/s), average particle fall velocity <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math> (m/s), supercooled water identification mark, liquid water content LWC (g/m<sup>3</sup>), and effective radius of supercooled cloud droplets <span class="html-italic">R</span><sub>e</sub> (μm), respectively.</p>
Full article ">Figure 9
<p>Water vapor density at different altitudes in three cases (g/m<sup>3</sup>): (<b>a</b>) 18 May 2023, 13:00–14:00; (<b>b</b>) 30 May 2023, 17:00–18:00; (<b>c</b>) 30 May 2023, 20:00–21:00.</p>
Full article ">Figure 10
<p>Trends in vertical air velocity (<span class="html-italic">V</span><sub>air</sub>) and liquid water content (LWC) (<b>a1</b>–<b>a3</b>) and the effective radius of supercooled water droplets (<span class="html-italic">R</span><sub>e</sub>) (<b>b1</b>–<b>b3</b>) within the supercooled water regions for the two cases. Panels (<b>a1</b>,<b>b1</b>) depict the temporal variations of these three variables at 3.5 km for Case 1, panels (<b>a2</b>,<b>b2</b>) show the temporal variations of these three variables at 4.05 km for Case 2, while panels (<b>a3</b>,<b>b3</b>) show the temporal variations in these three variables at 4.5 km for Case 3.</p>
Full article ">Figure 11
<p>Comparison of liquid water path (LWP, g/m<sup>2</sup>) retrieved by radar (MMCR) and microwave radiometer (MWR). (<b>a</b>–<b>c</b>) represent the periods from 13:00 to 14:00 on 18 May, from 17:00 to 17:45 on 30 May, and from 20:15 to 21:00 on 30 May, respectively.</p>
Full article ">
24 pages, 7359 KiB  
Article
Vegetation Water Content Retrieval from Spaceborne GNSS-R and Multi-Source Remote Sensing Data Using Ensemble Machine Learning Methods
by Yongfeng Zhang, Jinwei Bu, Xiaoqing Zuo, Kegen Yu, Qiulan Wang and Weimin Huang
Remote Sens. 2024, 16(15), 2793; https://doi.org/10.3390/rs16152793 - 30 Jul 2024
Viewed by 834
Abstract
Vegetation water content (VWC) is a crucial parameter for evaluating vegetation growth, climate change, natural disasters such as forest fires, and drought prediction. Spaceborne global navigation satellite system reflectometry (GNSS-R) has become a valuable tool for soil moisture (SM) and biomass remote sensing [...] Read more.
Vegetation water content (VWC) is a crucial parameter for evaluating vegetation growth, climate change, natural disasters such as forest fires, and drought prediction. Spaceborne global navigation satellite system reflectometry (GNSS-R) has become a valuable tool for soil moisture (SM) and biomass remote sensing (RS) due to its higher spatial resolution compared with microwave measurements. Although previous studies have confirmed the enormous potential of spaceborne GNSS-R for vegetation monitoring, the utilization of this technology to fuse multiple RS parameters to retrieve VWC is not yet mature. For this purpose, this paper constructs a local high-spatiotemporal-resolution spaceborne GNSS-R VWC retrieval model that integrates key information, such as bistatic radar cross section (BRCS), effective scattering area, CYGNSS variables, and surface auxiliary parameters based on five ensemble machine learning (ML) algorithms (i.e., bagging tree (BT), gradient boosting decision tree (GBDT), extreme gradient boosting (XGBoost), random forest (RF), and light gradient boosting machine (LightGBM)). We extensively tested the performance of different models using SMAP ancillary data as validation data, and the results show that the root mean square errors (RMSEs) of the BT, XGBoost, RF, and LightGBM models in VWC retrieval are better than 0.50 kg/m2. Among them, the BT and RF models performed the best in localized VWC retrieval, with RMSE values of 0.50 kg/m2. Conversely, the XGBoost model exhibits the worst performance, with an RMSE of 0.85 kg/m2. In terms of RMSE, the RF model demonstrates improvements of 70.00%, 52.00%, and 32.00% over the XGBoost, LightGBM, and GBDT models, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>IGBP land classification map (2021 year).</p>
Full article ">Figure 2
<p>Algorithm flow of bagging tree model algorithm.</p>
Full article ">Figure 3
<p>Construction and evaluation flow chart of VWC retrieval mode.</p>
Full article ">Figure 4
<p>Scatter density plots for the retrieval of VWC and SMAP VWC using five models: (<b>a</b>) GBDT; (<b>b</b>) BT; (<b>c</b>) XGBoost; (<b>d</b>) LightGBM; (<b>e</b>) RF.</p>
Full article ">Figure 5
<p>SMAP VWC (<b>a</b>) and the distribution of local bias (Australia) between SMAP VWC and the estimated VWC of five models: (<b>b</b>) BT; (<b>c</b>) GBDT; (<b>d</b>) XGBoost; (<b>e</b>) LightGBM; (<b>f</b>) RF.</p>
Full article ">Figure 6
<p>Histograms illustrating the distribution of discrepancies between SMAP VWC and VWC estimated by five models: (<b>a</b>) BT; (<b>b</b>) GBDT; (<b>c</b>) XGBoost; (<b>d</b>) LightGBM; (<b>e</b>) RF.</p>
Full article ">Figure 7
<p>The importance of 16 indices generated by five models: (<b>a</b>) BT; (<b>b</b>) GBDT; (<b>c</b>) XGBoost; (<b>d</b>) LightGBM; (<b>e</b>) RF.</p>
Full article ">Figure 8
<p>Performance evaluation of different parameter combination strategies (three schemes) on five models (GBDT, BT, XGBoost, LightGBM, and RF). (<b>a</b>) RMSE; (<b>b</b>) MAE; (<b>c</b>) MAPE; (<b>d</b>) R.</p>
Full article ">Figure 9
<p>Scatter density plots of VWC and SMAP VWC for four quarters retrieved by five models: (<b>a</b>–<b>e</b>) spring; (<b>f</b>–<b>j</b>) summer; (<b>k</b>–<b>o</b>) autumn; (<b>p</b>–<b>t</b>) winter.</p>
Full article ">Figure 10
<p>VWC retrieval performance of various models at different latitudes: (<b>a</b>–<b>e</b>) low latitudes; (<b>f</b>–<b>j</b>) mid-latitudes; (<b>k</b>–<b>o</b>) high latitudes.</p>
Full article ">Figure 11
<p>PDF distribution curves of VWC and SMAP VWC retrieved from five models with different vegetation cover: low (<b>a</b>), medium (<b>b</b>), and high (<b>c</b>).</p>
Full article ">
21 pages, 6948 KiB  
Article
Has IMERG_V07 Improved the Precision of Precipitation Retrieval in Mainland China Compared to IMERG_V06?
by Hao Guo, Yunfei Tian, Junli Li, Chunrui Guo, Xiangchen Meng, Wei Wang and Philippe De Maeyer
Remote Sens. 2024, 16(14), 2671; https://doi.org/10.3390/rs16142671 - 22 Jul 2024
Viewed by 562
Abstract
Integrated Multi-satellitE Retrievals for GPM (Global Precipitation Measurement) (IMERG) is the primary high spatiotemporal resolution precipitation product of the GPM era. To assess the applicability of the latest released IMERG_V07 in mainland China, this study systematically evaluates the error characteristics of IMERG_V07 from [...] Read more.
