[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (365)

Search Parameters:
Keywords = SMAP

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
28 pages, 5528 KiB  
Article
Estimating Rootzone Soil Moisture by Fusing Multiple Remote Sensing Products with Machine Learning
by Shukran A. Sahaar and Jeffrey D. Niemann
Remote Sens. 2024, 16(19), 3699; https://doi.org/10.3390/rs16193699 - 4 Oct 2024
Viewed by 911
Abstract
This study explores machine learning for estimating soil moisture at multiple depths (0–5 cm, 0–10 cm, 0–20 cm, 0–50 cm, and 0–100 cm) across the coterminous United States. A framework is developed that integrates soil moisture from Soil Moisture Active Passive (SMAP), precipitation [...] Read more.
This study explores machine learning for estimating soil moisture at multiple depths (0–5 cm, 0–10 cm, 0–20 cm, 0–50 cm, and 0–100 cm) across the coterminous United States. A framework is developed that integrates soil moisture from Soil Moisture Active Passive (SMAP), precipitation from the Global Precipitation Measurement (GPM), evapotranspiration from the Ecosystem Spaceborne Thermal Radiometer Experiment on Space Station (ECOSTRESS), vegetation data from the Moderate Resolution Imaging Spectroradiometer (MODIS), soil properties from gridded National Soil Survey Geographic (gNATSGO), and land cover information from the National Land Cover Database (NLCD). Five machine learning algorithms are evaluated including the feed-forward artificial neural network, random forest, extreme gradient boosting (XGBoost), Categorical Boosting, and Light Gradient Boosting Machine. The methods are tested by comparing to in situ soil moisture observations from several national and regional networks. XGBoost exhibits the best performance for estimating soil moisture, achieving higher correlation coefficients (ranging from 0.76 at 0–5 cm depth to 0.86 at 0–100 cm depth), lower root mean squared errors (from 0.024 cm3/cm3 at 0–100 cm depth to 0.039 cm3/cm3 at 0–5 cm depth), higher Nash–Sutcliffe Efficiencies (from 0.551 at 0–5 cm depth to 0.694 at 0–100 cm depth), and higher Kling–Gupta Efficiencies (0.511 at 0–5 cm depth to 0.696 at 0–100 cm depth). Additionally, XGBoost outperforms the SMAP Level 4 product in representing the time series of soil moisture for the networks. Key factors influencing the soil moisture estimation are elevation, clay content, aridity index, and antecedent soil moisture derived from SMAP. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Locations and climates of the in situ soil moisture stations used in this study.</p>
Full article ">Figure 2
<p>Performance metrics (<span class="html-italic">R</span>, MBE, RMSE, ubRMSE, NSE, and KGE) for the soil moisture estimates of the machine learning algorithms when compared to the testing data, including all depths and stations. For each performance metric, the line inside the box indicates the median value and the box represents the interquartile range.</p>
Full article ">Figure 3
<p>RMSE of the soil moisture estimates from the machine learning algorithms for the testing dataset when the data are divided according to the (<b>a</b>) in situ soil moisture networks and (<b>b</b>) depths. For each performance metric, the line inside the box indicates the median and the box represents the interquartile range.</p>
Full article ">Figure 4
<p>Density plots comparing the observed and XGBoost estimates of soil moisture for each depth using the testing datasets for each climate. Darker blues represent higher concentrations of data, while lighter blues represent lower concentrations.</p>
Full article ">Figure 5
<p>Time series of soil moisture at (<b>a</b>) 0–5 cm and (<b>b</b>) 0–100 cm depths at the arid USCRN Las Cruces 20N station (a member of the testing dataset). The plotted soil moisture data include hourly in situ measurements, estimates from the XGBoost model, and 3 h SMAP L4 soil moisture estimates. Daily GPM precipitation data at the site are also shown.</p>
Full article ">Figure 5 Cont.
<p>Time series of soil moisture at (<b>a</b>) 0–5 cm and (<b>b</b>) 0–100 cm depths at the arid USCRN Las Cruces 20N station (a member of the testing dataset). The plotted soil moisture data include hourly in situ measurements, estimates from the XGBoost model, and 3 h SMAP L4 soil moisture estimates. Daily GPM precipitation data at the site are also shown.</p>
Full article ">Figure 6
<p>Time series of soil moisture at (<b>a</b>) 0–5 cm and (<b>b</b>) 0–100 cm depths at the humid USCRN Versailles 3NNW station (a member of the testing dataset). The plotted soil moisture data include hourly in situ measurements, estimates from the XGBoost model, and 3 h SMAP L4 soil moisture estimates. Daily GPM precipitation data for the site are also shown.</p>
Full article ">Figure 7
<p>Correlations between predictor variables and in situ soil moisture at different depths. Positive correlations are shown in blue and negative correlations are shown in red.</p>
Full article ">Figure 8
<p>Relative importance of each predictor variable in the RF, XGBoost, CatBoost, and LightGBM models and the average importance among the four models.</p>
Full article ">
27 pages, 10360 KiB  
Article
Soil Moisture-Derived SWDI at 30 m Based on Multiple Satellite Datasets for Agricultural Drought Monitoring
by Jing Ning, Yunjun Yao, Joshua B. Fisher, Yufu Li, Xiaotong Zhang, Bo Jiang, Jia Xu, Ruiyang Yu, Lu Liu, Xueyi Zhang, Zijing Xie, Jiahui Fan and Luna Zhang
Remote Sens. 2024, 16(18), 3372; https://doi.org/10.3390/rs16183372 - 11 Sep 2024
Viewed by 745
Abstract
As a major agricultural hazard, drought frequently occurs due to a reduction in precipitation resulting in a continuously propagating soil moisture (SM) deficit. Assessment of the high spatial-resolution SM-derived drought index is crucial for monitoring agricultural drought. In this study, we generated a [...] Read more.
As a major agricultural hazard, drought frequently occurs due to a reduction in precipitation resulting in a continuously propagating soil moisture (SM) deficit. Assessment of the high spatial-resolution SM-derived drought index is crucial for monitoring agricultural drought. In this study, we generated a downscaled random forest SM dataset (RF-SM) and calculated the soil water deficit index (RF-SM-SWDI) at 30 m for agricultural drought monitoring. The results showed that the RF-SM dataset exhibited better consistency with in situ SM observations in the detection of extremes than did the SM products, including SMAP, SMOS, NCA-LDAS, and ESA CCI, for different land cover types in the U.S. and yielded a satisfactory performance, with the lowest root mean square error (RMSE, below 0.055 m3/m3) and the highest coefficient of determination (R2, above 0.8) for most observation networks, based on the number of sites. A vegetation health index (VHI), derived from a Landsat 8 optical remote sensing dataset, was also generated for comparison. The results illustrated that the RF-SM-SWDI and VHI exhibited high correlations (R ≥ 0.5) at approximately 70% of the stations. Furthermore, we mapped spatiotemporal drought monitoring indices in California. The RF-SM-SWDI provided drought conditions with more detailed spatial information than did the short-term drought blend (STDB) released by the U.S. Drought Monitor, which demonstrated the expected response of seasonal drought trends, while differences from the VHI were observed mainly in forest areas. Therefore, downscaled SM and SWDI, with a spatial resolution of 30 m, are promising for monitoring agricultural field drought within different contexts, and additional reliable factors could be incorporated to better guide agricultural management practices. Full article
(This article belongs to the Special Issue Quantitative Remote Sensing of Vegetation and Its Applications)
Show Figures

Figure 1

Figure 1
<p>The spatial distribution of training and testing stations used in the downscaling framework. The map of land cover types of the substudy area and the locations of the in situ observation stations appear at the <b>top left</b> and <b>bottom</b>, respectively.</p>
Full article ">Figure 2
<p>Downscaling framework for the surface SM at 30 m through the integration of multiple datasets.</p>
Full article ">Figure 3
<p>Scatterplots of the comparison for the RF-SM data and SM derived from in situ observations at (<b>a</b>) 170 training stations and (<b>b</b>) 72 independent validation stations. The color indicates the density of the samples distributed in the area.</p>
Full article ">Figure 4
<p>Permutation importance of RF-SM. The features (i.e., input variables) include the SM products (SMAP, SMOS, ESA CCI, and NCA-LDAS), the soil properties (clay, sand, and silt), and the reflectance at visible and near-infrared bands (from SR_b4 to SR_b7), as well as the surface temperature (ST_b10) derived from Landsat 8.</p>
Full article ">Figure 5
<p>Boxplots of the in situ SM, RF-SM data, and the four SM products (SMAP, SMOS, NCA-LDAS, and KGE) for different land cover types. In the single boxplots, the red cross-dots denote outliers; the lowest and highest lines denote minimum and maximum results, respectively, except for extreme values (outliers); and the lower bound of the box, red line in the box, and upper bound of the box represent the lower quartile (25%), the median, and upper quartile (75%), respectively.</p>
Full article ">Figure 6
<p>Diagrams of the statistics (R<sup>2</sup>, RMSE, Bias, and KGE) for the comparison between the RF-SM dataset and the four SM products (SMAP, SMOS, NCA-LDAS, KGE) for the different observation networks.</p>
Full article ">Figure 7
<p>Temporal variations in precipitation (P) and surface SM derived from RF-SM and the four products at the representative stations in the substudy area during 2016.</p>
Full article ">Figure 8
<p>Spatial distributions of the RF-SM in the substudy area during 2016.</p>
Full article ">Figure 9
<p>Comparison between the RF-SM-SWDI and VHI based on the Pearson correlation coefficient (R) from 242 in situ stations in 2016.</p>
Full article ">Figure 10
<p>Temporal variations in SM-SWDI, RF-SM-SWDI, VHI, and precipitation (P) anomalies at the representative stations in the substudy area in 2016.</p>
Full article ">Figure 11
<p>Spatial distributions of the RF-SM-SWDI and VHI in the substudy area in 2016.</p>
Full article ">Figure 12
<p>Comparison between the RF-SM-SWDI and VHI based on the Pearson correlation coefficient (R) in the substudy area in 2016.</p>
Full article ">Figure 13
<p>Comparison between the RF-SM-SWDI and two VHI components: (<b>a</b>) VCI and (<b>b</b>) TCI, based on the Pearson correlation coefficient (R) in the substudy area in 2016.</p>
Full article ">Figure 14
<p>Spatial distributions of the RF-SM-SWDI, RF-SM-SWDI after resampling, and the short-term drought blend (STDB) in the substudy area in 2016.</p>
Full article ">
17 pages, 2777 KiB  
Article
Comparing Satellite Soil Moisture Products Using In Situ Observations over an Instrumented Experimental Basin in Romania
by Sofia Ortenzi, Corrado Cencetti, Florentina-Iuliana Mincu, Gianina Neculau, Viorel Chendeş, Luca Ciabatta, Christian Massari and Lucio Di Matteo
Remote Sens. 2024, 16(17), 3283; https://doi.org/10.3390/rs16173283 - 4 Sep 2024
Viewed by 699
Abstract
This study assessed the performance of different remotely sensed soil moisture products with in situ observations; six profile probes for the water content monitoring were selected, operating during 2016–2021 from the Voineşti Experimental Basin in the Romanian Subcarpathian region. The reliability of satellite [...] Read more.
