[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 (559)

Search Parameters:
Keywords = radar rainfall

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
29 pages, 9650 KiB  
Article
Seasonal Variations in the Rainfall Kinetic Energy Estimation and the Dual-Polarization Radar Quantitative Precipitation Estimation Under Different Rainfall Types in the Tianshan Mountains, China
by Yong Zeng, Lianmei Yang, Zepeng Tong, Yufei Jiang, Abuduwaili Abulikemu, Xinyu Lu and Xiaomeng Li
Remote Sens. 2024, 16(20), 3859; https://doi.org/10.3390/rs16203859 (registering DOI) - 17 Oct 2024
Abstract
Raindrop size distribution (DSD) has an essential effect on rainfall kinetic energy estimation (RKEE) and dual-polarization radar quantitative precipitation estimation (QPE); DSD is a key factor for establishing a dual-polarization radar QPE scheme and RKEE scheme, particularly in mountainous areas. To improve the [...] Read more.
Raindrop size distribution (DSD) has an essential effect on rainfall kinetic energy estimation (RKEE) and dual-polarization radar quantitative precipitation estimation (QPE); DSD is a key factor for establishing a dual-polarization radar QPE scheme and RKEE scheme, particularly in mountainous areas. To improve the understanding of seasonal DSD-based RKEE, dual-polarization radar QPE, and the impact of rainfall types and classification methods, we investigated RKEE schemes and dual-polarimetric radar QPE algorithms across seasons and rainfall types based on two classic classification methods (BR09 and BR03) and DSD data from a disdrometer in the Tianshan Mountains during 2020–2022. Two RKEE schemes were established: the rainfall kinetic energy flux–rain rate (KEtimeR) and the rainfall kinetic energy content–mass-weighted mean diameter (KEmmDm). Both showed seasonal variation, whether it was stratiform rainfall or convective rainfall, under BR03 and BR09. Both schemes had excellent performance, especially the KEmmDm relationship across seasons and rainfall types. In addition, four QPE schemes for dual-polarimetric radar—R(Kdp), R(Zh), R(Kdp,Zdr), and R(Zh,Zdr)—were established, and exhibited characteristics that varied with season and rainfall type. Overall, the performance of the single-parameter algorithms was inferior to that of the double-parameter algorithms, and the performance of the R(Zh) algorithm was inferior to that of the R(Kdp) algorithm. The results of this study show that it is necessary to consider different rainfall types and seasons, as well as classification methods of rainfall types, when applying RKEE and dual-polarization radar QPE. In this process, choosing a suitable estimator—KEtime(R), KEmm(Dm), R(Kdp), R(Zh), R(Kdp,Zdr), or R(Zh,Zdr)—is key to improving the accuracy of estimating the rainfall KE and R. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Topography (m) and location of the Tianshan Mountains, and (<b>b</b>) locations of Zhaosu (red dot) and Xinyuan (black dot; Zeng et al. [<a href="#B55-remotesensing-16-03859" class="html-bibr">55</a>]).</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) Topography (m) and location of the Tianshan Mountains, and (<b>b</b>) locations of Zhaosu (red dot) and Xinyuan (black dot; Zeng et al. [<a href="#B55-remotesensing-16-03859" class="html-bibr">55</a>]).</p>
Full article ">Figure 2
<p>Seasonal variations in the distributions of (<b>a</b>) <span class="html-italic">KE<sub>time</sub></span> and (<b>b</b>) <span class="html-italic">KE<sub>mm</sub></span> at Zhaosu.</p>
Full article ">Figure 3
<p>Scatterplots of <span class="html-italic">KE<sub>time</sub></span> vs. <span class="html-italic">R</span> for the entire data and the fitted <span class="html-italic">KE<sub>time</sub></span>–<span class="html-italic">R</span> relationship across seasons at Zhaosu. Dashed lines represent the <span class="html-italic">KE<sub>time</sub></span>–<span class="html-italic">R</span> relationship reported by Zeng et al. [<a href="#B55-remotesensing-16-03859" class="html-bibr">55</a>], Seela et al. [<a href="#B83-remotesensing-16-03859" class="html-bibr">83</a>], and Wu et al. [<a href="#B36-remotesensing-16-03859" class="html-bibr">36</a>].</p>
Full article ">Figure 4
<p>Scatterplots of <span class="html-italic">KE<sub>mm</sub></span> vs. <span class="html-italic">D<sub>m</sub></span> for the entire data and the seasonal variation in fitted <span class="html-italic">KE<sub>mm</sub></span>–<span class="html-italic">D<sub>m</sub></span> at Zhaosu. Dashed lines represent the <span class="html-italic">KE<sub>mm</sub></span>–<span class="html-italic">D<sub>m</sub></span> relationship reported by Zeng et al. [<a href="#B55-remotesensing-16-03859" class="html-bibr">55</a>] and Seela et al. [<a href="#B83-remotesensing-16-03859" class="html-bibr">83</a>].</p>
Full article ">Figure 5
<p>Scatterplot of estimated <span class="html-italic">KE<sub>time</sub></span> from RKEE schemes versus <span class="html-italic">KE<sub>time</sub></span> calculated from DSD for (<b>a</b>) the entire data, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) fall at Zhaosu. Scatterplot of estimated <span class="html-italic">KE<sub>mm</sub></span> from RKEE schemes versus the <span class="html-italic">KE<sub>mm</sub></span> calculated from DSD for (<b>e</b>) the entire data, (<b>f</b>) spring, (<b>g</b>) summer, and (<b>h</b>) fall at Zhaosu in Tianshan Mountains. Black dashed lines represent the 1:1 relationship.</p>
Full article ">Figure 6
<p>Violin plots of seasonal variations in <span class="html-italic">KE<sub>time</sub></span> under (<b>a</b>) BR09_S, (<b>c</b>) BR09_C, (<b>e</b>) BR03_S, and (<b>g</b>) BR03_C, and violin plots of seasonal variations in <span class="html-italic">KE<sub>mm</sub></span> under (<b>b</b>) BR09_S, (<b>d</b>) BR09_C, (<b>f</b>) BR03_S, and (<b>h</b>) BR03_C at Zhaosu.</p>
Full article ">Figure 7
<p>Scatterplots of <span class="html-italic">KE<sub>time</sub></span> vs. <span class="html-italic">R</span> for the entire data and the seasonal variation of the fitted <span class="html-italic">KE<sub>time</sub></span>–<span class="html-italic">R</span> relationship at Zhaosu under (<b>a</b>) BR09_S, (<b>b</b>) BR09_C, (<b>c</b>) BR03_S, and (<b>d</b>) BR03_C.</p>
Full article ">Figure 8
<p>Scatterplot of estimated <span class="html-italic">KE<sub>time</sub></span> from RKEE schemes versus <span class="html-italic">KE<sub>time</sub></span> calculated from DSD for (<b>a</b>) the entire data, (<b>e</b>) spring, (<b>i</b>) summer, and (<b>m</b>) fall under BR09_S; those for (<b>b</b>) the entire data, (<b>f</b>) spring, (<b>j</b>) summer, and (<b>n</b>) fall under BR09_C; those for (<b>c</b>) the entire data, (<b>g</b>) spring, (<b>k</b>) summer, and (<b>o</b>) fall under BR03_S; and those for (<b>d</b>) the entire data, (<b>h</b>) spring, and (<b>l</b>) summer under BR03_C at Zhaosu. Black dashed lines represent the 1:1 relationship.</p>
Full article ">Figure 9