Integrated Multi-satellitE Retrievals for GPM (Global Precipitation Measurement) (IMERG) is the primary high spatiotemporal resolution precipitation product of the GPM era. To assess the applicability of the latest released IMERG_V07 in mainland China, this study systematically evaluates the error characteristics of IMERG_V07 from the perspective of different seasons, precipitation intensity, topography, and climate regions on an hourly scale. Ground-based meteorological observations are used as the reference, and the performance improvement of IMERG_V07 relative to IMERG_V06 is verified. Error evaluation is conducted in terms of precipitation amount and precipitation frequency, and an improved error component procedure is utilized to trace the error sources. The results indicate that IMERG_V07 exhibits a smaller RMSE in mainland China, especially with significant improvements in the southeastern region. IMERG_V07 shows better consistency with ground station data. IMERG_V07 shows an overall improvement of approximately 4% in capturing regional average precipitation events compared to IMERG_V06, with the northwest region showing particularly notable enhancement. The error components of IMERG_V06 and IMERG_V07 exhibit similar spatial distributions. IMERG_V07 outperforms V06 in terms of lower Missed bias but slightly underperforms in Hit bias and False bias compared to IMERG_V06. IMERG_V07 shows improved ability in capturing precipitation frequency for different intensities, but challenges remain in capturing heavy precipitation events, missing light precipitation, and winter precipitation events. Both IMERG_V06 and IMERG_V07 exhibit notable topography dependency in terms of Total bias and error components. False bias is the primary error source for both versions, except in winter, where high-altitude regions (DEM > 1200 m) primarily contribute to Missed bias. IMERG_V07 has enhanced the accuracy of precipitation retrieval in high-altitude areas, but there are still limitations in capturing precipitation events. Compared to IMERG_V06, IMERG_V07 demonstrates more concentrated error component values in the four climatic regions, with reduced data dispersion and significant improvement in Missed bias. The algorithm improvements in IMERG_V07 have the most significant impact in arid regions. False bias serves as the primary error source for both satellite-based precipitation estimations in the four climatic regions, with a secondary contribution from Hit bias. The evaluation results of this study offer scientific references for enhancing the algorithm of IMERG products and enhancing users’ understanding of error characteristics and sources in IMERG. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Digital elevation model (DEM) and (<b>b</b>) locations of meteorological stations in mainland China.</p>
Full article ">Figure 2
<p>Spatial distribution of precipitation for meteorological stations (<b>a</b>–<b>d</b>), IMERG_V06 (<b>e</b>–<b>h</b>), and IMERG_V07 (<b>i</b>–<b>l</b>) over four seasons (spring, summer, autumn, and winter).</p>
Full article ">Figure 3
<p>Spatial distribution of Bias (<b>a</b>,<b>e</b>), CC (<b>b</b>,<b>f</b>), RB (<b>c</b>,<b>g</b>), and RMSE (<b>d</b>,<b>h</b>) between hourly precipitation data from SPEs and observations. Regional averaged values for BIAS (mm), CC, RB (%), and RMSE (mm/h) are shown in the color-coded barplots. The <span class="html-italic">x</span>-axis numbers 1–4 correspond to humid, semi-humid, semi-arid, and arid regions, respectively.</p>
Full article ">Figure 4
<p>Taylor diagrams showing CC, STD, and RMSE of hourly mean precipitation between SPE and observations in different seasons: (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; and (<b>d</b>) winter.</p>
Full article ">Figure 5
<p>Hourly-scale spatial distribution of categorical statistical indexes (POD, MIS, FAR) for IMERG_V06 (<b>a</b>–<b>c</b>) and IMERG_V07 (<b>d</b>–<b>f</b>) with a 0.1 mm/hour precipitation/no precipitation threshold. The barplot with different colors indicates the regional averaged values for POD, MIS, and FAR with a unit of %. The <span class="html-italic">x</span>-axis numbers 1–4 correspond to humid, semi-humid, semi-arid, and arid regions, respectively.</p>
Full article ">Figure 6
<p>Spatial distribution of error components at hourly scales for IMERG_V06 (<b>a</b>–<b>d</b>) and IMERG_V07 (<b>e</b>–<b>h</b>). The inserted barplot with different colors indicates the regional averaged values for Total bias and different error components with a unit of mm/h. The <span class="html-italic">x</span>-axis numbers 1–4 correspond to humid, semi-humid, semi-arid, and arid regions, respectively.</p>
Full article ">Figure 7
<p>Hourly precipitation classification metrics for IMERG SPEs at different intensities in spring (<b>a</b>,<b>e</b>), summer (<b>b</b>,<b>f</b>), autumn (<b>c</b>,<b>g</b>) and winter (<b>d</b>,<b>h</b>) in mainland China. Note that different <span class="html-italic">Y</span>-axis limits are used in <a href="#remotesensing-16-02671-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 8
<p>Categorical statistical indices (<b>a</b>–<b>d</b>) and RMSE (<b>c</b>–<b>h</b>) as a function of elevation for IMERG SPEs in spring, summer, autumn, and winter.</p>
Full article ">Figure 9
<p>Schematic representation of the total bias and error components as a function of altitude for IMERG SPEs in spring (<b>a</b>,<b>b</b>), summer (<b>c</b>,<b>d</b>), autumn (<b>e</b>,<b>f</b>), and winter (<b>g</b>,<b>h</b>).</p>
Full article ">Figure 10
<p>Plots of Total bias and error component half violins for IMERG SPEs in the humid (<b>a</b>–<b>d</b>), semi-humid (<b>e</b>–<b>h</b>), semi-arid (<b>i</b>–<b>l</b>), and arid (<b>m</b>–<b>p</b>) regions.</p>
Full article ">
16 pages, 9526 KiB  
Article
High-Resolution Characterization of Aerosol Optical Depth and Its Correlation with Meteorological Factors in Afghanistan
by Sayed Esmatullah Torabi, Muhammad Amin, Worradorn Phairuang, Hyung-Min Lee, Mitsuhiko Hata and Masami Furuuchi
Atmosphere 2024, 15(7), 849; https://doi.org/10.3390/atmos15070849 - 19 Jul 2024
Viewed by 1232
Abstract
Atmospheric aerosols pose a significant global problem, particularly in urban areas in developing countries where the rapid urbanization and industrial activities degrade air quality. This study examined the spatiotemporal variations and trends in aerosol optical depth (AOD) at a 550 nm wavelength, alongside [...] Read more.