This study assessed the performance of different remotely sensed soil moisture products with in situ observations; six profile probes for the water content monitoring were selected, operating during 2016–2021 from the Voineşti Experimental Basin in the Romanian Subcarpathian region. The reliability of satellite observations has been analyzed on both single ground-based observation points and spatialized information, considering near-surface and root-zone soil moisture data. The physics-based index (HCI) and some statistical tests widely used in inter-comparison analyses have been computed. The study of HCI highlighted that the SMAP SP_L4_SM products have shown the best performances considering the near-surface and root-zone data evaluations. The comparison of SWI1km observations with in situ data produced good results for single-point and spatialized soil moisture estimations acquired at different depths over the experimental basin. The SSM1km and SMAP L2_SM_SP products exhibited the lowest performances. The results contribute to the validation of satellite products of surface and root-zone soil moisture in the Subcarpathian region, helping to provide information in an area not monitored by the International Soil Moisture Network. The findings offer valuable insights into evaluating the performance of satellite soil moisture products in the Romanian region. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the Voineşti Experimental Basin (VEB, red dot) with ground-based monitoring points (blue rhombus) of the Romanian Soil Moisture Network, RSMN (EPSG:4326).</p>
Full article ">Figure 2
<p>Digital elevation model of VEB with the land use and detail of the experimental plots with the location of the pluviometer and soil moisture probes. The four grid pixels of CGLS products, including the VEB, are also depicted. The grid pixels of SMAP products are not visualized because they are much larger than the extension of VEB. Probes P7 and P8 are placed in the apple orchard, the extent of which is very small, and the symbols of probes P7 and P8 cover it.</p>
Full article ">Figure 3
<p>HCI analysis of soil moisture datasets for the 2016–2021 period over the VEB. (<b>a</b>) SSM1km; (<b>b</b>) SWI1km; (<b>c</b>) L2_SM_SP am; (<b>d</b>) L2_SM_SP apm; (<b>e</b>) SP_L4_SM ssm; (<b>f</b>) Daily ground-based rainfall data recorded at G1 rain gauge (see <a href="#remotesensing-16-03283-f002" class="html-fig">Figure 2</a> for the location). The blue and yellow bands represent the autumn-winter and spring-summer periods, respectively.</p>
Full article ">Figure 4
<p>Comparison of the different satellite products with the spatialized ground-based soil moisture values at 10 cm (<math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="false"> <mrow> <mi mathvariant="sans-serif">θ</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi mathvariant="normal">n</mi> <mo>_</mo> <mn>10</mn> </mrow> </msub> </mrow> </semantics></math>) over the VEB. (<b>a</b>) SSM1km and the related cumulative distribution function SSM1km (CDF); (<b>b</b>) SWI1km and the related cumulative distribution function SWI1km (CDF); (<b>c</b>) L2_SM_SP am and the related cumulative distribution function L2_SM_SP am (CDF); (<b>d</b>) L2_SM_SP apm and the related cumulative distribution function L2_SM_SP apm (CDF); (<b>e</b>) SP_L4_SM ssm and the related cumulative distribution function SP_L4_SM ssm (CDF).</p>
Full article ">Figure 5
<p>Comparison of in situ spatialized observations at three different depths (<math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="false"> <mrow> <mi mathvariant="sans-serif">θ</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi mathvariant="normal">n</mi> <mo>_</mo> <mn>20</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="false"> <mrow> <mi mathvariant="sans-serif">θ</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi mathvariant="normal">n</mi> <mo>_</mo> <mn>40</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="false"> <mrow> <mi mathvariant="sans-serif">θ</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi mathvariant="normal">n</mi> <mo>_</mo> <mn>60</mn> </mrow> </msub> </mrow> </semantics></math>) and for the root zone (<math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="false"> <mrow> <mi mathvariant="sans-serif">θ</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi mathvariant="normal">n</mi> <mo>_</mo> <mo>_</mo> <mi mathvariant="normal">r</mi> <mi mathvariant="normal">z</mi> </mrow> </msub> </mrow> </semantics></math>). (<b>a</b>) SWI1km T = 5 and the related cumulative distribution function SWI1km (CDF); (<b>b</b>) SWI1km T = 10 and the related cumulative distribution function SWI1km (CDF); (<b>c</b>) SWI1km T = 15 and the related cumulative distribution function SWI1km (CDF) (<b>d</b>) SP_L4_SM rz and the related cumulative distribution function SP_L4_SM rz (CDF).</p>
Full article ">
18 pages, 20239 KiB  
Article
Geoclimatic Distribution of Satellite-Observed Salinity Bias Classified by Machine Learning Approach
by Yating Ouyang, Yuhong Zhang, Ming Feng, Fabio Boschetti and Yan Du
Remote Sens. 2024, 16(16), 3084; https://doi.org/10.3390/rs16163084 - 21 Aug 2024
Viewed by 615
Abstract
Sea surface salinity (SSS) observed by satellite has been widely used since the successful launch of the first salinity satellite in 2009. However, compared with other oceanographic satellite products (e.g., sea surface temperature, SST) that became operational in the 1980s, the SSS product [...] Read more.
Sea surface salinity (SSS) observed by satellite has been widely used since the successful launch of the first salinity satellite in 2009. However, compared with other oceanographic satellite products (e.g., sea surface temperature, SST) that became operational in the 1980s, the SSS product is less mature and lacks effective validation from the user end. We employed an unsupervised machine learning approach to classify the Level 3 SSS bias from the Soil Moisture Active Passive (SMAP) satellite and its observing environment. The classification model divides the samples into fifteen classes based on four variables: satellite SSS bias, SST, rain rate, and wind speed. SST is one of the most significant factors influencing the classification. In regions with cold SST, satellite SSS has an accuracy of less than 0.2 PSU (Practical Salinity Unit), mainly due to the higher uncertainty in the cold environment. A small number of observations near the seawater freezing point show a significant fresh bias caused by sea ice. A systematic bias of the SMAP SSS product is found in the mid-latitudes: positive bias tends to occur north (south) of 45°N(S) and negative bias is more common in 25°N(S)–45°N(S) bands, likely associated with the SMAP calibration scheme. A significant bias also occurs in regions with strong ocean currents and eddy activities, likely due to spatial mismatch in the highly dynamic background. Notably, satellite SSS and in situ data correlations remain good in similar environments with weaker ocean dynamic activities, implying that satellite salinity data are reliable in dynamically active regions for capturing high-resolution details. The features of the SMAP SSS shown in this work call for careful consideration by the data user community when interpreting biased values. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Flowchart of the data selection, assembly processing, and production of the final classification. The 15 maps are the geographical distribution of the 15 classes, and the coloured dots in the maps represent the ΔS value of the sample.</p>
Full article ">Figure 2
<p>Ensemble mean (blue line) and spread (grey shading) of the BIC score for increasing the number of GMM classes. The black bars are the standard deviation of the ensemble mean. The BIC scores are computed for 50 random sample groups, each consisting of 90% of the total profiles.</p>
Full article ">Figure 3
<p>Visualisation of the classification results. For each class, the mean value of SST, rain rate, and wind speed is plotted as a 3D coordinate. (<b>a</b>) is the mean values of each class, the size of the marker represents the sample size of the class, and the colour of the marker represents the mean ΔS of the class. To better illustrate the spread of the classes and without hiding the small classes, we subdivided the classes into 3 subplots according to different temperature ranges. (<b>b</b>) Classes with mean SST below 10 °C, corresponding to the triangle markers in (<b>a</b>); (<b>c</b>) between 10–20 °C, corresponding to the square markers in (<b>a</b>); (<b>d</b>) above 20 °C, corresponding to the round markers in (<b>a</b>). The <span class="html-italic">x</span>-axis is SST, the <span class="html-italic">y</span>-axis is wind speed, and the <span class="html-italic">z</span>-axis is rain rate. Rain rate is plotted in log scale for ease of visualisation in (<b>b</b>–<b>d</b>). The details of each class are referred to in <a href="#remotesensing-16-03084-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Classes with mean SST higher than 25 °C. (<b>a</b>,<b>b</b>,<b>e</b>–<b>g</b>) Scatterplot maps of ΔS (unit: PSU) in the class K11, K13, K3, K15, and K8, respectively. The dotted area in (<b>a</b>) is where the number of members exceeds 200 in a 5° × 5° grid cell and the samples exceed 12. Regions where samples are insufficient for identifying the predominant season are discarded. (<b>c</b>,<b>d</b>,<b>h</b>–<b>j</b>) Prevailing season of the observations in the same classes above. Colours represent over 50% of the observations in the area being taken in the same season: blue is December to February of next year, green is March to May, red is June to August, orange is September to November, and grey means there is no prevailing season in the area.</p>