<p>Scatterplots of <span class="html-italic">KE<sub>mm</sub></span> vs. <span class="html-italic">D<sub>m</sub></span> for the entire data and the fitted <span class="html-italic">KE<sub>mm</sub></span>–<span class="html-italic">D<sub>m</sub></span> relationship across seasons at Zhaosu under (<b>a</b>) BR09_S, (<b>b</b>) BR09_C, (<b>c</b>) BR03_S, and (<b>d</b>) BR03_C.</p>
Full article ">Figure 10
<p>Scatterplot of estimated <span class="html-italic">KE<sub>mm</sub></span> from RKEE schemes versus <span class="html-italic">KE<sub>mm</sub></span> calculated from DSD for (<b>a</b>) the entire data, (<b>e</b>) spring, (<b>i</b>) summer, and (<b>m</b>) fall under BR09_S; for (<b>b</b>) the entire data, (<b>f</b>) spring, (<b>j</b>) summer, and (<b>n</b>) fall under BR09_C; for (<b>c</b>) the entire data, (<b>g</b>) spring, (<b>k</b>) summer, and (<b>o</b>) fall under BR03_S; and for (<b>d</b>) the entire data, (<b>h</b>) spring, and (<b>l</b>) summer under BR03_C at Zhaosu. Black dashed lines represent the 1:1 relationship.</p>
Full article ">Figure 11
<p>Seasonal variations in the distributions of (<b>a</b>) <span class="html-italic">Z<sub>h</sub></span>, (<b>b</b>) <span class="html-italic">Z<sub>dr</sub></span>, and (<b>c</b>) <span class="html-italic">K<sub>dp</sub></span> at Zhaosu.</p>
Full article ">Figure 12
<p>Scatterplot of estimated <span class="html-italic">R</span> based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span>) algorithm for (<b>a</b>) the entire data, (<b>e</b>) spring, (<b>i</b>) summer, and (<b>m</b>) fall; estimated <span class="html-italic">R</span> based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span>) algorithm for (<b>b</b>) the entire data, (<b>f</b>) spring, (<b>j</b>) summer, and (<b>n</b>) fall; estimated <span class="html-italic">R</span> based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm for (<b>c</b>) the entire data, (<b>g</b>) spring, (<b>k</b>) summer, and (<b>o</b>) fall; and estimated <span class="html-italic">R</span> based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm for (<b>d</b>) the entire data, (<b>h</b>) spring, (<b>l</b>) summer, and (<b>p</b>) fall versus calculated <span class="html-italic">R</span> according to Equation (4) at Zhaosu. Black dashed lines represent the 1:1 relationship.</p>
Full article ">Figure 13
<p>Seasonal variations in the distributions of <span class="html-italic">Z<sub>h</sub></span> under (<b>a</b>) BR09_S, (<b>d</b>) BR09_C, (<b>g</b>) BR03_S, and (<b>j</b>) BR03_C; those of <span class="html-italic">Z<sub>dr</sub></span> under (<b>b</b>) BR09_S, (<b>e</b>) BR09_C, (<b>h</b>) BR03_S, and (<b>k</b>) BR03_C; and those of <span class="html-italic">K<sub>dp</sub></span> under (<b>c</b>) BR09_S, (<b>f</b>) BR09_C, (<b>i</b>) BR03_S, and (<b>l</b>) BR03_C at Zhaosu.</p>
Full article ">Figure 14
<p>Scatterplot of estimated <span class="html-italic">R</span> based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span>) algorithm during BR09_S for (<b>a</b>) the entire data, (<b>e</b>) spring, (<b>i</b>) summer, and (<b>m</b>) fall; that based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span>) algorithm during BR09_S for (<b>b</b>) the entire data, (<b>f</b>) spring, (<b>j</b>) summer, and (<b>n</b>) fall; that based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm during BR09_S for (<b>c</b>) the entire data, (<b>g</b>) spring, (<b>k</b>) summer, and (<b>o</b>) fall; and that based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm during BR09_S for (<b>d</b>) the entire data, (<b>h</b>) spring, (<b>l</b>) summer, and (<b>p</b>) fall versus calculated <span class="html-italic">R</span> according to Equation (4) at Zhaosu. Black dashed lines represent the 1:1 relationship.</p>
Full article ">Figure 15
<p>Scatterplot of estimated <span class="html-italic">R</span> based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span>) algorithm during BR09_C for (<b>a</b>) the entire data, (<b>e</b>) spring, (<b>i</b>) summer, and (<b>m</b>) fall; that based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span>) algorithm during BR09_C for (<b>b</b>) the entire data, (<b>f</b>) spring, (<b>j</b>) summer, and (<b>n</b>) fall; that based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm during BR09_C for (<b>c</b>) the entire data, (<b>g</b>) spring, (<b>k</b>) summer, and (<b>o</b>) fall; and that based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm during BR09_C for (<b>d</b>) the entire data, (<b>h</b>) spring, (<b>l</b>) summer, and (<b>p</b>) fall versus calculated <span class="html-italic">R</span> according to Equation (4) at Zhaosu. Black dashed lines represent the 1:1 relationship.</p>
Full article ">Figure 16
<p>Scatterplot of estimated <span class="html-italic">R</span> based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span>) algorithm during BR03_S for (<b>a</b>) the entire data, (<b>e</b>) spring, (<b>i</b>) summer, and (<b>m</b>) fall; that based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span>) algorithm during BR03_S for (<b>b</b>) the entire data, (<b>f</b>) spring, (<b>j</b>) summer, and (<b>n</b>) fall; that based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm during BR03_S for (<b>c</b>) the entire data, (<b>g</b>) spring, (<b>k</b>) summer, and (<b>o</b>) fall; and that based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm during BR03_S for (<b>d</b>) the entire data, (<b>h</b>) spring, (<b>l</b>) summer, and (<b>p</b>) fall versus calculated <span class="html-italic">R</span> according to Equation (4) at Zhaosu. Black dashed lines represent the 1:1 relationship.</p>
Full article ">Figure 17
<p>Scatterplot of estimated <span class="html-italic">R</span> based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span>) algorithm during BR03_C for (<b>a</b>) the entire data, (<b>e</b>) spring, and (<b>i</b>) summer; that based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span>) algorithm during BR03_C for (<b>b</b>) the entire data, (<b>f</b>) spring, and (<b>j</b>) summer; that based on the <span class="html-italic">R</span>(<span class="html-italic">Z<sub>h</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm during BR03_C for (<b>c</b>) the entire data, (<b>g</b>) spring, and (<b>k</b>) summer; and that based on the <span class="html-italic">R</span>(<span class="html-italic">K<sub>dp</sub></span><sub>,</sub><span class="html-italic">Z<sub>dr</sub></span>) algorithm during BR03_C for (<b>d</b>) the entire data, (<b>h</b>) spring, and (<b>l</b>) summer versus calculated <span class="html-italic">R</span> according to Equation (4) at Zhaosu. Black dashed lines represent the 1:1 relationship.</p>
Full article ">
20 pages, 11684 KiB  
Article
Development of a Storm-Tracking Algorithm for the Analysis of Radar Rainfall Patterns in Athens, Greece
by Apollon Bournas and Evangelos Baltas
Water 2024, 16(20), 2905; https://doi.org/10.3390/w16202905 - 12 Oct 2024
Viewed by 522
Abstract
This research work focuses on the development and application of a storm-tracking algorithm for identifying and tracking storm cells. The algorithm first identifies storm cells on the basis of reflectivity thresholds and then matches the cells in the tracking procedure on the basis [...] Read more.