Atmospheric aerosols pose a significant global problem, particularly in urban areas in developing countries where the rapid urbanization and industrial activities degrade air quality. This study examined the spatiotemporal variations and trends in aerosol optical depth (AOD) at a 550 nm wavelength, alongside key meteorological factors, in Kabul, Afghanistan, from 2000 to 2022. Using the Google Earth Engine geospatial analysis platform, daily AOD data were retrieved from the Moderate Resolution Imaging Spectroradiometer to assess monthly, seasonal, and annual spatiotemporal variations and long-term trends. Meteorological parameters such as temperature (T), relative humidity (RH), precipitation (PCP), wind speed (WS), wind direction, and solar radiation (SR) were obtained from the Modern Era Retrospective Analysis for Research and Applications. The Mann–Kendall test was employed to analyze the time-series trends, and a Pearson correlation matrix was calculated to assess the influence of the meteorological factors on AOD. Principal component analysis (PCA) was performed to understand the underlying structure. The results indicated high AOD levels in spring and summer, with a significant upward trend from 2000 to 2022. The findings revealed a positive correlation of AOD value with T, RH, WS, and PCP and a negative correlation with SR. The PCA results highlighted complex interactions among these factors and their impact on the AOD. These insights underscore the need for stringent air quality regulations and emission control measures in Kabul. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of this study’s process.</p>
Full article ">Figure 2
<p>Location and base map of the study site. The blue lines represent the boundary of Kabul, including 22 districts.</p>
Full article ">Figure 3
<p>Spatial distribution of monthly average AOD from 2000 to 2022 over Kabul. The blue and red colors indicate clear and polluted environments, respectively.</p>
Full article ">Figure 4
<p>Seasonal spatial variations in AOD over the study site in different periods: 2000, 2005, 2010, 2015, and 2020.</p>
Full article ">Figure 5
<p>Frequency distribution based on daily AOD values at 550 nm across four seasons in Kabul from 2000 to 2022.</p>
Full article ">Figure 6
<p>Annual spatial distribution of AOD from 2000 to 2022 over the study site.</p>
Full article ">Figure 7
<p>Difference in average AOD between 2000 and 2022.</p>
Full article ">Figure 8
<p>Average and standard deviation of AOD from 2000 to 2022. The red line indicates the trend.</p>
Full article ">Figure 9
<p>(<b>a</b>) Correlation matrices among AOD, T, RH, WS, PCP, and SR, calculated using Pearson correlation based on the monthly mean of these variables from 2000 to 2022. Red and blue scatterplots indicate positive and negative correlations, respectively. (<b>b</b>) The RMSE values for AOD are based on meteorological factors.</p>
Full article ">Figure 10
<p>Scatter plot of principal components (PC1 and PC2) colored by AOD value.</p>
Full article ">
20 pages, 10820 KiB  
Article
Mapping Crop Evapotranspiration by Combining the Unmixing and Weight Image Fusion Methods
by Xiaochun Zhang, Hongsi Gao, Liangsheng Shi, Xiaolong Hu, Liao Zhong and Jiang Bian
Remote Sens. 2024, 16(13), 2414; https://doi.org/10.3390/rs16132414 - 1 Jul 2024
Viewed by 599
Abstract
The demand for freshwater is increasing with population growth and rapid socio-economic development. It is more and more important for refined irrigation water management to conduct research on crop evapotranspiration (ET) data with a high spatiotemporal resolution in agricultural regions. We propose the [...] Read more.
The demand for freshwater is increasing with population growth and rapid socio-economic development. It is more and more important for refined irrigation water management to conduct research on crop evapotranspiration (ET) data with a high spatiotemporal resolution in agricultural regions. We propose the unmixing–weight ET image fusion model (UWET), which integrates the advantages of the unmixing method in spatial downscaling and the weight-based method in temporal prediction to produce daily ET maps with a high spatial resolution. The Landsat-ET and MODIS-ET datasets for the UWET fusion data are retrieved from Landsat and MODIS images based on the surface energy balance model. The UWET model considers the effects of crop phenology, precipitation, and land cover in the process of the ET image fusion. The precision evaluation is conducted on the UWET results, and the measured ET values are monitored by eddy covariance at the Luancheng station, with average MAE values of 0.57 mm/day. The image results of UWET show fine spatial details and capture the dynamic ET changes. The seasonal ET values of winter wheat from the ET map mainly range from 350 to 660 mm in 2019–2020 and from 300 to 620 mm in 2020–2021. The average seasonal ET in 2019–2020 is 499.89 mm, and in 2020–2021, it is 459.44 mm. The performance of UWET is compared with two other fusion models: the Spatial and Temporal Adaptive Reflectance Fusion Model (STARFM) and the Spatial and Temporal Reflectance Unmixing Model (STRUM). UWET performs better in the spatial details than the STARFM and is better in the temporal characteristics than the STRUM. The results indicate that UWET is suitable for generating ET products with a high spatial–temporal resolution in agricultural regions. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area and ground test station.</p>
Full article ">Figure 2
<p>The flowchart for the extraction of land types.</p>
Full article ">Figure 3
<p>The spatial distribution of sampling points.</p>
Full article ">Figure 4
<p>The NDVI curve of winter wheat and corresponding feature points from 2019 to 2020.</p>
Full article ">Figure 5
<p>Land cover map. (<b>a</b>) 2019–2020; (<b>b</b>) 2020–2021.</p>
Full article ">Figure 6
<p>The UWET framework.</p>
Full article ">Figure 7
<p>Unmixing-based spatial downscaling of Landsat-ET and MODIS-ET coarse pixels.</p>
Full article ">Figure 8
<p>The base dates and prediction dates matching process of LS-MS ET image pairs.</p>