Full article ">Figure 5
<p>Classes with mean SST between 10 °C and 25 °C. (<b>a</b>,<b>b</b>,<b>e</b>–<b>g</b>) Scatterplot maps of ΔS in classes K1, K6, K9, K14, and K7, respectively. (<b>c</b>,<b>d</b>,<b>h</b>–<b>j</b>) Prevailing season of the observations. The legend is the same as <a href="#remotesensing-16-03084-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 6
<p>Scatterplot of all SMAP SSS bias observations over a PSU (<span class="html-italic">x</span>-axis) and latitude (<span class="html-italic">y</span>-axis) plot. The coloured shading represents the observation count in a 0.02 PSU and 0.5° grid size. The overlaid dashed lines are the mean rain rate (black) and the mean SSS (red), respectively, along the latitude. The mean rain rate and SSS values are in the top <span class="html-italic">x</span>-axis.</p>
Full article ">Figure 7
<p>Classes with mean SST lower than 10 °C. (<b>a</b>,<b>b</b>) Scatterplot maps of ΔS in classes K2 and K10, respectively. (<b>c</b>,<b>d</b>) Prevailing season of the observations in classes K2 and K10, respectively. The legend is the same as <a href="#remotesensing-16-03084-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 8
<p>The distribution of members in K12 and its relationship with sea ice concentration. (<b>a</b>) Scatterplot map of K12, where the colour represents ΔS. (<b>b</b>) Prevailing season of the observations. (<b>c</b>) Scatter plot of observations with sea ice presence within 50 km, with the colour representing the percentage of ice concentration. (<b>d</b>) Observations and mean ΔS concerning sea ice concentration. (<b>e</b>) Scatterplot within the classification parameter space, with the <span class="html-italic">x</span>-, <span class="html-italic">y</span>-, and <span class="html-italic">z</span>-axes representing SST, wind speed, and rain rate, respectively, and the colour of the marker representing ΔS.</p>
Full article ">Figure 9
<p>The distribution of members in K4 and its relationship with precipitation. (<b>a</b>) Scatterplot map of K4. (<b>b</b>) Prevailing season of the observations. (<b>c</b>) Annual mean precipitation. (<b>d</b>) Relations between ΔS and rain rate, the colour is the member count in the corresponding ΔS and rain rate. (<b>e</b>) Scatterplot for classification parameters, same as in <a href="#remotesensing-16-03084-f008" class="html-fig">Figure 8</a>e. The observation count in (<b>d</b>) is calculated with the bin size of 0.1 PSU along the <span class="html-italic">x</span>-axis and 2.5 mm/day along the <span class="html-italic">y</span>-axis.</p>
Full article ">Figure 10
<p>The distribution of members in K5 and its relationship with sea surface current. (<b>a</b>) Scatter plot of K5. (<b>b</b>) Prevailing season of the observations. (<b>c</b>) Annual mean Eddy Kinetic Energy (EKE) of surface current (shading) overlaps with the mean velocity of sea surface current (contour, unit: m/s). (<b>d</b>) Snapshot of SMAP SSS and ocean surface current. The colour shading is SSS, the quiver is current, and the red pentagram marker is Argo observation.</p>
Full article ">
15 pages, 2158 KiB  
Article
How Can Seasonality Influence the Performance of Recent Microwave Satellite Soil Moisture Products?
by Raffaele Albano, Teodosio Lacava, Arianna Mazzariello, Salvatore Manfreda, Jan Adamowski and Aurelia Sole
Remote Sens. 2024, 16(16), 3044; https://doi.org/10.3390/rs16163044 - 19 Aug 2024
Viewed by 537
Abstract
In addition to technical issues related to the instruments used, differences between soil moisture (SM) measured using ground-based methods and microwave remote sensing (RS) can be related to the main features of the study areas, which are intricately connected to hydraulic–hydrological conditions and [...] Read more.
In addition to technical issues related to the instruments used, differences between soil moisture (SM) measured using ground-based methods and microwave remote sensing (RS) can be related to the main features of the study areas, which are intricately connected to hydraulic–hydrological conditions and soil properties. When long-term analysis is performed, these discrepancies are mitigated by the contribution of SM seasonality and are only evident when high-frequency variations (i.e., signal anomalies) are investigated. This study sought to examine the responsiveness of SM to seasonal variations in terrestrial ecoregions located in areas covered by the in situ Romanian Soil Moisture Network (RSMN). To achieve this aim, several remote sensing-derived retrievals were considered: (i) NASA’s Soil Moisture Active and Passive (SMAP) L4 V5 model assimilated product data; (ii) the European Space Agency’s Soil Moisture and Ocean Salinity INRA–CESBIO (SMOS-IC) V2.0 data; (iii) time-series data extracted from the H115 and H116 SM products, which are derived from the analysis of Advanced Scatterometer (ASCAT) data acquired via MetOp satellites; (iv) Copernicus Global Land Service SSM 1 km data; and (v) the “combined” European Space Agency’s Climate Change Initiative for Soil Moisture (ESA CCI SM) product v06.1. An initial assessment of the performance of these products was conducted by checking the anomaly of long-term fluctuations, quantified using the Absolute Variation of Local Change of Environment (ALICE) index, within a time frame spanning 2015 to 2020. These correlations were then compared with those based on raw data and anomalies computed using a moving window of 35 days. Prominent correlations were observed with the SMAP L4 dataset and across all ecoregions, and the Balkan mixed forests (646) exhibited strong concordance regardless of the satellite source (with a correlation coefficient RALICE > 0.5). In contrast, neither the Central European mixed forests (No. 654) nor the Pontic steppe (No. 735) were adequately characterized by any satellite dataset (RALICE < 0.5). Subsequently, the phenological seasonality and dynamic behavior of SM were computed to investigate the effects of the wetting and drying processes. Notably, the Central European mixed forests (654) underwent an extended dry phase (with an extremely low p-value of 2.20 × 10−16) during both the growth and dormancy phases. This finding explains why the RSMN showcases divergent behavior and underscores why no satellite dataset can effectively capture the complexities of the ecoregions covered by this in situ SM network. Full article
(This article belongs to the Special Issue Remote Sensing of Climate-Related Hazards)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>General study workflow: the blue boxes indicate pre-processing phases, and the grey and dark gray boxes indicate the steps related to the ALICE index and the phenological seasonality and dynamic behavior of soil moisture (SM), respectively. The figure is adapted from [<a href="#B13-remotesensing-16-03044" class="html-bibr">13</a>]. Please note that the “SM Dynamic” analysis refers to the work of Manfreda et al., 2007 in [<a href="#B22-remotesensing-16-03044" class="html-bibr">22</a>].</p>
Full article ">Figure 2
<p>Ecoregions analyzed in the present study and the Romanian Soil Moisture Network (RSMN).</p>
Full article ">Figure 3
<p>Growth (blue) and dormancy (light gray) phases in each ecoregion. Derived from [<a href="#B13-remotesensing-16-03044" class="html-bibr">13</a>].</p>
Full article ">Figure 4
<p>Pearson Correlation Coefficient between ASCAT time series and in situ ISMN boxplots for the following ecoregions: No. 646 Balkan mixed forests, No. 654 Central European mixed forests, No. 661 East European forest steppe, No. 674 Pannonian mixed forests, and No. 735 Pontic steppe. Derived from [<a href="#B13-remotesensing-16-03044" class="html-bibr">13</a>].</p>
Full article ">Figure 5
<p>Overall, the growth phase and dormancy phase surface soil moisture (SSM) frequency distribution in the Balkan mixed forests (646), Central European mixed forests (654), East European forest steppe (661), Pannonian mixed forests (674), and Pontic steppe (735). N.B. A different maximum <span class="html-italic">y</span>-value was employed depending on the sample distribution in order to avoid losing details of the curve shape. Adapted from [<a href="#B13-remotesensing-16-03044" class="html-bibr">13</a>].</p>
Full article ">
22 pages, 30326 KiB  
Article
Spatially Interpolated CYGNSS Data Improve Downscaled 3 km SMAP/CYGNSS Soil Moisture
by Liza J. Wernicke, Clara C. Chew and Eric E. Small
Remote Sens. 2024, 16(16), 2924; https://doi.org/10.3390/rs16162924 - 9 Aug 2024
Viewed by 912
Abstract
Soil moisture data with both a fine spatial scale and a short global repeat period would benefit many hydrologic and climatic applications. Since the radar transmitter malfunctioned on NASA’s Soil Moisture Active Passive (SMAP) in 2015, SMAP soil moisture has been downscaled using [...] Read more.