This research work focuses on the development and application of a storm-tracking algorithm for identifying and tracking storm cells. The algorithm first identifies storm cells on the basis of reflectivity thresholds and then matches the cells in the tracking procedure on the basis of their geometrical characteristics and the distance within the weather radar image. A sensitivity analysis was performed to evaluate the preferable thresholds for each case and test the algorithm’s ability to perform in different time step resolutions. Following this, we applied the algorithm to 54 rainfall events recorded by the National Technical University X-Band weather radar, the rainscanner system, from 2018 to 2023 in the Attica region of Greece. Testing of the algorithm demonstrated its efficiency in tracking storm cells over various time intervals and reflecting changes such as merging or dissipation. The results reveal the predominant southwest-to-east storm directions in 40% of cases examined, followed by northwest-to-east and south-to-north patterns. Additionally, stratiform storms showed slower north-to-west trajectories, while convective storms exhibited faster west-to-east movement. These findings provide valuable insights into storm behavior in Athens and highlight the algorithm’s potential for integration into nowcasting systems, particularly for flood early warning systems. Full article
Show Figures

Figure 1

Figure 1
<p>The study area, the Attica region, along with the beam blockage area.</p>
Full article ">Figure 2
<p>The rainscanner pre-process data correction procedure.</p>
Full article ">Figure 3
<p>The storm-tracking algorithm flow chart.</p>
Full article ">Figure 4
<p>Monthly distribution of recorded events used in the study.</p>
Full article ">Figure 5
<p>The results of the cell algorithm are shown when a different reflectivity threshold is used; 25-, 30-, and 35-dBZ are shown in each row, while in each column when a polygon or fitting ellipse is used to form the boundaries. In each frame, the red shapes illustrate the boundaries of the cells, the star symbol the cell’s centroid’s location and the number the ID of each cell.</p>
Full article ">Figure 6
<p>The matchmaking procedure between six 2-min reflectivity fields time frames with a 2-min step. In each frame, the red shapes illustrate the boundaries of the cells, while the star and cross symbols illustrate the previous and current frame cell’s centroid’s location, respectively.</p>
Full article ">Figure 7
<p>The matchmaking procedure between six 10-min reflectivity fields time frames with a 2-min step. In each frame, the red shapes illustrate the boundaries of the cells, while the star and cross symbols illustrate the previous and current frame cell’s centroid’s location, respectively.</p>
Full article ">Figure 8
<p>The matchmaking procedure between six 10-min reflectivity fields time frames with a 10-min step. In each frame, the red shapes illustrate the boundaries of the cells, while the star and cross symbols illustrate the previous and current frame cell’s centroid’s location, respectively.</p>
Full article ">Figure 8 Cont.
<p>The matchmaking procedure between six 10-min reflectivity fields time frames with a 10-min step. In each frame, the red shapes illustrate the boundaries of the cells, while the star and cross symbols illustrate the previous and current frame cell’s centroid’s location, respectively.</p>
Full article ">Figure 9
<p>Storm cells of events 31 and 50, for two reflectivity thresholds, illustrated with blue polygons; first column 35 dBZ, second column 40 dBZ. The triangle symbol illustrates the Rainscanner location, wher eas the dashed circles an increasing 10 km distance from its location.</p>
Full article ">Figure 9 Cont.
<p>Storm cells of events 31 and 50, for two reflectivity thresholds, illustrated with blue polygons; first column 35 dBZ, second column 40 dBZ. The triangle symbol illustrates the Rainscanner location, wher eas the dashed circles an increasing 10 km distance from its location.</p>
Full article ">Figure 10
<p>Centroids and areas covered by storm cells with a reflectivity threshold &gt;40 dBZ. The triangle symbol illustrates the Rainscanner location, whereas the dashed circles an increasing 10 km distance from its location. The dashed arrow illustrates the main trajectory of the storm.</p>
Full article ">Figure 10 Cont.
<p>Centroids and areas covered by storm cells with a reflectivity threshold &gt;40 dBZ. The triangle symbol illustrates the Rainscanner location, whereas the dashed circles an increasing 10 km distance from its location. The dashed arrow illustrates the main trajectory of the storm.</p>
Full article ">Figure 11
<p>Main directions of storm cells in the Attica Regio: red lines are the most reoccurring, followed by blue, yellow, and purple, according to the attached chart. The triangle symbol illustrates the Rainscanner location, whereas the dashed circles an increasing 10 km distance from its location.</p>
Full article ">
15 pages, 2753 KiB  
Article
Assessing Soil Physical Quality in a Layered Agricultural Soil: A Comprehensive Approach Using Infiltration Experiments and Time-Lapse Ground-Penetrating Radar Surveys
by Simone Di Prima, Gersende Fernandes, Maria Burguet, Ludmila Ribeiro Roder, Vittoria Giannini, Filippo Giadrossich, Laurent Lassabatere and Alessandro Comegna
Appl. Sci. 2024, 14(20), 9268; https://doi.org/10.3390/app14209268 - 11 Oct 2024
Viewed by 505
Abstract
Time-lapse ground-penetrating radar (GPR) surveys, combined with automated infiltration experiments, provide a non-invasive approach for investigating the distribution of infiltrated water within the soil medium and creating three-dimensional images of the wetting bulb. This study developed and validated an experimental protocol aimed at [...] Read more.