Full article ">Figure 9
<p>The framework of weight-based temporal prediction process.</p>
Full article ">Figure 10
<p>The variation of daily UWET.</p>
Full article ">Figure 11
<p>The comparison of measured ET, Landsat-ET, and UWET at the Luancheng station.</p>
Full article ">Figure 12
<p>The spatial pattern comparison between UWET and Landsat-ET. (<b>a</b>) Land cover map of Field 1; (<b>b</b>–<b>g</b>) Landsat-ET and UWET maps on different dates in Field 1; (<b>h</b>) Land cover map of Field 2; (<b>i</b>–<b>n</b>): Landsat-ET and UWET maps on different dates in Field 2.</p>
Full article ">Figure 13
<p>Spatial distribution of wheat ET. (<b>a</b>) Accumulated ET between 2019 and 2020; (<b>b</b>) accumulated ET between 2020 and 2021.</p>
Full article ">Figure 14
<p>Validation of crop ET. (<b>a</b>) MODIS-ET in 2019–2020; (<b>b</b>) MODIS-ET in 2020–2021; (<b>c</b>) UWET in 2019–2020; (<b>d</b>) UWET in 2020–2021.</p>
Full article ">Figure 15
<p>The spatial characteristics comparison of three fusion models on 22 May 2020. (<b>a</b>) Land cover map; (<b>b</b>) MODIS-ET map; (<b>c</b>) Landsat-ET map; (<b>d</b>) STARFM-ET map; (<b>e</b>) STRUM-ET map; (<b>f</b>) UWET map.</p>
Full article ">Figure 16
<p>Daily ET of the three models during the growing season. (<b>a</b>) 2019–2020; (<b>b</b>) 2020–2021.</p>
Full article ">
24 pages, 13599 KiB  
Article
Dual-Mode Sea Ice Extent Retrieval for the Rotating Fan Beam Scatterometer
by Liling Liu, Xiaolong Dong, Wenming Lin and Shuyan Lang
Remote Sens. 2024, 16(13), 2378; https://doi.org/10.3390/rs16132378 - 28 Jun 2024
Viewed by 517
Abstract
Scatterometers are highlighted in polar applications, such as sea ice extent retrieval. There are currently three types of spaceborne scatterometer in operation, among which the rotating pencil beam scatterometer and the rotating fan beam scatterometer have similar rotating observation geometry, but different beam [...] Read more.
Scatterometers are highlighted in polar applications, such as sea ice extent retrieval. There are currently three types of spaceborne scatterometer in operation, among which the rotating pencil beam scatterometer and the rotating fan beam scatterometer have similar rotating observation geometry, but different beam sampling. However, it is difficult to objectively evaluate the performance of the two rotating beam scatterometers using the obtained data. This is because there are significant differences in their system parameters, which in turn affects the objectivity of the evaluation. Considering the high flexibility of the rotating fan beam scatterometer, this study proposes a dual-mode sea ice extent retrieval method for the rotating fan beam scatterometer. The dual modes refer to the rotating fan beam mode (or full incidence mode) and the equivalent rotating pencil beam mode (or single incidence mode). The two modes share the same system and spatiotemporal synchronous backscatter measurements provide the possibility of objectively comparing the rotating pencil beam and rotating fan beam scatterometers. The comparison, validation, and evaluation of the dual-mode sea ice extent of China France Oceanography Satellite Scatterometer (CSCAT) were performed. The results indicate that the sea ice extent retrieval of the equivalent rotating pencil beam mode of the rotating fan beam scatterometer is realizable, and compared to the existing rotating pencil beam scatterometers (such as the OceanSat Scatterometer on ScatSat-1, OSCAT, on ScatSat-1, and the Hai Yang 2B Scatterometer, HSCAT-B), the derived sea ice extent is closer to that of Advanced Microwave Scanning Radiometer 2 (AMSR2). For the two modes of CSCAT, when compared to AMSR2, the sea ice extent of the CSCAT full incidence mode has smaller values of root mean squared error (RMSE), error-of-ice (EI), and ice edge location distance (LD) than those of the CSCAT single incidence mode. These suggest that the rotating fan beam scatterometer shows better observation abilities for sea ice extent than the rotating pencil beam scatterometers. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Scan geometry of the rotating beam scatterometer: (<b>a</b>) pencil beam; (<b>b</b>) fan beam.</p>
Full article ">Figure 2
<p>Top view of CSCAT observation geometry: (<b>a</b>) rotating fan beam mode; (<b>b</b>) equivalent rotating pencil beam mode.</p>
Full article ">Figure 3
<p>The incidence and antenna azimuth angle for different CSCAT WVC values on 1 January 2019 (Revolution 12): (<b>a</b>) incidence; (<b>b</b>) antenna azimuth angle.</p>
Full article ">Figure 4
<p>The extracted data in the fore/after plane on 1 January 2019 (Revolution 12): (<b>a</b>) probability distribution of the average incidences; (<b>b</b>) distribution of the antenna azimuth angles.</p>
Full article ">Figure 5
<p>The backscattering of sea ice and open water of the CSCAT equivalent rotating pencil beam mode data in the Arctic region on 15 March 2019: (<b>a</b>) inner WVC (40°); (<b>b</b>) outer WVC (48°).</p>
Full article ">Figure 6
<p>Distribution of slope and intercept values derived from preprocessed Arctic daily backscatter data from January to March in 2019–2022.</p>
Full article ">Figure 7
<p>Sea ice GMF for the equivalent rotating pencil beam mode of CSCAT.</p>
Full article ">Figure 8
<p>Distribution of the distance between measured backscatter and the sea ice GMF model for the inner and outer WVCs on the fore/after backscatter plane using the preprocessed CSCAT data on 15 March 2019: (<b>a</b>) Arctic Region; (<b>b</b>) Antarctic Region.</p>
Full article ">Figure 8 Cont.
<p>Distribution of the distance between measured backscatter and the sea ice GMF model for the inner and outer WVCs on the fore/after backscatter plane using the preprocessed CSCAT data on 15 March 2019: (<b>a</b>) Arctic Region; (<b>b</b>) Antarctic Region.</p>
Full article ">Figure 9
<p>The value of <span class="html-italic">μ</span> and std for the inner and outer WVCs in the Arctic and Antarctic regions using the preprocessed CSCAT data: (<b>a</b>) 2019; (<b>b</b>) 2020; (<b>c</b>) 2021; (<b>d</b>) 2022.</p>
Full article ">Figure 9 Cont.