Soil moisture data with both a fine spatial scale and a short global repeat period would benefit many hydrologic and climatic applications. Since the radar transmitter malfunctioned on NASA’s Soil Moisture Active Passive (SMAP) in 2015, SMAP soil moisture has been downscaled using numerous alternative fine-resolution data. In this paper, we describe the creation and validation of a new downscaled 3 km soil moisture dataset, which is the culmination of previous work. We downscaled SMAP enhanced 9 km brightness temperatures by merging them with L-band Cyclone Global Navigation Satellite System (CYGNSS) reflectivity data, using a modified version of the SMAP active–passive brightness temperature algorithm. We then calculated 3 km SMAP/CYGNSS soil moisture using the resulting 3 km SMAP/CYGNSS brightness temperatures and the SMAP single-channel vertically polarized soil moisture algorithm (SCA-V). To remedy the sparse daily coverage of CYGNSS data at a 3 km spatial resolution, we used spatially interpolated CYGNSS data to downscale SMAP soil moisture. 3 km interpolated SMAP/CYGNSS soil moisture matches the SMAP repeat period of ~2–3 days, providing a soil moisture dataset with both a fine spatial scale and a short repeat period. 3 km interpolated SMAP/CYGNSS soil moisture, upscaled to 9 km, has an average correlation of 0.82 and an average unbiased root mean square difference (ubRMSD) of 0.035 cm3/cm3 using all SMAP 9 km core validation sites (CVSs) within ±38° latitude. The observed (not interpolated) SMAP/CYGNSS soil moisture did not perform as well at the SMAP 9 km CVSs, with an average correlation of 0.68 and an average ubRMSD of 0.048 cm3/cm3. A sensitivity analysis shows that CYGNSS reflectivity is likely responsible for most of the uncertainty in downscaled SMAP/CYGNSS soil moisture. The success of 3 km SMAP/CYGNSS soil moisture demonstrates that Global Navigation Satellite System–Reflectometry (GNSS-R) observations are effective for downscaling soil moisture. Full article
(This article belongs to the Special Issue Microwave Remote Sensing of Soil Moisture II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A spatial comparison of (<b>a</b>) 3 km observed CYGNSS reflectivity, (<b>b</b>) 3 km interpolated CYGNSS reflectivity, and (<b>c</b>) 9 km SMAP brightness temperature. All data are from 31 March 2018. The spatial coverage of observed CYGNSS reflectivity over the displayed landmass is 10.9%.</p>
Full article ">Figure 2
<p>Workflow depicting the steps required to calculate 3 km SMAP/CYGNSS soil moisture.</p>
Full article ">Figure 3
<p>(<b>a</b>) Depiction of SMAP/CYGNSS temporal merging periods and brightness temperature time series from May 2020 to June 2020. Black lines denote the occurrence of a SMAP observation and alternating white and gray shaded regions denote the temporal merging periods of ±half the time between successive SMAP observations. Time series shows the increased temporal frequency of 3 km interpolated SMAP/CYGNSS brightness temperature compared to 3 km observed SMAP/CYGNSS brightness temperature. All data are from the 3 km grid cell at 33.194°N and 88.024°W, denoted with a red diamond in (<b>b</b>,<b>d</b>). Bottom row: the difference in spatial coverage for an example 33 × 33 km box (black square), centered on the 9 × 9 km grid cell (red square) at 33.166°N and 87.993°W, on 18 June 2020. (<b>b</b>) All observed CYGNSS reflectivity values (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Γ</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> </mrow> </semantics></math>) used to calculate (<b>c</b>) observed <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Γ</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math>. (<b>d</b>) All interpolated CYGNSS reflectivity values (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Γ</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> </mrow> </semantics></math>) used to calculate (<b>e</b>) interpolated <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Γ</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Boxplots of the daily standard deviation of soil moisture over the continental United States (25–38°N, 75–125°E) for each day from April 2017 to December 2021. The comparison includes 3 km observed and interpolated SMAP/CYGNSS soil moisture, observed and interpolated SMAP/CYGNSS soil moisture upscaled to 9 km, and 9 km SMAP soil moisture. (<b>b</b>) Boxplots of the fractional spatial coverage of 3 km observed and interpolated SMAP/CYGNSS soil moisture, compared to 9 km SMAP soil moisture, calculated using all data within the latitudinal range of ±37° for each day during the year 2020. Blue boxes indicate the interquartile ranges, red lines indicate the medians, and black plus signs denote all values outside of the interquartile range.</p>
Full article ">Figure 5
<p>All images and maps show a ~150 km × 100 km region of northwest Texas, USA (35.4°–36.4°N and 101°–102.5°W). (<b>a</b>) Terra/MODIS reflectance image on 20 August 2020 [<a href="#B50-remotesensing-16-02924" class="html-bibr">50</a>], showing areas with irrigated cropland adjacent to non-agricultural areas. (<b>b</b>) 3 km interpolated SMAP/CYGNSS soil moisture. (<b>c</b>) 3 km observed SMAP/CYGNSS soil moisture. (<b>d</b>) 3 km SMAP/Sentinel soil moisture [<a href="#B51-remotesensing-16-02924" class="html-bibr">51</a>]. (<b>e</b>) 9 km SMAP soil moisture [<a href="#B42-remotesensing-16-02924" class="html-bibr">42</a>]. All soil moisture maps are averaged or aggregated over two periods: 1–31 July 2020 and 1–31 August 2020.</p>
Full article ">Figure 6
<p>Soil moisture maps showing the full CYGNSS latitudinal band of ±38° for (<b>a</b>) 3 km SMAP/Sentinel [<a href="#B51-remotesensing-16-02924" class="html-bibr">51</a>], (<b>b</b>) 3 km interpolated SMAP/CYGNSS, and (<b>c</b>) 9 km SMAP [<a href="#B42-remotesensing-16-02924" class="html-bibr">42</a>]. Regional soil moisture maps for (<b>d</b>) 3 km observed SMAP/CYGNSS, (<b>e</b>) 3 km interpolated SMAP/CYGNSS, and (<b>f</b>) 9 km SMAP [<a href="#B42-remotesensing-16-02924" class="html-bibr">42</a>]. All data are aggregated from 14 to 17 July 2020 to create a SMAP soil moisture map with no data gaps. The red rectangles in (<b>b</b>,<b>c</b>) indicate the location of the maps in (<b>d</b>–<b>f</b>), and the red rectangles in (<b>d</b>–<b>f</b>) indicate the location of the maps in <a href="#remotesensing-16-02924-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 7
<p>(<b>a</b>) SMAP/CYGNSS soil moisture ubRMSD (cm<sup>3</sup>/cm<sup>3</sup>) versus 9 km SMAP enhanced soil moisture ubRMSD (cm<sup>3</sup>/cm<sup>3</sup>) calculated using in situ soil moisture at SMAP 9 km CVSs. The dashed black lines denote the SMAP accuracy requirement of 0.04 cm<sup>3</sup>/cm<sup>3</sup>. (<b>b</b>) SMAP/CYGNSS soil moisture correlation versus 9 km SMAP enhanced soil moisture correlation calculated using in situ soil moisture at SMAP 9 km CVSs. Interpolated SMAP/CYGNSS soil moisture is represented with a filled circle and observed SMAP/CYGNSS soil moisture is represented with an ‘x’. Each SMAP CVS used in the study is represented by a unique color, shown in the legend. Walnut Gulch, Yanco, and TxSON each have two separate 9 km validation regions.</p>
Full article ">Figure 8
<p>Boxplots showing (<b>a</b>) ubRMSD and (<b>b</b>) correlation for the 9 km CVSs used in this study. The blue boxes show the interquartile ranges, the red lines denote the medians, and the black plus signs show all values that fall outside of the interquartile range. Each panel depicts a comparison of the in situ CVS soil moisture and 1) 9 km SMAP enhanced soil moisture, 2) upscaled 3 km interpolated SMAP/CYGNSS soil moisture, 3) upscaled 3 km interpolated SMAP/CYGNSS soil moisture, thinned to only include data on the same days as observed SMAP/CYGNSS soil moisture, and 4) upscaled 3 km observed SMAP/CYGNSS soil moisture.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) include time series of (1) 9 km averaged in situ soil moisture, (2) upscaled 3 km interpolated SMAP/CYGNSS soil moisture, (3) upscaled 3 km observed SMAP/CYGNSS soil moisture, and (4) 9 km SMAP enhanced soil moisture. (<b>a</b>) Time series spanning from 1 April 2017 to 31 March 2021. The dashed gray box indicates the period of (<b>b</b>), which is a time series spanning from 1 January 2020 to 31 December 2020. (<b>c</b>) Scatter plot comparing TxSON in situ soil moisture with upscaled 3 km observed SMAP/CYGNSS soil moisture. (<b>d</b>) Scatter plot comparing TxSON in situ soil moisture with upscaled 3 km interpolated SMAP/CYGNSS soil moisture. (<b>e</b>) Scatter plot comparing TxSON in situ soil moisture with 9 km SMAP enhanced soil moisture. All scatter plots were created using data spanning from 1 April 2017 to 31 March 2021. The location for all data is the TxSON CVS in Texas, USA (30.271°N, 98.729°W). TxSON in situ data were retrieved from [<a href="#B53-remotesensing-16-02924" class="html-bibr">53</a>]. All satellite soil moisture is corrected for bias.</p>
Full article ">Figure 10
<p>The RMSE caused by the uncertainty of each parameter in the SMAP/CYGNSS brightness temperature algorithm and the SMAP SCA-V soil moisture algorithm, determined by varying each parameter by its estimated uncertainty. Because the CYGNSS uncertainty is unknown and significantly affects SMAP/CYGNSS soil moisture, various CYGNSS reflectivity uncertainty estimates are included. All RMSE estimates are for 3 km interpolated SMAP/CYGNSS soil moisture.</p>
Full article ">
25 pages, 5377 KiB  
Article
Investigating FWI Moisture Codes in Relation to Satellite-Derived Soil Moisture Data across Varied Resolutions
by Hatice Atalay, Ayse Filiz Sunar and Adalet Dervisoglu
Fire 2024, 7(8), 272; https://doi.org/10.3390/fire7080272 - 5 Aug 2024
Viewed by 749
Abstract
In the Mediterranean region, particularly in Antalya, southern Türkiye, rising forest fire risks due to climate change threaten ecosystems, property, and lives. Reduced soil moisture during the growing season is a key factor increasing fire risk by stressing plants and lowering fuel moisture [...] Read more.