Time-lapse ground-penetrating radar (GPR) surveys, combined with automated infiltration experiments, provide a non-invasive approach for investigating the distribution of infiltrated water within the soil medium and creating three-dimensional images of the wetting bulb. This study developed and validated an experimental protocol aimed at quantifying and visualizing water distribution fluxes in layered soils under both unsaturated and saturated conditions. The 3D images of the wetting bulb significantly enhanced the interpretation of infiltration data, enabling a detailed analysis of water movement through the layered system. We used the infiltrometer data and the Beerkan Estimation of Soil Transfer parameters (BEST) method to determine soil capacitive indicators and evaluate the physical quality of the upper soil layer. The field survey involved conducting time-lapse GPR surveys alongside infiltration experiments between GPR repetitions. These experiments included both tension and ponding tests, designed to sequentially activate the soil matrix and the full pore network. The results showed that the soil under study exhibited significant soil aeration and macroporosity (represented by AC and pMAC), while indicators related to microporosity (such as PAWC and RFC) were notably low. The RFC value of 0.55 m3 m−3 indicated the soil’s limited capacity to retain water relative to its total pore volume. The PAWC value of 0.10 m3 m−3 indicated a scarcity of micropores ranging from 0.2 to 30 μm in diameter, which typically hold water accessible to plant roots within the total porosity. The saturated soil hydraulic conductivity, Ks, values ranged from 192.2 to 1031.0 mm h−1, with a mean of 424.4 mm h−1, which was 7.9 times higher than the corresponding unsaturated hydraulic conductivity measured at a pressure head of h = −30 mm (K−30). The results indicated that the upper soil layer supports root proliferation and effectively drains excess water to the underlying limestone layer. However, this layer has limited capacity to store and supply water to plant roots and acts as a restrictive barrier, promoting non-uniform downward water movement, as revealed by the 3D GPR images. The observed difference in hydraulic conductivity between the two layers suggests that surface ponding and overland flow are generated through a saturation excess mechanism. Water percolating through the soil can accumulate above the limestone layer, creating a shallow perched water table. During extreme rainfall events, this water table may rise, leading to the complete saturation of the soil profile. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart outlining the process to generate a 3D image of the wetting bulb. The arrow indicates the funneling flow path through the limestone layer.</p>
Full article ">Figure 2
<p>Three-dimensional representations of the wetting zones obtained from ground-penetrating radar surveys conducted before and after wetting, during (<b>a</b>) tension and (<b>e</b>) ponding infiltrometer experiments at the Ottava site. Panels (<b>b</b>,<b>f</b>) illustrate horizontal cross-sections taken from the 3D models at a depth of −0.1m from the soil surface. Panels (<b>c</b>,<b>g</b>) present vertical cross-sections oriented north–south with a view to the east, while panels (<b>d</b>,<b>h</b>) show vertical cross-sections oriented west–east within a view to the north. The red arrows highlight the detected flow channeling through the limestone layer (see <a href="#applsci-14-09268-f001" class="html-fig">Figure 1</a> for reference).</p>
Full article ">Figure 3
<p>Example of the procedure adopted for detecting flow impedance owing to the hydraulic resistance exerted by the underlying limestone layer. (<b>a</b>): Entire cumulative infiltration curve [<span class="html-italic">I</span>(<span class="html-italic">t</span>) vs. <span class="html-italic">t</span>]. (<b>b</b>): Data linearized according to the cumulative linearization (CL, Smiles and Knight, 1976) method (<span class="html-italic">I</span>√<span class="html-italic">t</span> vs. √<span class="html-italic">t</span>). The abscissa (√<span class="html-italic">t</span>) of the intersection point of the two straight lines splits the infiltration data into two subsets. (<b>c</b>): Cumulative infiltration data representative of the first stage when water infiltrates into the upper layer.</p>
Full article ">Figure 4
<p>θ<span class="html-italic"><sub>PWP</sub></span> [m<sup>3</sup> m<sup>−3</sup>] is the permanent wilting point soil water content, corresponding to <span class="html-italic">h</span> = −150 m. θ<span class="html-italic"><sub>FC</sub></span> [m<sup>3</sup> m<sup>−3</sup>] is the field capacity (gravity drained) soil water content, corresponding to <span class="html-italic">h</span> = −1 m. θ<span class="html-italic"><sub>m</sub></span> [m<sup>3</sup> m<sup>−3</sup>] is the saturated volumetric water content of the soil matrix, corresponding to <span class="html-italic">h</span> = −0.1 m. θ<span class="html-italic"><sub>TI</sub></span> [m<sup>3</sup> m<sup>−3</sup>] is the final volumetric water content at the end of the TI test (corresponding to <span class="html-italic">h</span> = −0.03 m), θ<span class="html-italic"><sub>s</sub></span> [m<sup>3</sup> m<sup>−3</sup>] is the saturated volumetric water content. <span class="html-italic">AC</span> [m<sup>3</sup> m<sup>−3</sup>] is the air capacity. <span class="html-italic">PAWC</span> [m<sup>3</sup> m<sup>−3</sup>] is the plant-available water capacity. <span class="html-italic">RFC</span> [−] is the relative field capacity. <span class="html-italic">p<sub>MAC</sub></span> [m<sup>3</sup> m<sup>−3</sup>] is the soil macroporosity. <sup>†</sup> Water content values determined from wet soil samples collected after the tension (θ<span class="html-italic"><sub>TI</sub></span>) and Beerkan (θ<span class="html-italic"><sub>s</sub></span>) infiltration tests.</p>
Full article ">
19 pages, 5199 KiB  
Article
Enhanced Precipitation Nowcasting via Temporal Correlation Attention Mechanism and Innovative Jump Connection Strategy
by Wenbin Yu, Daoyong Fu, Chengjun Zhang, Yadang Chen, Alex X. Liu and Jingjing An
Remote Sens. 2024, 16(20), 3757; https://doi.org/10.3390/rs16203757 - 10 Oct 2024
Viewed by 314
Abstract
This study advances the precision and efficiency of precipitation nowcasting, particularly under extreme weather conditions. Traditional forecasting methods struggle with precision, spatial feature generalization, and recognizing long-range spatial correlations, challenges that intensify during extreme weather events. The Enhanced Temporal Correlation Jump Prediction Network [...] Read more.