<p>The value of <span class="html-italic">μ</span> and std for the inner and outer WVCs in the Arctic and Antarctic regions using the preprocessed CSCAT data: (<b>a</b>) 2019; (<b>b</b>) 2020; (<b>c</b>) 2021; (<b>d</b>) 2022.</p>
Full article ">Figure 10
<p>Gaussian parameters for the distance normalization.</p>
Full article ">Figure 11
<p>The probability distribution of <span class="html-italic">MLE</span><sub>ice</sub> of the preprocessed CSCAT data on 24 April 2020: (<b>a</b>) Arctic region; (<b>b</b>) Antarctic region.</p>
Full article ">Figure 11 Cont.
<p>The probability distribution of <span class="html-italic">MLE</span><sub>ice</sub> of the preprocessed CSCAT data on 24 April 2020: (<b>a</b>) Arctic region; (<b>b</b>) Antarctic region.</p>
Full article ">Figure 12
<p>The probability distribution of <span class="html-italic">MLE</span><sub>wind</sub> of the preprocessed CSCAT data on 24 April 2020: (<b>a</b>) Arctic region; (<b>b</b>) Antarctic region.</p>
Full article ">Figure 13
<p>Flowchart of the dual-mode sea ice extent retrieval for the rotating fan beam scatterometer.</p>
Full article ">Figure 14
<p>Arctic Bayesian probability images (<b>left</b>) and sea ice extent images (<b>right</b>) on 10 December 2019: (<b>a</b>) CSCAT full incidence mode; (<b>b</b>) CSCAT single incidence mode. The colorbar gives the Bayesian probability and the water/ice classification is signified with 0/1.</p>
Full article ">Figure 15
<p>Comparison of sea ice extent from CSCAT dual modes during 2019–2022: (<b>a</b>) Arctic region; (<b>b</b>) Antarctic region.</p>
Full article ">Figure 16
<p>Timeline of the satellite scatterometer data used for comparison.</p>
Full article ">Figure 17
<p>Sea ice extent comparison among CSCAT dual modes, OSCAT, HSCAT-B, and AMSR2, for three years.</p>
Full article ">Figure 18
<p>Monthly RMSE distribution of CSCAT dual modes in the Antarctic region.</p>
Full article ">Figure 19
<p>Sea ice extent comparison for the different rotating pencil beam scatterometer data.</p>
Full article ">Figure 20
<p>Spatial distributions of the overestimated (red), underestimated (light blue), and overlapping (light gray) ice pixels in the sea ice extent images of CSCAT compared to AMSR2 on 10 June 2019: (<b>a</b>) CSCAT full incidence mode; (<b>b</b>) CSCAT single incidence mode. Non-ice pixels are set to white. The black lines represent the sea ice edges of AMSR2 at 15% sea ice concentration. <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi mathvariant="normal">O</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi mathvariant="normal">U</mi> </msub> </mrow> </semantics></math> correspond to the sum of all red and light blue ice pixels.</p>
Full article ">Figure 21
<p>Comparison of EI, EO, and EU between the sea ice extent resulting from the CSCAT dual mode for the three years (2019, 2020, 2022): (<b>a</b>) Arctic region; (<b>b</b>) Antarctic region.</p>
Full article ">Figure 22
<p>Comparison of LD between the sea ice extent resulted from CSCAT dual modes for the three years (2019, 2020, 2022): (<b>a</b>) Arctic region; (<b>b</b>) Antarctic region.</p>
Full article ">
15 pages, 4607 KiB  
Article
A First Exploration of the Ts/VI “Analytical Triangle” Technique with UAV Imagery for Deriving Key Surface Energy Balance Parameters at Very High Spatial Resolution
by George P. Petropoulos, Spyridon E. Detsikas, Kleomenis Kalogeropoulos and Andrew Pavlides
Drones 2024, 8(7), 290; https://doi.org/10.3390/drones8070290 - 27 Jun 2024
Cited by 1 | Viewed by 631
Abstract
Knowledge on the spatiotemporal patterns of surface energy balance parameters is crucial for understanding climate system processes. To this end, the assimilation of Earth Observation data with land biosphere models has shown promising results, but they are still hampered by several limitations related [...] Read more.
Knowledge on the spatiotemporal patterns of surface energy balance parameters is crucial for understanding climate system processes. To this end, the assimilation of Earth Observation data with land biosphere models has shown promising results, but they are still hampered by several limitations related to the spatiotemporal resolution of EO sensors and cloud contamination. With the recent developments on Unmanned Aerial Vehicles (UAVs), there is a great opportunity to overcome these challenges and gain knowledge of surface energy balance parameters at unprecedented resolutions. The present study examines, for the first time, the ability of an inversion-modeling scheme, the so-called “analytical triangle” method, to retrieve estimates of surface energy fluxes and soil surface moisture (SSM) at high spatial resolution using UAV data. A further aim of our study was to examine the representativeness of the SSM estimates for the SM measurements taken at different depths. The selected experimental site is an agricultural site of citrus trees located near the city of Palermo on 30 July 2019. The results of comparisons showed that the sensible and latent heat fluxes from UAV were consistent with those measured from the ground, with absolute differences in comparison to ground measurements being 5.00 Wm−2 for the latent heat (LE) flux and 65.02 Wm−2 for H flux, whereas for the daytime fluxes H/Rn and LE/Rn were 0.161 and 0.012, respectively. When comparing analytical triangle SSM estimates with SM measurements made at different depths, it was found that there was a gradual increase in underestimation with increasing measurement depth. All in all, this study’s results provide a credible demonstration of the significant potential of the technique investigated herein as a cost-effective and rapid solution for estimating key parameters characterizing land surface processes. As those parameters are required by a wide range of disciplines and applications, utilization of the investigated technique in research and practical applications is expected to be seen in the future. Full article
Show Figures

Figure 1

Figure 1
<p>The geographical location of this study’s experimental site located in the vicinity of Palermo, Italy (38°4′53.4″ N, 13°25′8.2″ E). The yellow labels correspond to the specific soil moisture sampling points used in the study. The top right map provides the general location of the experimental site within a broader region, marked by a circle. The bottom right map shows the location within Italy, highlighted by a square.</p>
Full article ">Figure 2
<p>A graphical illustration of methodological approach employed in the present study.</p>
Full article ">Figure 3
<p>The final pre-processed maps of Fr (<b>a</b>) and T<sub>scaled</sub> (<b>b</b>). The histograms of both variables are included in the figure.</p>
Full article ">Figure 4
<p>Maps of SSM (<b>a</b>), LE (<b>b</b>), HA (<b>c</b>), LE/R<sub>n</sub> (<b>d</b>), and H/R<sub>n</sub> (<b>e</b>) and their histogram as derived from the analytical triangle.</p>
Full article ">Figure 5
<p>Comparison results of the SSM derived from the analytical triangle based against the in situ measurements for the different soil depths. Bias, Scatter, and RMSD values are expressed in cm<sup>3</sup> cm<sup>−3</sup>.</p>
Full article ">
22 pages, 6370 KiB  
Article
Unveiling the Accuracy of New-Generation Satellite Rainfall Estimates across Bolivia’s Complex Terrain
by Silvia Roxana Mattos Gutierrez, Ayele Almaw Fenta, Taye Minichil Meshesha and Ashebir Sewale Belay
Remote Sens. 2024, 16(12), 2211; https://doi.org/10.3390/rs16122211 - 18 Jun 2024
Viewed by 1047
Abstract
This study evaluated the accuracy of two new generation satellite rainfall estimates (SREs): Climate Hazards Group InfraRed Precipitation with Station data (CHIRPS) and Integrated Multi-satellite Retrieval for GPM (IMERG) over Bolivia’s complex terrain. These SREs were compared against rainfall data from rain gauge [...] Read more.