In the Mediterranean region, particularly in Antalya, southern Türkiye, rising forest fire risks due to climate change threaten ecosystems, property, and lives. Reduced soil moisture during the growing season is a key factor increasing fire risk by stressing plants and lowering fuel moisture content. This study assessed soil moisture and fuel moisture content (FMC) in ten fires (2019–2021) affecting over 50 hectares. The Fire Weather Index (FWI) and its components (FFMC, DMC, DC) were calculated using data from the General Directorate of Meteorology, EFFIS (8 km), and ERA5 (≈28 km) satellite sources. Relationships between FMCs, satellite-based soil moisture datasets (SMAP, SMOS), and land surface temperature (LST) data (MODIS, Landsat 8) were analyzed. Strong correlations were found between FWI codes and satellite soil moisture, particularly with SMAP. Positive correlations were observed between LST and FWIs, while negative correlations were evident with soil moisture. Statistical models integrating in situ soil moisture and EFFIS FWI (R: −0.86, −0.84, −0.83 for FFMC, DMC, DC) predicted soil moisture levels during extended fire events effectively, with model accuracy assessed through RMSE (0.60–3.64%). The SMAP surface (0–5 cm) dataset yielded a lower RMSE of 0.60–2.08%, aligning with its higher correlation. This study underlines the critical role of soil moisture in comprehensive fire risk assessments and highlights the necessity of incorporating modeled soil moisture data in fire management strategies, particularly in regions lacking comprehensive in situ monitoring. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the study area (Antalya region); created using ArcGIS Pro (version 3.1, Esri, Redlands, CA, USA). (<b>b</b>) geographical distribution of climate types in Türkiye based on Köppen–Geiger climate system (the map is retrieved from [<a href="#B70-fire-07-00272" class="html-bibr">70</a>,<a href="#B71-fire-07-00272" class="html-bibr">71</a>]).</p>
Full article ">Figure 2
<p>Locations and sizes of ten forest fires analyzed (data sourced from EFFIS), along with meteorological stations in the study area; created using ArcGIS Pro (version 3.1, Esri, Redlands, CA, USA).</p>
Full article ">Figure 3
<p>Important meteorological parameters considered for each date in the analysis.</p>
Full article ">Figure 4
<p>Distribution of LULC classes as a percentage of burned areas in ten fires analyzed.</p>
Full article ">Figure 5
<p>Categorization of forest floor fuels based on the fuel moisture codes of the FWI System [<a href="#B28-fire-07-00272" class="html-bibr">28</a>].</p>
Full article ">Figure 6
<p>Flowchart of the study.</p>
Full article ">
24 pages, 13466 KiB  
Article
Significance of Multi-Variable Model Calibration in Hydrological Simulations within Data-Scarce River Basins: A Case Study in the Dry-Zone of Sri Lanka
by Kavini Pabasara, Luminda Gunawardhana, Janaka Bamunawala, Jeewanthi Sirisena and Lalith Rajapakse
Hydrology 2024, 11(8), 116; https://doi.org/10.3390/hydrology11080116 - 5 Aug 2024
Viewed by 1235
Abstract
Traditional hydrological model calibration using limitedly available streamflow data often becomes inadequate, particularly in dry climates, as the flow regimes may abruptly vary from arid conditions to devastating floods. Newly available remote-sensing-based datasets can be supplemented to overcome such inadequacies in hydrological simulations. [...] Read more.
Traditional hydrological model calibration using limitedly available streamflow data often becomes inadequate, particularly in dry climates, as the flow regimes may abruptly vary from arid conditions to devastating floods. Newly available remote-sensing-based datasets can be supplemented to overcome such inadequacies in hydrological simulations. To address this shortcoming, we use multi-variable-based calibration by setting up and calibrating a lumped-hydrological model using observed streamflow and remote-sensing-based soil moisture data from Soil Moisture Active Passive Level 4. The proposed method was piloted at the Maduru Oya River Basin, Sri Lanka, as a proof of concept. The relative contributions from streamflow and soil moisture were assessed and optimised via the Kling–Gupta Efficiency (KGE). The Generalized Reduced Gradient non-linear solver function was used to optimise the Tank Model parameters. The findings revealed satisfactory performance in streamflow simulations under single-variable model validation (KGE of 0.85). Model performances were enhanced by incorporating soil moisture data (KGE of 0.89), highlighting the capability of the proposed multi-variable calibration technique for improving the overall model performance. Further, the findings of this study highlighted the instrumental role of remote sensing data in representing the soil moisture dynamics of the study area and the importance of using multi-variable calibration to ensure robust hydrological simulations of river basins in dry climates. Full article
Show Figures

Figure 1

Figure 1
<p>Padiyathalawa sub-watershed in Maduru Oya river basin.</p>
Full article ">Figure 2
<p>Conceptual illustration of the Tank Model setup.</p>
Full article ">Figure 3
<p>Sensitivity analysis of the Tank Model parameters for (<b>A</b>) peak flow and (<b>B</b>) first quartile (i.e., 25th percentile) flow. Note: A<sub>11</sub>-runoff coefficient to estimate surface flow, Z<sub>11</sub>-runoff hole depth to estimate surface flow, A<sub>12</sub>-runoff coefficient to estimate sub-surface flow, Z<sub>12</sub>-runoff hole depth to estimate sub-surface flow, B<sub>1</sub>-infiltration coefficient of the first tank, A<sub>2</sub>-runoff coefficient to estimate intermediate flow, Z<sub>2</sub>-runoff hole depth to estimate intermediate flow, B<sub>2</sub>-infiltration coefficient of the second tank, A<sub>3</sub>-runoff coefficient to estimate sub-base flow, Z<sub>3</sub>-runoff hole depth to estimate sub-base flow, B<sub>3</sub>-infiltration coefficient of the third tank, A<sub>4</sub> runoff coefficient to estimate base flow.</p>
Full article ">Figure 4
<p>Comparison of observed and simulated hydrographs (<b>A</b>) and flow-duration curves (<b>B</b>) at the basin outlet for the calibration period (between August 2011 and September 2012).</p>
Full article ">Figure 5
<p>Comparison of observed and simulated hydrographs (<b>A</b>) and flow-duration curves (<b>B</b>) at the basin outlet for the validation period (between September 2015 and April 2016).</p>
Full article ">Figure 6
<p>Variation of model performances (i.e., the King–Gupta Efficiency (<span class="html-italic">KGE</span>)) with weighted factor (<span class="html-italic">ϑ</span>) under multi-variable calibration.</p>
Full article ">Figure 7
<p>Percentage change in Kling–Gupta Efficiency for streamflow (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>K</mi> <mi>G</mi> <mi>E</mi> </mrow> <mrow> <mi mathvariant="normal">Q</mi> </mrow> </msub> </mrow> </semantics></math>) when the optimal parameters from multi-variable calibration are used in the simulation, compared to the simulations with parameters optimised from single-variable calibration (i.e., weighted factor (<span class="html-italic">ϑ</span>) of 0).</p>
Full article ">Figure 8
<p>(<b>A</b>) Hydrograph and (<b>B</b>) FDCs for the Tank Model simulations (between September 2015 and April 2016) under calibration schemes with 0.1, 0.8, and 0.9 weighted factors (<span class="html-italic">ϑ</span>).</p>
Full article ">Figure 9
<p>(<b>A</b>) Hydrograph and (<b>B</b>) FDCs for the Tank Model simulations (between August 2011 and September 2012) under validation schemes with 0.1, 0.8, and 0.9 weighted factors (<span class="html-italic">ϑ</span>).</p>
Full article ">Figure 10
<p>Soil moisture simulation results under calibration schemes with 0.1, 0.8, and 0.9 weighted factors (<span class="html-italic">ϑ</span>) (between September 2015 and April 2016).</p>
Full article ">Figure A1
<p>Monthly precipitation at Welipitiya Coconut Estate gauging station between October 2010 and September 2016. The boxes are limited to the 25th and 75th percentiles, and the horizontal line shows the median (i.e., 50th percentile) value of the monthly data sets. Whiskers are extended to 1.5 times inter-quartile range to the top and bottom of the boxes.</p>
Full article ">Figure A2
<p>Average minimum temperature measured at Maduru Oya gauging station between October 2010 and September 2016. The boxes are limited to the 25th and 75th percentiles, and the horizontal line shows the median (i.e., 50th percentile) value of the monthly data sets. Whiskers are extended to 1.5 times inter-quartile range to the top and bottom of the boxes.</p>
Full article ">Figure A3
<p>Average maximum temperature measured at Maduru Oya gauging station between October 2010 and September 2016. The boxes are limited to the 25th and 75th percentiles, and the horizontal line shows the median (i.e., 50th percentile) value of the monthly data sets. Whiskers are extended to 1.5 times inter-quartile range to the top and bottom of the boxes.</p>
Full article ">Figure A4
<p>Monthly averaged streamflow measured at Padiyathalawa gauging station between October 2010 and September 2016. The boxes are limited to the 25th and 75th percentiles, and the horizontal line shows the median (i.e., 50th percentile) value of the monthly data sets. Whiskers are extended to 1.5 times inter-quartile range to the top and bottom of the boxes.</p>
Full article ">Figure A5
<p>Cumulative Distribution Function (CDF) Matching method-based rescaling of spatially and temporally aggregated Root Zone Soil Moisture data.</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 818
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 ">
14 pages, 5235 KiB  
Article
Mapping Extreme Wildfires Using a Critical Threshold in SMAP Soil Moisture
by Benjamin D. Goffin, Aashutosh Aryal, Quinton Deppert, Kenton W. Ross and Venkataraman Lakshmi
Remote Sens. 2024, 16(13), 2457; https://doi.org/10.3390/rs16132457 - 4 Jul 2024
Viewed by 1041
Abstract
This study analyzed the ground conditions that allowed some extreme wildfires in 2017 and 2023 to take such proportions and burn around 750,000 ha across Central Chile. Using publicly available satellite data, we examined the relationship between the burned areas from the Moderate [...] Read more.