This study advances the precision and efficiency of precipitation nowcasting, particularly under extreme weather conditions. Traditional forecasting methods struggle with precision, spatial feature generalization, and recognizing long-range spatial correlations, challenges that intensify during extreme weather events. The Enhanced Temporal Correlation Jump Prediction Network (ETCJ-PredNet) introduces a novel attention mechanism that optimally leverages spatiotemporal data correlations. This model scrutinizes and encodes information from previous frames, enhancing predictions of high-intensity radar echoes. Additionally, ETCJ-PredNet addresses the issue of gradient vanishing through an innovative jump connection strategy. Comparative experiments on the Moving Modified National Institute of Standards and Technology (Moving-MNIST) and Hong Kong Observatory Dataset Number 7 (HKO-7) validate that ETCJ-PredNet outperforms existing models, particularly under extreme precipitation conditions. Detailed evaluations using Critical Success Index (CSI), Heidke Skill Score (HSS), Probability of Detection (POD), and False Alarm Ratio (FAR) across various rainfall intensities further underscore its superior predictive capabilities, especially as rainfall intensity exceeds 30 dbz,40 dbz, and 50 dbz. These results confirm ETCJ-PredNet’s robustness and utility in real-time extreme weather forecasting. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Advanced schematic diagram of the PredRNN-V2 architecture [<a href="#B20-remotesensing-16-03757" class="html-bibr">20</a>].</p>
Full article ">Figure 2
<p>Architectural framework diagram with jump connection strategy.</p>
Full article ">Figure 3
<p>On the left is the Schematic Diagram of the Time Memory Flow Architecture in the PredRNN Model, and on the right is the Improved Schematic Diagram of the Time Memory Flow Architecture with Jump Connections Introduced.</p>
Full article ">Figure 4
<p>(<b>a</b>) Improved architecture incorporating temporal correlation attention mechanism. (<b>b</b>) Traditional SDPA architecture and proposed temporal correlation attention architecture.</p>
Full article ">Figure 5
<p>Display of prediction results on the Moving-MNIST test dataset.</p>
Full article ">Figure 6
<p>Comparison of MSE, LPIPS, SSIM, and PSNR metrics under different jump connection strategies.</p>
Full article ">Figure 7
<p>On the left are the original training loss curves for different models, and on the right are the smoothed training loss curves with smoothing techniques applied.</p>
Full article ">Figure 8
<p>Performance trends of various models over time steps on the HKO-7 dataset.</p>
Full article ">Figure 9
<p>Examples of predictions on the radar echo test set, generating 20 future frames from 10 past observations, demonstrating the prediction of echo growth.</p>
Full article ">Figure 10
<p>Examples of predictions on the radar echo test set, generating 20 future frames from 10 past observations, demonstrating the prediction of precipitation decay.</p>
Full article ">
29 pages, 11100 KiB  
Article
Assessing the Impact of Rainfall Inputs on Short-Term Flood Simulation with Cell2Flood: A Case Study of the Waryong Reservoir Basin
by Hyunjun Kim, Dae-Sik Kim, Won-Ho Nam and Min-Won Jang
Hydrology 2024, 11(10), 162; https://doi.org/10.3390/hydrology11100162 - 2 Oct 2024
Viewed by 503
Abstract
This study explored the impacts of various rainfall input types on short-term runoff simulations using the Cell2Flood model in the Waryong Reservoir Basin, South Korea. Six types of rainfall data were assessed: on-site gauge measurements, spatially interpolated data from 39 Automated Synoptic Observing [...] Read more.
This study explored the impacts of various rainfall input types on short-term runoff simulations using the Cell2Flood model in the Waryong Reservoir Basin, South Korea. Six types of rainfall data were assessed: on-site gauge measurements, spatially interpolated data from 39 Automated Synoptic Observing System (ASOS) and 117 Automatic Weather System (AWS) stations using inverse distance weighting (IDW), and Hybrid Surface Rainfall (HSR) data from the Korea Meteorological Administration. The choice of rainfall input significantly affected model accuracy across the three rainfall events. The point-gauged ASOS (P-ASOS) data demonstrated the highest reliability in capturing the observed rainfall patterns, with Pearson’s r values of up to 0.84, whereas the radar-derived HSR data had the lowest correlations (Pearson’s r below 0.2), highlighting substantial discrepancies. For runoff simulation, the P-ASOS and ASOS-AWS combined interpolated dataset (R-AWS) achieved relatively accurate predictions, with P-ASOS and R-AWS exhibiting Normalized Peak Error (NPE) values of approximately 0.03 and Peak Time Error (PTE) within 20 min. In contrast, the HSR data produced large errors, with NPE up to 4.66 and PTE deviations exceeding 200 min, indicating poor temporal accuracy. Although input-specific calibration improved performance, significant errors persisted because of the inherent uncertainty of rainfall data. These findings underscore the importance of selecting and calibrating appropriate rainfall inputs to enhance the reliability of short-term flood modeling, particularly in ungauged and data-sparse basins. Full article
Show Figures

Figure 1

Figure 1
<p>Study site located in southern Gyeongsangnam-do, with observation spots with one radar flow level gauge and one AWS.</p>
Full article ">Figure 2
<p>Screen interface of Cell2Flood model showing its graphical user interface and auto-calibration module.</p>
Full article ">Figure 3
<p>Spatial input data for Cell2Flood, which were converted into an ASCII file with 500 m spatial resolution as the input data for Cell2Flood.</p>
Full article ">Figure 4
<p>Geographical distribution of rain gauge (ASOS and AWS) stations around the study site.</p>
Full article ">Figure 5
<p>Comparison of accumulated rainfall for each rainfall input by events.</p>
Full article ">Figure 6
<p>Change in accumulated rainfall in 1 h intervals of each event by rainfall inputs.</p>
Full article ">Figure 7
<p>Comparison of 1 h maximum rainfall intensity for each rainfall input by events.</p>
Full article ">Figure 8
<p>Comparison of the spatial distribution of 10-min accumulated rainfall for each rainfall input across different events. Event (<b>I</b>) 2023-06-27 22:30, (<b>II</b>) 2023-08-10 07:10, and (<b>III</b>) 2024-06-29 19:00, respectively, and (<b>a</b>–<b>c</b>) corresponding to R-ASOS, R-AWS, R-RADAR.</p>
Full article ">Figure 9
<p>Results of rainfall–runoff simulations for Event I by rainfall inputs.</p>
Full article ">Figure 9 Cont.
<p>Results of rainfall–runoff simulations for Event I by rainfall inputs.</p>
Full article ">Figure 10
<p>Results of rainfall–runoff simulations for Event II by rainfall inputs.</p>
Full article ">Figure 10 Cont.
<p>Results of rainfall–runoff simulations for Event II by rainfall inputs.</p>
Full article ">Figure 11
<p>Results of rainfall–runoff simulations for Event III by rainfall inputs.</p>
Full article ">Figure 11 Cont.
<p>Results of rainfall–runoff simulations for Event III by rainfall inputs.</p>
Full article ">Figure 12
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for P-ASOS.</p>
Full article ">Figure 13
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for P-AWS.</p>
Full article ">Figure 14
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for R-ASOS.</p>
Full article ">Figure 15
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for R-AWS.</p>
Full article ">Figure 16
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for R-RADAR.</p>
Full article ">
20 pages, 4521 KiB  
Article
Optimizing the Activation of WWTP Wet-Weather Operation Using Radar-Based Flow and Volume Forecasting with the Relative Economic Value (REV) Approach
by Vianney Courdent, Thomas Munk-Nielsen and Peter Steen Mikkelsen
Water 2024, 16(19), 2806; https://doi.org/10.3390/w16192806 - 2 Oct 2024
Viewed by 390
Abstract
Wastewater treatment plants (WWTPs) connected to combined sewer systems must cope with high flows during wet-weather conditions, often leading to bypass and thus pollution of water bodies. Radar rainfall forecasts coupled with a rainfall-runoff model provides flow and volume forecasts that can be [...] Read more.