This study evaluated the accuracy of two new generation satellite rainfall estimates (SREs): Climate Hazards Group InfraRed Precipitation with Station data (CHIRPS) and Integrated Multi-satellite Retrieval for GPM (IMERG) over Bolivia’s complex terrain. These SREs were compared against rainfall data from rain gauge measurements on a point-to-pixel basis for the period 2002–2020. The evaluation was performed across three regions with distinct topographical settings: Altiplano (Highland), Valles (Midland), and Llanos (Lowland). IMERG exhibited better accuracy in rainfall detection than CHIRPS, with the highest rainfall detection skills observed in the Highland region. However, IMERG’s higher rainfall detection skill was countered by its higher false alarm ratio. CHIRPS provided a more accurate estimation of rainfall amounts across the three regions, exhibiting low random errors and relative biases below 10%. IMERG tended to overestimate rainfall amounts, with marked overestimation by up to 75% in the Highland region. Bias decomposition revealed that IMERG’s high false rainfall bias contributed to its marked overestimation of rainfall. We showcase the utility of long-term CHIRPS data to investigate spatio-temporal rainfall patterns and meteorological drought occurrence in Bolivia. The findings of this study offer valuable insights for choosing appropriate SREs for informed decision-making, particularly in regions of complex topography lacking reliable gauge data. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location map of Bolivia showing the rainfall gauging stations used for this study. The background provides elevation information extracted from ALOS PALSAR (Phased Array L-band Synthetic Aperture Radar on the Advanced Land Observing Satellite).</p>
Full article ">Figure 2
<p>Monthly mean gauge-based rainfall (2002–2020) averaged for the Highland, Midland, and Lowland stations. Vertical bars represent standard deviation monthly rainfall.</p>
Full article ">Figure 3
<p>Scatter plots of daily SREs versus gauge measurements for the period 2002–2020 across the Highland, Midland, and Lowland regions of Bolivia.</p>
Full article ">Figure 4
<p>Scatter plots of monthly SREs versus gauge measurements for the period 2002–2020 across the Highland, Midland, and Lowland regions of Bolivia.</p>
Full article ">Figure 5
<p>Comparison of long-term mean monthly rainfall estimates by the SREs for the period 2002–2020 across the three regions of Bolivia. Vertical bars represent standard deviation monthly rainfall.</p>
Full article ">Figure 6
<p>Quantile maps showing spatial distribution of (<b>a</b>) average annual rainfall (mm year–1), (<b>b</b>) coefficient of variation of annual rainfall, (<b>c</b>) wet season rainfall (mm season–1), (<b>d</b>) coefficient of variation of wet season rainfall, (<b>e</b>) dry season rainfall (mm season–1), and (<b>f</b>) coefficient of variation of dry season rainfall, during the period 1981–2020 based on CHIRPS data.</p>
Full article ">Figure 7
<p>Standardized annual rainfall over the three regions of Bolivia normalized with respect to the 1981–2020 average based on CHIRPS data.</p>
Full article ">Figure 8
<p>Spatial distribution of standardized precipitation index (SPI) for December–February of 1995 and 2016 (drought years) and 2018 (a normal year) based on CHIRPS data.</p>
Full article ">
22 pages, 4281 KiB  
Article
Non-Uniform Spatial Partitions and Optimized Trajectory Segments for Storage and Indexing of Massive GPS Trajectory Data
by Yuqi Yang, Xiaoqing Zuo, Kang Zhao and Yongfa Li
ISPRS Int. J. Geo-Inf. 2024, 13(6), 197; https://doi.org/10.3390/ijgi13060197 - 12 Jun 2024
Viewed by 855
Abstract
The presence of abundant spatio-temporal information based on the location of mobile objects in publicly accessible GPS mobile devices makes it crucial to collect, analyze, and mine such information. Therefore, it is necessary to index a large volume of trajectory data to facilitate [...] Read more.