This study analyzed the ground conditions that allowed some extreme wildfires in 2017 and 2023 to take such proportions and burn around 750,000 ha across Central Chile. Using publicly available satellite data, we examined the relationship between the burned areas from the Moderate Resolution Imaging Spectroradiometers (MODIS) and their antecedent soil moisture from the Soil Moisture Active Passive (SMAP) mission. We found that a small number of fires were responsible for disproportionately large burned areas and that these megafires (i.e., >10,000 ha) were more likely to exhibit relatively drier conditions in the months and days prior. Based on this, we tested various thresholds in low antecedent soil moisture to identify areas more prone to megafires. By differentiating the moisture conditions below and above 0.14 m3/m3, we were able to map all of the 2017 megafires, at least in part. Our classification balanced the success and errors in prediction, yielding 54.1% recall and 75.9% precision (well above the 56.3% baseline). For 2023, the burned areas could not be classified as accurately, due to differences in pre-fire conditions. Overall, our research provided new insights into the link between satellite-based soil moisture and extreme wildfire events. Among other things, this study demonstrated that certain critical thresholds in SMAP had predictive skill to identify conditions more conducive to megafires. Ultimately, this work can be expanded to other parts of the world in support of enhanced wildfire mitigation and management. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area over (<b>a</b>) regions of Central Chile affected by (<b>b</b>) major wildfires in recent years with (<b>c</b>) various climate types.</p>
Full article ">Figure 2
<p>Flowchart of the classification of burned areas from the Moderate Resolution Imaging Spectroradiometers (MODIS) using one critical threshold in antecedent soil moisture obtained from the Soil Moisture Active Passive (SMAP) mission.</p>
Full article ">Figure 3
<p>Spatial variability in 10-day soil moisture across the megafires and control areas (111 and 86 pixels, respectively) throughout the six months prior to the spread of the 2017 wildfires (shaded in gray).</p>
Full article ">Figure 4
<p>Comparison of the distributions of soil moisture pixels across the megafires and control areas over the 10 days ahead of the 2017 wildfires.</p>
Full article ">Figure 5
<p>Contrast in soil moisture across the (<b>a</b>) megafires and (<b>b</b>) control areas over the 10 days ahead of 2017 wildfires. Red and blue indicate drier and wetter conditions, respectively.</p>
Full article ">Figure 6
<p>Performance of various critical thresholds in antecedent soil moisture to classify burned areas ahead of the 2017 wildfires using (<b>a</b>) Receiver Operating Characteristic (ROC) curves and (<b>b</b>) Precision–Recall Curves (PRC).</p>
Full article ">Figure 7
<p>Classification of the burned areas using a critical threshold of 0.14 m<sup>3</sup>/m<sup>3</sup> in soil moisture over the 10 days prior to the 2017 wildfires.</p>
Full article ">Figure 8
<p>Classification of the burned areas using a critical threshold of 0.14 m<sup>3</sup>/m<sup>3</sup> in soil moisture over the 10 days prior to the 2023 wildfires.</p>
Full article ">
17 pages, 5274 KiB  
Review
Reviewing Space-Borne GNSS-Reflectometry for Detecting Freeze/Thaw Conditions of Near-Surface Soils
by Haishan Liang and Xuerui Wu
Remote Sens. 2024, 16(11), 1828; https://doi.org/10.3390/rs16111828 - 21 May 2024
Viewed by 807
Abstract
GNSS-Reflectometry, a technique that harnesses the power of microwave remote sensing, is poised to revolutionize our ability to detect and monitor near-surface soil freeze/thaw processes. This technique’s theoretical underpinnings are deeply rooted in the comprehensive explanation of the Zhang–Zhao dielectric constant model, which [...] Read more.
GNSS-Reflectometry, a technique that harnesses the power of microwave remote sensing, is poised to revolutionize our ability to detect and monitor near-surface soil freeze/thaw processes. This technique’s theoretical underpinnings are deeply rooted in the comprehensive explanation of the Zhang–Zhao dielectric constant model, which provides crucial insights into the behavior of frozen and thawed soils. The model elucidates how the dielectric properties of soil change as it transitions between frozen and thawed states, offering a scientific basis for understanding reflectivity variations. Furthermore, the theoretical framework includes a set of formulas that are instrumental in calculating reflectivity at Lower Right (LR) polarization and in deriving Dual-Polarization Differential Observables (DDMs). These calculations are pivotal for interpreting the signals captured by GNSS-R sensors, allowing for the detection of subtle changes in the soil’s surface conditions. The evolution of GNSS-R as a tool for detecting freeze/thaw phenomena has been substantiated through qualitative analyses involving multiple satellite missions, such as SMAP-R, TDS-1, and CYGNSS. These analyses have provided empirical evidence of the technique’s effectiveness, illustrating its capacity to capture the dynamics of soil freezing and thawing processes. In addition to these qualitative assessments, the application of a discriminant retrieval algorithm using data from CYGNSS and F3E GNOS-R has further solidified the technique’s potential. This algorithm contributes to refining the accuracy of freeze/thaw detection by distinguishing between frozen and thawed soil states with greater precision. The deployment of space-borne GNSS-R for monitoring near-surface freeze/thaw cycles has yielded commendable results, exhibiting robust consistency and delivering relatively precise retrieval outcomes. These achievements stand as testaments to the technique’s viability and its growing significance in the field of remote sensing. However, it is imperative to recognize and actively address certain limitations that have been highlighted in this review. These limitations serve as critical focal points for future research endeavors, directing the efforts toward enhancing the technique’s overall performance and applicability. Addressing these challenges will be essential for leveraging the full potential of GNSS-R to advance our understanding and management of near-surface soil freeze/thaw processes. Full article
(This article belongs to the Special Issue SoOP-Reflectometry or GNSS-Reflectometry: Theory and Applications)
Show Figures

Figure 1

Figure 1
<p>The relationship between the real (<b>left</b>) and imaginary (<b>right</b>) parts of the dielectric constant and soil temperature under different soil moisture conditions. There are no units for the real part and the imaginary part of the dielectric constant.</p>
Full article ">Figure 2
<p>DDMs for frozen soil (<b>a</b>) and thawed soil (<b>b</b>); Delay waveform (DW) in (<b>c</b>) for frozen soil (blue line) and thawed soil (red line).</p>
Full article ">Figure 3
<p>Development of the space-borne GNSS-R missions.</p>
Full article ">Figure 4
<p>SMAP F/T cycle fraction and the percentage of frozen pixels versus the TDS−1 reflectivity [<a href="#B27-remotesensing-16-01828" class="html-bibr">27</a>]. While the green line indicate the surface reflectivity, and the red line demonstrate the percentage of frozen pixels.</p>
Full article ">Figure 5
<p>Specular points of CYGNSS in the Qinghai–Tibet Plateau on 1 January 2018; water bodies are colored blue [<a href="#B29-remotesensing-16-01828" class="html-bibr">29</a>].</p>
Full article ">Figure 6
<p>The International Geosphere-Biosphere Programme (IGBP) land cover type on the Tibetan Plateau [<a href="#B29-remotesensing-16-01828" class="html-bibr">29</a>].</p>
Full article ">Figure 7
<p>The time series of surface reflectivity versus the Soil Moisture Active Passive (SMAP) freeze ratio area under different land cover types on the Tibetan Plateau. The blue lines are the surface reflectivity and the red lines demonstrated the F/T pixels.</p>
Full article ">Figure 8
<p>The IGBP Land Cover Types in the target area.</p>
Full article ">Figure 9
<p>The SCA distribution in January (<b>a</b>) and July (<b>b</b>) of 2019 [<a href="#B33-remotesensing-16-01828" class="html-bibr">33</a>].</p>
Full article ">Figure 10
<p>Surface thawed state (<b>a</b>–<b>c</b>) from January 2019 to March 2019, the months for (<b>a</b>–<b>c</b>) are January, February and March. and surface frozen state (<b>d</b>–<b>f</b>) from July 2019 to September 2019 while the months for (<b>d</b>–<b>f</b>) are July, August and September, obtained using CYGNSS data [<a href="#B33-remotesensing-16-01828" class="html-bibr">33</a>].</p>
Full article ">Figure 11
<p>The land cover and land use map in the Arctic Circle [<a href="#B31-remotesensing-16-01828" class="html-bibr">31</a>].</p>
Full article ">Figure 12
<p>The SR (Surface reflectivity) ratio factors (left <span class="html-italic">Y</span>-axis) in blue color and the SMAP F/T cycle values (right <span class="html-italic">Y</span>-axis) for three types of land surfaces ((<b>a</b>) barren, (<b>b</b>) LowVeg, (<b>c</b>) Forest) in the Arctic Circle during the period from 10 July 2021 to 10 July 2022; the day of the year (DOY) is used to represent the date during the studied period in the <span class="html-italic">X</span>-axis.</p>
Full article ">
16 pages, 4039 KiB  
Article
A Soil Moisture and Vegetation-Based Susceptibility Mapping Approach to Wildfire Events in Greece
by Kyriakos Chaleplis, Avery Walters, Bin Fang, Venkataraman Lakshmi and Alexandra Gemitzi
Remote Sens. 2024, 16(10), 1816; https://doi.org/10.3390/rs16101816 - 20 May 2024
Cited by 2 | Viewed by 997
Abstract
Wildfires in Mediterranean areas are becoming more frequent, and the fire season is extending toward the spring and autumn months. These alarming findings indicate an urgent need to develop fire susceptibility methods capable of identifying areas vulnerable to wildfires. The present work aims [...] Read more.