Wastewater treatment plants (WWTPs) connected to combined sewer systems must cope with high flows during wet-weather conditions, often leading to bypass and thus pollution of water bodies. Radar rainfall forecasts coupled with a rainfall-runoff model provides flow and volume forecasts that can be used for deciding when to switch from normal to wet-weather operation, which temporarily allows for higher inflow. However, forecasts are by definition uncertain and may lead to potential mismanagement, e.g., false alarms and misses. Our study focused on two years of operational data from the Damhuså sewer catchment and WWTP. We used the Relative Economic Value (REV) framework to optimize the control parameters of a baseline control strategy (thresholds on flow measurements and radar flow prognosis) and to test new control strategies based on volume instead of flow thresholds. We investigated two situations with different objective functions, considering higher negative impact from misses than false alarms and vice versa, and obtained in both cases a reduction of the rate of false alarms, higher flow thresholds and lower bypass compared to the baseline control. We also assess a new control strategy that employs thresholds of predicted accumulated volume instead of predicted flow and achieved even better results. Full article
(This article belongs to the Section Wastewater Treatment and Reuse)
Show Figures

Figure 1

Figure 1
<p>Benefit from use of a flow forecast for wet-weather control switching, which leads to an avoided discharge (bypass) of untreated wastewater. Based on [<a href="#B7-water-16-02806" class="html-bibr">7</a>].</p>
Full article ">Figure 2
<p>The Damhuså catchment (<b>top</b>) and a process diagram of the Damhuså WWTP (<b>bottom</b>).</p>
Full article ">Figure 3
<p>Main loads and concentrations during dry weather operation (<b>a</b>) and wet weather with the two operation modes: conventional wet-weather operation (<b>b</b>) and ATS operation (<b>c</b>) (inspired from [<a href="#B7-water-16-02806" class="html-bibr">7</a>].</p>
Full article ">Figure 4
<p>Baseline control scheme for the ATS switch at the Damhuså WWTP (June 2015–June 2017), based on three different inputs, (<b>A</b>) the measured inflow at the WWTP, (<b>B</b>) the measured flow at Dæmning upstream in the drainage system, and (<b>C</b>) the flow prognosis at the WWTP using radar extrapolation data.</p>
Full article ">Figure 5
<p>Examples of ATS control switch for two events in August 2015 and June 2016, based on (see <a href="#water-16-02806-f004" class="html-fig">Figure 4</a>) flow measurements at the WWTP (A) and the upstream Dæmning location (B) and on radar flow prognosis (C). The maximal hydraulic capacity to the biological treatment varies under different conditions: (a) dry weather, (b) preparation of the ATS operation, (c) ATS operation, (d) critical sludge blanket level in secondary settlers and the wastewater bypassed (cross hatched).</p>
Full article ">Figure 6
<p>Example of volume based approach (dotted rectangles in purple and green) compared to the flow threshold approach (dotted rectangles in red), for three examples showing that use of a flow threshold can lead to both (<b>a</b>) false alarms and (<b>b</b>) hits, and that ((<b>c</b>), compared with (<b>b</b>)) volume forecasts can be made with different coupled volume-duration. “On” means that in a given situation ATS would be activated, and “off” that the ATS would not be activated. The time of activation of the ATS is represented by the time <span class="html-italic">t<sub>1</sub></span> and <span class="html-italic">t<sub>2</sub></span>.</p>
Full article ">Figure 7
<p>Average flow-duration criterion to start the ATS operation based on the radar flow prog-nosis.</p>
Full article ">Figure 8
<p>Histograms of the ATS event duration for the different control strategies outlined in <a href="#water-16-02806-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 9
<p><span class="html-italic">REV</span> response surface (*) for the current ATS control (FOR-2), with <span class="html-italic">k</span> and <span class="html-italic">α</span> as independent parameters, for the 3 different REF. Notice that the <span class="html-italic">k</span> and <span class="html-italic">α</span> axis are reversed for better visibility of the 3D plots.</p>
Full article ">Figure 10
<p>Cross-section of the <span class="html-italic">REV</span> surface response for high impact of misses with <span class="html-italic">k</span> = 0.2 (<b>a</b>–<b>c</b>) and high impact of false alarms with <span class="html-italic">k</span> = 0.8 (<b>d</b>–<b>f</b>).</p>
Full article ">
18 pages, 3143 KiB  
Article
Estimating Rainfall Intensity Using an Image-Based Convolutional Neural Network Inversion Technique for Potential Crowdsourcing Applications in Urban Areas
by Youssef Shalaby, Mohammed I. I. Alkhatib, Amin Talei, Tak Kwin Chang, Ming Fai Chow and Valentijn R. N. Pauwels
Big Data Cogn. Comput. 2024, 8(10), 126; https://doi.org/10.3390/bdcc8100126 - 29 Sep 2024
Viewed by 510
Abstract
High-quality rainfall data are essential in many water management problems, including stormwater management, water resources management, and more. Due to the high spatial–temporal variations, rainfall measurement could be challenging and costly, especially in urban areas. This could be even more challenging in tropical [...] Read more.