The presence of abundant spatio-temporal information based on the location of mobile objects in publicly accessible GPS mobile devices makes it crucial to collect, analyze, and mine such information. Therefore, it is necessary to index a large volume of trajectory data to facilitate efficient trajectory retrieval and access. It is difficult for existing indexing methods that primarily rely on data-driven indexing structures (such as R-Tree) or space-driven indexing structures (such as Quadtree) to support efficient analysis and computation of data based on spatio-temporal range queries as a service basis, especially when applied to massive trajectory data. In this study, we propose a massive GPS data storage and indexing method based on uneven spatial segmentation and trajectory optimization segmentation. Primarily, the method divides GPS trajectories in a large spatio-temporal data space into multiple MBR sequences by greedy algorithm. Then, a hybrid indexing model for segmented trajectories is constructed to form a global spatio-temporal segmentation scheme, called HHBITS index, to achieve hierarchical organization of trajectory data. Eventually, a spatio-temporal range query processing method is proposed based on this index. This paper implements and evaluates the index in MongoDB and compares it with two other spatio-temporal composite indexes for performing spatio-temporal range queries efficiently. The experimental results show that the method in this paper has high performance in responding to spatio-temporal queries on large-scale trajectory data. Full article
Show Figures

Figure 1

Figure 1
<p>Spatio-temporal trajectories.</p>
Full article ">Figure 2
<p>Data model of trajectory segmentation.</p>
Full article ">Figure 3
<p>A noisy point in trajectory and clustering.</p>
Full article ">Figure 4
<p>Example of spatio-temporal object splitting.</p>
Full article ">Figure 5
<p>The trajectory greedy splitting process.</p>
Full article ">Figure 6
<p>The architecture of the time index.</p>
Full article ">Figure 7
<p>Hilbert curve.</p>
Full article ">Figure 8
<p>Correspondence between non-uniform spatial divisions and Hilbert curve.</p>
Full article ">Figure 9
<p>Framework of trajectory segment index for NoSQL.</p>
Full article ">Figure 10
<p>The general framework of spatio-temporal range query.</p>
Full article ">Figure 11
<p>The MBR size ratio of segmentation optimization in different segment lengths.</p>
Full article ">Figure 12
<p>The comparison between basic segmentation and optimized segmentation in spatial query with different segment lengths: (<b>a</b>) index-filtering stage; (<b>b</b>) traversing the refinement stage; (<b>c</b>) total time consumption for spatio-temporal range query.</p>
Full article ">Figure 13
<p>The comparison of query accuracy between basic segmentation and optimized segmentation with different segment lengths.</p>
Full article ">Figure 14
<p>The query time for different segment lengths.</p>
Full article ">Figure 15
<p>The comparison of the spatio-temporal query performance of methods under different spatio-temporal ranges: (<b>a</b>) time interval: 1 day; (<b>b</b>) time interval: 7 days; (<b>c</b>) time interval: 30 days; (<b>d</b>) time interval: 120 days.</p>
Full article ">
18 pages, 2812 KiB  
Article
Analysing Spatiotemporal Variability of Chlorophyll-a Concentration and Water Surface Temperature in Coastal Lagoons of the Ebro Delta (NW Mediterranean Sea, Spain)
by Lara Talavera, José Antonio Domínguez-Gómez, Nuria Navarro and Inmaculada Rodríguez-Santalla
J. Mar. Sci. Eng. 2024, 12(6), 941; https://doi.org/10.3390/jmse12060941 - 3 Jun 2024
Viewed by 639
Abstract
Coastal lagoons are highly productive transitional water bodies threatened by human factors and vulnerable to global climate change effects. Monitoring biophysical parameters in these ecosystems is crucial for their preservation. In this work, we used Sentinel-2 and Landsat imagery combined with in situ [...] Read more.
Coastal lagoons are highly productive transitional water bodies threatened by human factors and vulnerable to global climate change effects. Monitoring biophysical parameters in these ecosystems is crucial for their preservation. In this work, we used Sentinel-2 and Landsat imagery combined with in situ data to (1) develop preliminary algorithms for retrieving the Chl-a concentration and water surface temperature of six lagoons located in the Ebro Delta (NE Mediterranean Sea, Spain), and to (2) compute maps and trend lines for analysing their spatiotemporal evolution from 2015 to 2022. Our findings showed that the algorithms’ accuracy ranged from 72% to 78% and had limited potential under high Chl-a concentration regimes. Even so, they revealed the lagoons’ trophic status, usual fluctuations, and deviations of both parameters attributed to seasonal (i.e., light and temperature) and short-term physical (i.e., winds) forcing, as well as valuable spatial patterns potentially useful for conservation efforts and land use planning. Future work will focus on the acquisition of a larger in situ data sample under a range of environmental conditions to improve the algorithms’ robustness, which in turn will allow the investigation of natural and human factors controlling the dynamics of the two investigated parameters. Full article
(This article belongs to the Section Marine Environmental Science)
Show Figures

Figure 1

Figure 1
<p>Location of the Ebro Delta habitats, land use types (CORINE 2018), and the two bays (1 and 6) and four coastal lagoons (2, 3, 4 and 5) of interest (Background image: Sentinel S2A-L1C image in November 2022). The in situ measurement points from the ACA (Catalan Water Agency) and ZOCOMAR research group are highlighted with red and orange triangles, respectively.</p>
Full article ">Figure 2
<p>Performance of the algorithms to extract: (<b>a</b>) Chl<span class="html-italic">-a</span> concentration from El Fangar and Els Alfacs bays (Algorithm A1), (<b>b</b>) Chl<span class="html-italic">-a</span> concentration from Bassa de L’Estella, El Garxal, Calaixos de Buda, and La Tancada coastal lagoons (Algorithm A2), and (<b>c</b>) water surface temperature in all the lagoons (Note: the displayed lines correspond to the 1:1 lines, which help to visually understand the differences between the measured and modelled data. The regression lines were also calculated and displayed R<sup>2</sup> values of 0.95, 0.96, and 0.90, for the A1, A2, and temperature algorithms, respectively).</p>
Full article ">Figure 3
<p>Mean Chl<span class="html-italic">-a</span> concentration and standard deviation for the two bays (<b>a</b>) and the four coastal lagoons (<b>b</b>) during the period of study (Note: W = Winter, Sp = Spring, Su = Summer, and A = Autumn).</p>
Full article ">Figure 4
<p>Maps displaying Chl<span class="html-italic">-a</span> concentration patterns for the bays and coastal lagoons analysed in 2021.</p>
Full article ">Figure 5
<p>Modelled water surface temperature trend lines from 2013 to 2022 in (<b>a</b>) El Fangar, (<b>b</b>) Bassa de L’Estella, (<b>c</b>) El Garxal, (<b>d</b>) Calaixos de Buda, (<b>e</b>) La Tancada, and (<b>f</b>) Els Alfacs.</p>
Full article ">Figure 6
<p>Maps displaying water surface temperature patterns for the bays and coastal lagoons analysed in 2018.</p>
Full article ">Figure 7
<p>Maps displaying water surface temperature patterns for the bays and coastal lagoons analysed in 2022.</p>
Full article ">
22 pages, 21135 KiB  
Article
Assessing the Nonlinear Relationship between Land Cover Change and PM10 Concentration Change in China
by Xiankang Xu, Jian Hao, Yuxin Liang and Jingwei Shen
Land 2024, 13(6), 766; https://doi.org/10.3390/land13060766 - 29 May 2024
Viewed by 553
Abstract
Inhalable particulate matter (PM10) is a major air pollutant that has significant impacts on environmental climate and human health. Land-cover change is also a key factor influencing changes in atmospheric pollution. Changes in land-cover types can lead to changes in the [...] Read more.