Wildfires in Mediterranean areas are becoming more frequent, and the fire season is extending toward the spring and autumn months. These alarming findings indicate an urgent need to develop fire susceptibility methods capable of identifying areas vulnerable to wildfires. The present work aims to uncover possible soil moisture and vegetation condition precursory signals of the largest and most devastating wildfires in Greece that occurred in 2021, 2022, and 2023. Therefore, the time series of two remotely sensed datasets–MAP L4 Soil Moisture (SM) and Landsat 8 NDVI, which represent vegetation and soil moisture conditions—were examined before five destructive wildfires in Greece during the study period. The results of the analysis highlighted specific properties indicative of fire-susceptible areas. NDVI in all fire-affected areas ranged from 0.13 to 0.35, while mean monthly soil moisture showed negative anomalies in the spring periods preceding fires. Accordingly, fire susceptibility maps were developed, verifying the usefulness of remotely sensed information related to soil moisture and NDVI. This information should be used to enhance fire models and identify areas at risk of wildfires in the near future. Full article
(This article belongs to the Special Issue Satellite Soil Moisture Estimation, Assessment, and Applications)
Show Figures

Figure 1

Figure 1
<p>Location map of the wildfires analyzed in the present work. Numbers correspond to those indicated in <a href="#remotesensing-16-01816-t001" class="html-table">Table 1</a>. Red designated areas correspond to wildfires that are used for the method development (Fires 1–5), while yellow is used for verification (Fires 6–11) only.</p>
Full article ">Figure 2
<p>Fire susceptibility map flow chart.</p>
Full article ">Figure 3
<p>Time series of soil moisture and (<b>a</b>) number of fires, (<b>b</b>) burned area, in Greece (2015–2023).</p>
Full article ">Figure 4
<p>Comparison of in situ SM data and SMAP L4 surface SM and root zone SM in the Rhodope area (NE Greece) during the study period.</p>
Full article ">Figure 5
<p>Time series graphs of NDVI and SMAP L4 surface SM anomalies for five wildfires during 2021–2023 in Greece. Arrows indicate the date of wildfire occurrence.</p>
Full article ">Figure 6
<p>Spatial distribution of NDVI (<b>a</b>–<b>c</b>) and SM anomalies (<b>d</b>–<b>f</b>) satisfying the fire susceptibility criteria in the broader area of Greece for 2021, 2022, and 2023.</p>
Full article ">Figure 7
<p>Spatial distribution of fire susceptibility (<b>a</b>–<b>c</b>), VDI at the NUTS3 level (<b>d</b>–<b>f</b>), and cross-tabulation results with fire susceptibility mapping of the present work (<b>g</b>–<b>i</b>) for Greece during 2021, 2022, and 2023.</p>
Full article ">
25 pages, 9092 KiB  
Article
Constrained Iterative Adaptive Algorithm for the Detection and Localization of RFI Sources Based on the SMAP L-Band Microwave Radiometer
by Xinxin Wang, Xiang Wang, Lin Wang, Jianchao Fan and Enbo Wei
Remote Sens. 2024, 16(10), 1791; https://doi.org/10.3390/rs16101791 - 18 May 2024
Viewed by 655
Abstract
The Soil Moisture Active Passive (SMAP) satellite carries an L-band microwave radiometer. This sensor can be used to observe global soil moisture (SM) and sea surface salinity (SSS) within the protected L-band spectrum (1400–1427 MHz). Owing to the complex effects of radio frequency [...] Read more.
The Soil Moisture Active Passive (SMAP) satellite carries an L-band microwave radiometer. This sensor can be used to observe global soil moisture (SM) and sea surface salinity (SSS) within the protected L-band spectrum (1400–1427 MHz). Owing to the complex effects of radio frequency interference (RFI), the SM and SSS data are missing or have low accuracy. In this paper, a constrained iterative adaptive algorithm for the detection, identification, and localization of RFI sources is designed, named MICA-BEID. The algorithm synthesizes antenna temperatures for the third and fourth Stokes parameters before RFI filtering, creating a new polarization parameter called WSPDA, designed to approximate the level of RFI interference on the L-band microwave radiometer. The algorithm then utilizes the WSPDA intensity and distribution density of RFI detection samples to enhance the identification and classification of RFI sources across various intensity levels. By utilizing statistical methods such as the probability density function (PDF) and the cumulative distribution function (CDF), the algorithm dynamically adjusts adaptive parameters, including the RFI detection threshold and the maximum effective radius of RFI sources. Through the application of multiple iterative clustering methods, the algorithm can adaptively detect and identify RFI sources at various satellite orbits and intensity levels. Through extensive comparative analysis with other localization results and known RFI sources, the MICA-BEID algorithm can achieve optimal localization accuracy of approximately 1.2 km. The localization of RFI sources provides important guidance for identifying and turning off illegal RFI sources. Moreover, the localization and long-time-series characteristic analysis of RFI sources that cannot be turned off is of significant value for simulating the spatial distribution characteristics of localized RFI source intensity in local areas. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart for the constrained iterative adaptive algorithm.</p>
Full article ">Figure 2
<p>Statistical histogram of calculation results of SMAP L1B half-orbit data parameter <span class="html-italic">W<sub>SPDA</sub></span> for (<b>a</b>) the ascending orbit and (<b>b</b>) the descending orbit.</p>
Full article ">Figure 3
<p>Spatial distribution of SMAP RFI detection sample data. (<b>a</b>,<b>b</b>) RFI detection samples inside (green) and on the edges (red) of the ascending and descending orbits, respectively; (<b>c</b>,<b>d</b>) SMAP RFI SRDF samples (blue), SPDA samples (red) and overlapped samples (green) in ascending and descending orbits, respectively, with the samples on the edges being removed.</p>
Full article ">Figure 3 Cont.
<p>Spatial distribution of SMAP RFI detection sample data. (<b>a</b>,<b>b</b>) RFI detection samples inside (green) and on the edges (red) of the ascending and descending orbits, respectively; (<b>c</b>,<b>d</b>) SMAP RFI SRDF samples (blue), SPDA samples (red) and overlapped samples (green) in ascending and descending orbits, respectively, with the samples on the edges being removed.</p>
Full article ">Figure 4
<p>Local spatial distribution of SMAP RFI detection sample data at the water–land boundary. The background is the spatial interpolation results of <span class="html-italic">W<sub>SPDA</sub></span>. The detection results of the descending orbit at a detection threshold of 5.5 K are in red, and those at a detection threshold of 9.3 K are in green.</p>
Full article ">Figure 5
<p>Spatial distribution of RFI detection samples after interpolation for (<b>a</b>) land only and (<b>b</b>) at the water–land boundary.</p>
Full article ">Figure 6
<p><span class="html-italic">R<sub>max</sub></span> statistical analysis examples. (<b>a</b>,<b>b</b>) Statistical results of RFI detection samples for fore and aft looks, respectively; (<b>c</b>,<b>d</b>) spatial distribution of <span class="html-italic">W<sub>SPDA</sub></span> intensity of clusters for fore and aft looks, respectively. The background is the spatial interpolation results of <span class="html-italic">W<sub>SPDA</sub></span> for all fore and aft detection samples.</p>
Full article ">Figure 7
<p>The changing characteristics of clustering identification results with iteration count. (<b>a</b>) Number of detected samples; (<b>b</b>) number of clustered samples; (<b>c</b>) number of RFI sources identified; (<b>d</b>) <span class="html-italic">W<sub>max</sub></span> of RFI sources.</p>
Full article ">Figure 8
<p>Spatial distribution of RFI detection samples for the SMAP satellite’s fore and aft looks footprints.</p>
Full article ">Figure 9
<p>Filtering results of clusters at the water–land boundary. (<b>a</b>) All clustering results for all samples, where various colors denote different clusters; (<b>b</b>) results of cluster filtering, with red indicating the removed clusters and green indicating the retained clusters.</p>
Full article ">Figure 10
<p>A typical example of the effects of Faraday rotation during a period of high geomagnetic disturbance on 27 February 2023.</p>
Full article ">Figure 11
<p>Statistical analysis results of the impact of Faraday rotation. (<b>a</b>) Data from half-orbit of the SMAP satellite. (<b>b</b>) Data within the blue box range, as depicted in <a href="#remotesensing-16-01791-f010" class="html-fig">Figure 10</a>, of a typical example. The red line in both (<b>a</b>) and (<b>b</b>) represents the result of linear fitting.</p>
Full article ">Figure 12
<p>(<b>a</b>) Two- and (<b>b</b>) three-dimensional intensity distributions of <span class="html-italic">W<sub>SPDA</sub></span> of clusters for descending orbit. Red and black circles represent fore and aft antenna scanning orbits, respectively.</p>
Full article ">Figure 13
<p>Statistical diagrams of spatial distribution of the normalized parameter with distance for descending orbits.</p>
Full article ">Figure 14
<p>RFI identification and determination criteria. The red stars represent outlier data that have been removed.</p>
Full article ">Figure 15
<p>Spatial distribution characteristic of long-time-series RFI location data points and the RFI location centroid.</p>
Full article ">Figure 16
<p>The characteristics of the relationship between the <span class="html-italic">W<sub>SPDA</sub></span> parameter and the antenna temperatures for (<b>a</b>) horizontal (<span class="html-italic">Ta_h</span>) and (<b>b</b>) vertical (<span class="html-italic">Ta_v</span>) polarizations.</p>
Full article ">Figure 17
<p>Comparison results between MICA-BEID algorithm and SMAP RFI surveys. (<b>a</b>) Comparison of RFI location results; (<b>b</b>) comparison of RFI levels.</p>
Full article ">Figure 18
<p>Comparative analysis of MICA-BEID algorithm RFI localization results with known RFI sources on a long-term scale. (<b>a</b>) Hebei. (<b>b</b>) Tianjin. (<b>c</b>) Shandong. (<b>d</b>) Hunan.</p>
Full article ">Figure 19
<p>Comparative analysis of RFI localization results with MICA-BEID algorithm with <span class="html-italic">W<sub>SPDA</sub></span> of the SMOS satellite.</p>
Full article ">Figure 20
<p>Comparative analysis of RFI localization results with MICA-BEID algorithm with the average TB-RMSE.</p>
Full article ">
26 pages, 7121 KiB  
Article
Morphological Study before and after Thermal Treatment of Polymer-Polymer Mixed-Matrix Membranes for Gas Separations
by Pedro Pradanos, Cenit Soto, Francisco Javier Carmona, Ángel E. Lozano, Antonio Hernández and Laura Palacio
Polymers 2024, 16(10), 1397; https://doi.org/10.3390/polym16101397 - 14 May 2024
Viewed by 1068
Abstract
A good integration of the polymer materials that form a mixed-matrix membrane (MMM) for gas separation is essential to reaching interesting permselective properties. In this work, a porous polymer network (PPN), obtained by combining triptycene and trifluoroacetophenone, has been used as a filler, [...] Read more.