High-quality rainfall data are essential in many water management problems, including stormwater management, water resources management, and more. Due to the high spatial–temporal variations, rainfall measurement could be challenging and costly, especially in urban areas. This could be even more challenging in tropical regions with their typical short-duration and high-intensity rainfall events, as some of the undeveloped or developing countries in those regions lack a dense rain gauge network and have limited resources to use radar and satellite readings. Thus, exploring alternative rainfall estimation methods could be helpful to back up some shortcomings. Recently, a few studies have examined the utilisation of citizen science methods to collect rainfall data as a complement to the existing rain gauge networks. However, these attempts are in the early stages, and limited works have been published on improving the quality of such data. Therefore, this study focuses on image-based rainfall estimation with potential usage in citizen science. For this, a novel convolutional neural network (CNN) model is developed to predict rainfall intensity by processing the images captured by citizens (e.g., by smartphones or security cameras) in an urban area. The developed model is merely a complementary sensing tool (e.g., better spatial coverage) to the existing rain gauge network in an urban area and is not meant to replace it. This study also presents one of the most extensive datasets of rain image data ever published in the literature. The estimated rainfall data by the proposed CNN model of this study using images captured by surveillance cameras and smartphone cameras are compared with observed rainfall by a weather station and exhibit strong R2 values of 0.955 and 0.840, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>Locations of rainfall images captured using surveillance cameras and smartphones during rain events near the Monash University campus in Malaysia.</p>
Full article ">Figure 2
<p>Sample images captured by smartphone on campus during rainfall events.</p>
Full article ">Figure 3
<p>Schematic representation of the CNN architecture in this study.</p>
Full article ">Figure 4
<p>The regression CNN workflow.</p>
Full article ">Figure 5
<p>Schematic structure of the CNN model importing image data, thresholding, partitioning, training, and testing (deep learning) for rainfall intensity prediction.</p>
Full article ">Figure 6
<p>Rainfall intensity distribution corresponding to (<b>a</b>) surveillance camera and (<b>b</b>) smartphone camera images.</p>
Full article ">Figure 7
<p>(<b>a</b>) displays the raw rain image, (<b>b</b>) is a sharpened image of the raw image input, (<b>c</b>) is a greyscale image that shows pixel intensity, and (<b>d</b>) displays the outcome of applying Otsu’s thresholding method. (<b>e</b>) combines the thresholding approach with image processing to merge two images.</p>
Full article ">Figure 8
<p>Samples of pre-processed images using Otsu’s method under different rainfall conditions: (<b>a</b>) no or low rain, (<b>b</b>) moderate rain, and (<b>c</b>) heavy rain.</p>
Full article ">Figure 9
<p>Observed vs. predicted rainfall intensity by CNN Model 4 using rainfall images captured by a surveillance camera. The blue dot line shows the fitted line corresponding to the R<sup>2</sup>.</p>
Full article ">Figure 10
<p>Observed vs. simulated rainfall intensities by CNN Model 4 using Approach 2 on the smartphone testing dataset.</p>
Full article ">
23 pages, 21344 KiB  
Article
Vertical Structure of Heavy Rainfall Events in Brazil
by Eliana Cristine Gatti, Izabelly Carvalho da Costa and Daniel Vila
Meteorology 2024, 3(3), 310-332; https://doi.org/10.3390/meteorology3030016 - 23 Sep 2024
Viewed by 376
Abstract
Intense rainfall events frequently occur in Brazil, often leading to rapid flooding. Despite their recurrence, there is a notable lack of sub-daily studies in the country. This research aims to assess patterns related to the structure and microphysics of clouds driving intense rainfall [...] Read more.
Intense rainfall events frequently occur in Brazil, often leading to rapid flooding. Despite their recurrence, there is a notable lack of sub-daily studies in the country. This research aims to assess patterns related to the structure and microphysics of clouds driving intense rainfall in Brazil, resulting in high accumulation within 1 h. Employing a 40 mm/h threshold and validation criteria, 83 events were selected for study, observed by both single and dual-polarization radars. Contoured Frequency by Altitude Diagrams (CFADs) of reflectivity, Vertical Integrated Liquid (VIL), and Vertical Integrated Ice (VII) are employed to scrutinize the vertical cloud characteristics in each region. To address limitations arising from the absence of polarimetric coverage in some events, one case study focusing on polarimetric variables is included. The results reveal that the generating system (synoptic or mesoscale) of intense rain events significantly influences the rainfall pattern, mainly in the South, Southeast, and Midwest regions. Regional CFADs unveil primary convective columns with 40–50 dBZ reflectivity, extending to approximately 6 km. The microphysical analysis highlights the rapid structural intensification, challenging the event predictability and the issuance of timely, specific warnings. Full article
(This article belongs to the Special Issue Early Career Scientists' (ECS) Contributions to Meteorology (2024))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spatial distribution of CEMADEN, DECEA, and SIPAM radars in the Brazilian territory, with INMET stations selected for the study.</p>
Full article ">Figure 2
<p>Examples of case validations. For each time analysis, an area 5 × 5 km was created centered on the same station coordinate, and the time of the pixel with the highest value (PMAX) is recorded.</p>
Full article ">Figure 3
<p>Example of the tracking carried out for a case that occurred in the municipality of Feira de Santana-BA, which is covered by the Salvador radar. The colors indicate the shapefile extracted at each time step of the storm.</p>
Full article ">Figure 4
<p>Procedure carried out to construct CFADs.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>e</b>) North region, Northeast region, Midwest region, Southeast region and South region. VIL calculation for the analyzed time instants. The PMAX is the reference period for the highest reflectivity value over the station location during the event. The colors indicate the median values of the VIL values, with shades of blue referring to higher medians (higher VIL values) and shades of brown to lower median values (lower VIL values). The red dots are the outliars.</p>
Full article ">Figure 6
<p>Similar to <a href="#meteorology-03-00016-f005" class="html-fig">Figure 5</a> but for cloud-integrated ice content (VII). (<b>a</b>–<b>d</b>) North region, Northeast region, Midwest region, Southeast region. Note that the y-axis in panel (<b>e</b>) differs from the other panels.</p>
Full article ">Figure 7
<p>CFAD of the Northern region of Brazil created from an area 5 × 5 km (25 km<sup>2</sup>), centered on the pixel with the highest VIL for 15 intense rain events selected in the region. n = 375 refers to the number of vertical profiles used in generating the CFAD. As 25 vertical profiles were extracted for each event (due to the size of the area) and 15 cases were studied in this region, there were a total of 375 vertical profiles in analyzing the events as a whole.</p>
Full article ">Figure 8
<p>Similar to <a href="#meteorology-03-00016-f007" class="html-fig">Figure 7</a> but for the Northeast region of Brazil. In total, 200 vertical profiles were used, referring to 08 selected events.</p>
Full article ">Figure 9
<p>Similar to <a href="#meteorology-03-00016-f007" class="html-fig">Figure 7</a> but for the Midwest region of Brazil. In total, 350 vertical profiles were used, referring to 08 selected events.</p>
Full article ">Figure 10
<p>Similar to <a href="#meteorology-03-00016-f007" class="html-fig">Figure 7</a> but for the Southeast region of Brazil. In total, 725 vertical profiles were used, referring to 08 selected events.</p>
Full article ">Figure 11
<p>Similar to <a href="#meteorology-03-00016-f007" class="html-fig">Figure 7</a> but for the South region of Brazil. In total, 425 vertical profiles were used, referring to 08 selected events.</p>
Full article ">Figure 12
<p>Standard deviations of reflectivity values as a function of height for each instant analyzed in the creation of the CFADs. The vertical line represents the 75th percentile (P75) of the entire deviation dataset. The horizontal line represents the height at which the deviation values are above P75. The colors represent the deviations for each instant and height separated by regions.</p>
Full article ">Figure 13
<p>Location of the Santa Tereza (blue) and Três Marias (green) radars and the INMET automatic stations (red) used in the study. The black dots represent the position of the weather radars.</p>
Full article ">Figure 14
<p>Life cycle of the water (VIL) and ice (VII) contents integrated in the cloud in the highest-intensity pixel (VIL and VII) of each analyzed event (<b>a</b>–<b>g</b>). The dashed vertical line indicates the PMAX instant.</p>
Full article ">Figure 15
<p>CFAD frequency diagram of the reflectivity variable using a 25 km<sup>2</sup> sample centered on the maximum VIL value for each instant analyzed. The CFAD was built from the 7 cases studied, and therefore, with 175 vertical profiles. The PMAX is the reference period in which the maximum reflectivity value on the rain gauge was observed within the hour of recording the accumulated rainfall. The y-axis refers to height in km and the x-axis to reflectivity intervals in dBZ.</p>
Full article ">Figure 16
<p>Similar to <a href="#meteorology-03-00016-f015" class="html-fig">Figure 15</a> but for the <math display="inline"><semantics> <msub> <mi>Z</mi> <mrow> <mi>D</mi> <mi>R</mi> </mrow> </msub> </semantics></math> variable.</p>
Full article ">
21 pages, 6113 KiB  
Article
Cumulative Rainfall Radar Recalibration with Rain Gauge Data Using the Colour Pattern Regression Algorithm QGIS Plugin
by Pablo Blanco-Gómez, Pau Estrany-Planas and José Luis Jiménez-García
Remote Sens. 2024, 16(18), 3496; https://doi.org/10.3390/rs16183496 - 20 Sep 2024
Viewed by 606
Abstract
Climate change is a major issue in wastewater management at local and regional levels, as it affects the frequency of flooding and therefore the need to update infrastructure and design regulations. To this end, rainfall data are the main input to hydraulic models [...] Read more.