Inhalable particulate matter (PM10) is a major air pollutant that has significant impacts on environmental climate and human health. Land-cover change is also a key factor influencing changes in atmospheric pollution. Changes in land-cover types can lead to changes in the sources and sinks of air pollutants, thus affecting the spatial distribution of PM10, which poses a threat to human health. Therefore, exploring the relationship between PM10 concentration change and land-cover change is of great significance. In this study, we constructed an extreme randomized trees model (ET) based on ground PM10 monitoring data, satellite-based aerosol optical depth (AOD) data, and auxiliary data including meteorological, vegetation, and population data to retrieve ground-level PM10 concentrations across China. The coefficient of determination (R2), the mean absolute error (MAE), and the root mean square error (RMSE) of the model were 0.878, 5.742 μg/m3, and 8.826 μg/m3, respectively. Based on this, we analyzed the spatio-temporal distribution of PM10 concentrations in China from 2015 to 2021. High PM10 values were mainly observed in the desert areas of northwestern China and the Beijing–Tianjin–Hebei urban agglomeration. The majority of China showed a significant decrease in PM10 concentrations. Additionally, we also analyzed the nonlinear response mechanism of the PM10 concentration change to land-cover change. The PM10 concentration is sensitive to forest and barren land change. Therefore, strengthening the protection of forests and desertification control can significantly reduce air pollution. Attention should also be paid to emission management in agricultural activities and urbanization processes. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study area.</p>
Full article ">Figure 2
<p>The framework of the study.</p>
Full article ">Figure 3
<p>Model accuracy at each monitoring station from 2015 to 2021: (<b>a</b>) R<sup>2</sup>, (<b>b</b>) MAE, (<b>c</b>) RMSE.</p>
Full article ">Figure 4
<p>Annual and multiple-year mean PM<sub>10</sub> maps (1 km × 1 km) in China.</p>
Full article ">Figure 5
<p><span class="html-italic">p</span>-value of PM<sub>10</sub> trend over China.</p>
Full article ">Figure 6
<p>Spatial distribution of the annual PM<sub>10</sub> concentration trend in China (μg/m<sup>3</sup>/year).</p>
Full article ">Figure 7
<p>Spatial distribution of land cover in China in 2015 and 2021.</p>
Full article ">Figure 8
<p>Land-cover change spatial distribution in China.</p>
Full article ">Figure 9
<p>The proportion of PM<sub>10</sub> concentrations at different levels across land-cover types.</p>
Full article ">Figure 10
<p>Nonlinear relationship between PM<sub>10</sub> concentrations and land-cover types in 2015. (<b>a</b>) Forest; (<b>b</b>) Glassland; (<b>c</b>) Wetland and Water; (<b>d</b>) Cropland; (<b>e</b>) Urban; (<b>f</b>) Barren land.</p>
Full article ">Figure 11
<p>Nonlinear relationship between PM<sub>10</sub> concentrations and land-cover types in 2021. (<b>a</b>) Forest; (<b>b</b>) Glassland; (<b>c</b>) Wetland and Water; (<b>d</b>) Cropland; (<b>e</b>) Urban; (<b>f</b>) Barren land.</p>
Full article ">
16 pages, 7648 KiB  
Article
Comprehensive Assessment of NDVI Products Derived from Fengyun Satellites across China
by Lei Wang, Xiuzhen Han, Shibo Fang and Fengjin Xiao
Remote Sens. 2024, 16(8), 1363; https://doi.org/10.3390/rs16081363 - 12 Apr 2024
Viewed by 1008
Abstract
NDVI data are crucial for agricultural and environmental research. The Fengyun-3 (FY-3) series satellites are recognized as primary sources for retrieving NDVI products on a global scale. To apply FY-3 NDVI data for long-term studies, such as climate change, this study conducted a [...] Read more.
NDVI data are crucial for agricultural and environmental research. The Fengyun-3 (FY-3) series satellites are recognized as primary sources for retrieving NDVI products on a global scale. To apply FY-3 NDVI data for long-term studies, such as climate change, this study conducted a thorough evaluation to detect the potentials of the FY-3B and FY-3D satellites for generating a long time series NDVI dataset. For this purpose, the spatiotemporal consistency between the FY-3B and FY-3D satellites was evaluated, and their performances were compared. Then, a grey relational analysis (GRA) method was applied to detect the factors influencing the consistency among the different satellites, and a gradient boosting regression (GBR) model was constructed to create a long-term FY-3 NDVI product. The results indicate an overall high consistency between the FY-3B and FY-3D NDVIs, suggesting that they could be used as complementary datasets for generating a long-term NDVI dataset. The correlations between the FY-3D NDVI and the MODIS NDVI, as well as the leaf area index (LAI) measurements, were both higher than those of FY-3B, which indicates a better performance of FY-3D in retrieving NDVI data. The grey correlation degrees between the NDVI differences and four parameters, which were land cover (LC), DEM, latitude (LAT) and longitude (LON), were calculated, revealing that the LC was the most related to the NDVI differences. Finally, a GBR model with FY-3B NDVI, LC, DEM, LAT and LON as the input variables and FY-3D NDVI as the target variable was established and achieved a robust performance. The R values between the GBR-estimated NDVI and FY-3D NDVI reached 0.947, 0.867 and 0.829 in the training, testing and validation datasets, respectively, indicating the feasibility of the established model for generating long time series NDVI data by combining data from the FY-3B and FY-3D satellites. 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>The DEM (<b>a</b>) and land cover types (<b>b</b>) in China.</p>
Full article ">Figure 2
<p>The varying patterns of the FY-3B and FY-3D NDVI values for different land cover types from January 2020 to December 2020.</p>
Full article ">Figure 3
<p>The spatial distribution of the difference in the NDVI values between FY-3B and FY-3D: (<b>a</b>) January; (<b>b</b>) April; (<b>c</b>) July; (<b>d</b>) October.</p>
Full article ">Figure 3 Cont.
<p>The spatial distribution of the difference in the NDVI values between FY-3B and FY-3D: (<b>a</b>) January; (<b>b</b>) April; (<b>c</b>) July; (<b>d</b>) October.</p>
Full article ">Figure 4
<p>Scatterplots between the station-based LAI measurements and the NDVI values derived from (<b>a</b>) FY-3B and (<b>b</b>) FY-3D.</p>
Full article ">Figure 5
<p>The correlation coefficients and mean absolute errors between the MODIS NDVI and the NDVI derived from FY-3B and FY-3D for different land cover types.</p>
Full article ">Figure 6
<p>The grey correlation degree between the NDVI difference values and multiple factors using the GRA method.</p>
Full article ">Figure 7
<p>Features’ importance generated for the GBR model.</p>
Full article ">Figure 8
<p>The spatial patterns of the NDVI values: (<b>a</b>) FY-3D; (<b>b</b>) FY-3B; (<b>c</b>) GBR model.</p>
Full article ">
Back to TopTop