A good integration of the polymer materials that form a mixed-matrix membrane (MMM) for gas separation is essential to reaching interesting permselective properties. In this work, a porous polymer network (PPN), obtained by combining triptycene and trifluoroacetophenone, has been used as a filler, which was blended with two o-hydroxypolyamides (HPAs) that act as polymer matrices. These polymer matrices have been thermally treated to induce a thermal rearrangement (TR) of the HPAs to polybenzoxazoles (β-TR-PBOs) through a solid-state reaction. For its structural study, various techniques have been proposed that allow us to undertake a morphological investigation into the integration of these materials. To access the internal structure of the MMMs, three different methods were used: a polishing process for the material surface, the partial dissolution of the polymer matrix, or argon plasma etching. The argon plasma technique has not only revealed its potential to visualize the internal structure of these materials; it has also been proven to allow for the transformation of their permselective properties. Force modulation and phase contrast in lift-mode techniques, along with the topographic images obtained via the tapping mode using a scanning probe microscope (SPM), have allowed us to study the distribution of the filler particles and the interaction of the polymer and the filler. The morphological information obtained via SPM, along with that of other more commonly used techniques (SEM, TGA, DSC, FTIR, WASX, gas adsorption, and permeability measurements), has allowed us to postulate the most probable structural configuration in this type of system. Full article
(This article belongs to the Special Issue Advanced Polymer Membranes for Adsorption and Separation Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Simple 6F0% (<b>a</b>–<b>d</b>): without grinding treatment (<b>a</b>); griding paper 15 µm (<b>b</b>); hard polishing cloth + diamond suspension 3 µm (<b>c</b>); soft polishing cloth + alumina suspension 0.05 µm (<b>d</b>); 6F10% with all polishing treatments (<b>e</b>); 6Ftr10% with all polishing treatments (<b>f</b>).</p>
Full article ">Figure 2
<p>Percentage of surface corresponding to holes versus % <span class="html-italic">w</span>/<span class="html-italic">w</span> of PPN2 (<b>a</b>) and area fraction <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <msubsup> <mo>Φ</mo> <mi>A</mi> <mrow> <mn>3</mn> <mo>/</mo> <mn>2</mn> </mrow> </msubsup> </mrow> </semantics></math>) versus weight fraction (<math display="inline"><semantics> <mrow> <msub> <mo>Φ</mo> <mi>w</mi> </msub> </mrow> </semantics></math>) (see the text for the precise definitions of <math display="inline"><semantics> <mrow> <msubsup> <mo>Φ</mo> <mi>A</mi> <mrow> <mn>3</mn> <mo>/</mo> <mn>2</mn> </mrow> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mo>Φ</mo> <mi>w</mi> </msub> </mrow> </semantics></math>) (<b>b</b>) for the series 6F#%.</p>
Full article ">Figure 3
<p>MMM 6F10% polished images: topography (<b>a</b>); phase contrast (<b>b</b>); amplitude obtained with the interleave scanning technique in negative-lift mode in FM (FM-Map) (<b>c</b>).</p>
Full article ">Figure 4
<p>6F15% with its surface partially dissolved with NMP: topography images (<b>a</b>,<b>c</b>); FM-Map (<b>b</b>,<b>d</b>).</p>
Full article ">Figure 5
<p>6F0% sample sequentially treated with Ar plasma (<b>a</b>–<b>f</b>). Times and power are indicated in the images.</p>
Full article ">Figure 6
<p>Samples treated with Ar plasma for 30 min at 10.2 W plus 12.5 h at 29.6 W. 6F0% (<b>a</b>), 6Ftr0% (<b>b</b>), tB0% (<b>c</b>), and 6F10% (<b>d</b>).</p>
Full article ">Figure 7
<p>The 6F10% samples treated with Ar plasma for 30 min at 10.2 W plus 12.5 h at 29.6 W. Top row: images with FM tip (70 kHz) of topography (<b>a</b>); phase (<b>b</b>); FM-Map (<b>c</b>). Botton row: images with tapping mode tip (300 kHz) of topography (<b>d</b>); phase contrast (<b>e</b>); phase contrast by using negative lift (<b>f</b>).</p>
Full article ">Figure 8
<p>The 6F0% sample treated with Ar plasma for 30 min at 10.2 W plus 12.5 h at 29.6 W, highlighting both a nanosphere and a surface network cell.</p>
Full article ">Figure 9
<p>He/N<sub>2</sub> selectivity versus He permeability for the studied samples. Void circles correspond to the plasma-untreated membranes and the filled circles to the samples treated with Ar plasma for 8.5 h at 29.6 W. Other literature data have been included [<a href="#B40-polymers-16-01397" class="html-bibr">40</a>].</p>
Full article ">Figure 10
<p>Scheme of the cross-section of an MMM showing the change in the effective thickness for the permeation before (<b>a</b>) and after (<b>b</b>) the plasma treatment.</p>
Full article ">Figure 11
<p>Pore size distribution obtained using NLDFT from CO<sub>2</sub> adsorption isotherms at 273.15 K for pure PPN2 and for membranes with and without filler: 6Ftr30% plasma-untreated and plasma-treated (<b>a</b>); 6Ftr0% plasma-untreated and plasma-treated (<b>b</b>) (Ar plasma treatment: 29.6 W for 8.5 h).</p>
Full article ">Figure 12
<p>Pore size distribution via NLDFT from N<sub>2</sub> isotherms (at 77 K) for samples with and without filler: 6Ftr30% plasma-treated and non-treated (<b>a</b>); 6Ftr0% plasma-treated and non-treated (<b>b</b>) (Ar plasma treatment: 29.6 W for 8.5 h).</p>
Full article ">Figure 13
<p>SEM images of MMM 6F20%: untreated (<b>a</b>); plasma-treated (<b>b</b>) (Ar plasma for 30 min at 10.2 W plus 12.5 h at 29.6 W).</p>
Full article ">Figure 14
<p>SEM surface images for plasma treated samples (Ar plasma for 30 min at 10.2 W plus 12.5 h at 29.6 W): tB0% (<b>a</b>); tB20% (<b>b</b>); 6Ftr20% (<b>c</b>). Red arrows point to some PPN2 particles.</p>
Full article ">Figure 15
<p>Magnification of <a href="#polymers-16-01397-f014" class="html-fig">Figure 14</a>a.</p>
Full article ">Figure 16
<p>Scheme of possible structural configurations of the MMMS (<b>A</b>–<b>F</b>) of this work and their possible equivalence (correspondence marked with green arrows) with the permeation model proposed by Moore and Koros [<a href="#B23-polymers-16-01397" class="html-bibr">23</a>].</p>
Full article ">Figure 17
<p>O<sub>2</sub>/N<sub>2</sub> selectivity versus O<sub>2</sub> permeability for the four families of polymers studied. The arrow marks the increase in PPN2 content from the pure polymer to the 30% of filler content.</p>
Full article ">
21 pages, 6337 KiB  
Article
An Improved Pedotransfer Function for Soil Hydrological Properties in New Zealand
by Stephen McNeill, Linda Lilburne, Shirley Vickers, Trevor Webb and Samuel Carrick
Appl. Sci. 2024, 14(10), 3997; https://doi.org/10.3390/app14103997 - 8 May 2024
Viewed by 919
Abstract
This paper describes a new pedotransfer function (PTF) for the soil water content of New Zealand soils at seven specific tensions (0, −5, −10, −20, −40, −100, −1500 kPa) using explanatory variables derived from the S-map soil mapping system. The model produces unbiased [...] Read more.
This paper describes a new pedotransfer function (PTF) for the soil water content of New Zealand soils at seven specific tensions (0, −5, −10, −20, −40, −100, −1500 kPa) using explanatory variables derived from the S-map soil mapping system. The model produces unbiased and physically plausible estimates of the response at each tension, as well as unbiased and physically plausible estimates of the response differences that define derived properties (e.g., macroporosity and total available water content). The PTF is a development of an earlier model using approximately double the number of sites compared with the earlier study, a change in fitting methodology to a semi-parametric GAM Beta response, and the inclusion of sample depth. The results show that the new model has resulted in significant improvements for the soil water content estimates and derived quantities using standard goodness-of-fit measures, based on validation data. A comparison with an international PTF using explanatory variables compatible with variables available from S-map (EUPTF2) suggests that the model is better for prediction of soil water content using the limited information available from the S-map system. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of points across New Zealand in the assembled NSDR dataset. One point is missing from this map since there were no spatial coordinates available.</p>
Full article ">Figure 2
<p>Measured versus modeled soil–water response [0, 1] for all tensions by soil order, as well as total available water content (TAW), using validation data. The error bars are the 95% prediction intervals, calculated by posterior simulation.</p>
Full article ">Figure 3
<p>Probability density of the response residuals (i.e., predicted minus measured values of the response) for training and validation data, for all tensions. TAW—total available water content.</p>
Full article ">Figure 4
<p>Plot of measured against predicted response using a Beta regression GAM with a smooth function for sample depth and carbon concentration, by composition class. TAW—total available water content.</p>
Full article ">Figure 5
<p>Contour plot of the estimated median total available water content (TAW) for organic horizons as a function of sample depth and carbon fraction for: (<b>a</b>) fibrous peaty soils; (<b>b</b>) humic peat composition soils. The markers are the data points used for model fitting.</p>
Full article ">Figure 6
<p>Plot of selected goodness-of-fit (GOF) measures for training and validation subsets of the data for each response, for non-organic soils. The dashed line corresponds to the value for the GOF measure for an ideal or perfect fit between the observed and predicted data. TAW—total available water content, MP—Macroporosity.</p>
Full article ">Figure 7
<p>Plot of the estimated median of total available water content (TAW) (left axis, solid line) and average accuracy (mean of the upper and lower 95% confidence values of TAW; right axis, dashed line) as a function of the sample depth. The shaded region in each case is plus-and-minus one standard error of the estimate.</p>
Full article ">Figure 8
<p>Measured against predicted total available water content (TAW) response for the EUPTF2 model. The predicted EUPTF2 TAW value was calculated from the difference between the EUPTF2 predicted responses at −10 and −1500 kPa. The long-dashed line is the linear fit between measured and modeled values, while the short-dashed line is the 1:1 line.</p>
Full article ">Figure 9
<p>Goodness-of-fit (GOF) measures for the 2018 logistic regression model from [<a href="#B1-applsci-14-03997" class="html-bibr">1</a>], the 2024 GAM model from this study, and the EUPTF V2 models [<a href="#B26-applsci-14-03997" class="html-bibr">26</a>], by measure. The preferred direction varies between measures; the dashed red line shows the ideal value for each of the measures.</p>
Full article ">
Back to TopTop