Climate change is a major issue in wastewater management at local and regional levels, as it affects the frequency of flooding and therefore the need to update infrastructure and design regulations. To this end, rainfall data are the main input to hydraulic models used for the design of drainage systems and, in advanced contexts, for their real-time monitoring. Field observations are of great interest and water authorities are increasing the number of existing rain gauges, but at present they are scarce and require maintenance, so their number needs to be considered with their O&M costs. Remote sensors, including both the existing satellite rain products (SRPs) and radar imagery (RI), can complete the spatial distribution of rainfall and optimise the cost of observations. While most SRPs are based on re-analysis and have a lag in availability, RI can be obtained in near real time and is becoming increasingly popular in weather forecasting applications. Unfortunately, actual rainfall forecasts from RI observations are not accurate enough for real-time monitoring of drainage systems. In this paper, the Colour Pattern Regression (CPR) algorithm is used to recalibrate the 6 h rainfall values from RI provided by the Agencia Estatal de Meteorología (AEMET) with the observed rain gauge data, using as a case study the metropolitan area of Palma (Spain). Full article
(This article belongs to the Special Issue Remote Sensing of Floods: Progress, Challenges and Opportunities)
Show Figures

Figure 1

Figure 1
<p>Daily discharge calendar for the period 2015–2021. Green colour intensity indicates the number of discharges affected. Source: EMAYA.</p>
Full article ">Figure 2
<p>Location map: location of the city of Palma in the island of Mallorca (Spain); hydro-meteorological and radar stations of the present study and sewage overflow discharges (in orange) into the Mediterranean Sea.</p>
Full article ">Figure 3
<p>Flowchart for the methodology adopted in the present study.</p>
Full article ">Figure 4
<p>Monthly distribution of moderate to extreme rainfall storm events in Palma de Mallorca (Spain) for the period 2015–2021.</p>
Full article ">Figure 5
<p>CPR algorithm results including, for each date and time evaluated, the calculated R, G, and B correlation coefficients and the maximum 6 h cumulative rainfall value observed at any of the existing gauging stations.</p>
Full article ">Figure 6
<p>Maximum 6 h observed rainfall for meteorological radar images that include only <span class="html-italic">Blue</span> and also <span class="html-italic">Green</span> values. (<b>a</b>) Shows the dispersion of both CPR algorithm parameters, and (<b>b</b>) presents the box and whisker results.</p>
Full article ">Figure 7
<p>Performance rating result analysis for (<b>a</b>) NNSE and (<b>b</b>) PBIAS statistics.</p>
Full article ">Figure 8
<p>Minimum and maximum values of the 6 h observed rainfall for meteorological radar image colour ranges. Corresponding to (<b>a</b>) <span class="html-italic">Satisfactory</span>, (<b>b</b>) <span class="html-italic">Good</span>, and (<b>c</b>) <span class="html-italic">Very good</span> performance models.</p>
Full article ">Figure 9
<p>Comparison of the resulting curve with several Z-R relationships adopted from [<a href="#B7-remotesensing-16-03496" class="html-bibr">7</a>,<a href="#B32-remotesensing-16-03496" class="html-bibr">32</a>].</p>
Full article ">
27 pages, 21981 KiB  
Article
A Multi-Scale Analysis of the Extreme Precipitation in Southern Brazil in April/May 2024
by Michelle Simões Reboita, Enrique Vieira Mattos, Bruno César Capucin, Diego Oliveira de Souza and Glauber Willian de Souza Ferreira
Atmosphere 2024, 15(9), 1123; https://doi.org/10.3390/atmos15091123 - 16 Sep 2024
Viewed by 810
Abstract
Since 2020, southern Brazil’s Rio Grande do Sul (RS) State has been affected by extreme precipitation episodes caused by different atmospheric systems. However, the most extreme was registered between the end of April and the beginning of May 2024. This extreme precipitation caused [...] Read more.
Since 2020, southern Brazil’s Rio Grande do Sul (RS) State has been affected by extreme precipitation episodes caused by different atmospheric systems. However, the most extreme was registered between the end of April and the beginning of May 2024. This extreme precipitation caused floods in most parts of the state, affecting 2,398,255 people and leading to 183 deaths and 27 missing persons. Due to the severity of this episode, we need to understand its drivers. In this context, the main objective of this study is a multi-scale analysis of the extreme precipitation between 26 April and 5 May, i.e., an analysis of the large-scale patterns of the atmosphere, a description of the synoptic environment, and an analysis of the mesoscale viewpoint (cloud-top features and lightning). Data from different sources (reanalysis, satellite, radar, and pluviometers) were used in this study, and different methods were applied. The National Center for Monitoring and Early Warning of Natural Disasters (CEMADEN) registered accumulated rainfall above 400 mm between 26 April and 5 May using 27 pluviometers located in the central-northern part of RS. The monthly volumes reached 667 mm and 803 mm, respectively, for April and May 2024, against a climatological average of 151 mm and 137 mm for these months. The maximum precipitation recorded was 300 mm in a single day on 30 April 2024. From a large-scale point of view, an anomalous heat source in the western Indian Ocean triggered a Rossby wave that contributed to a barotropic anticyclonic anomalous circulation over mid-southeastern Brazil. While the precipitant systems were inhibited over this region (the synoptic view), the anomalous stronger subtropical jet southward of the anticyclonic circulation caused uplift over RS State and, consequently, conditions leading to mesoscale convective system (MCS) development. In addition, the low-level jet east of the Andes transported warm and moist air to southern Brazil, which also interacted with two cold fronts that reached RS during the 10-day period, helping to establish the precipitation. Severe deep MCSs (with a cloud-top temperature lower than −80 °C) were responsible for a high lightning rate (above 10 flashes km−2 in 10 days) and accumulated precipitation (above 600 mm in 10 days), as observed by satellite measurements. This high volume of rainfall caused an increase in soil moisture, which exceeded a volume fraction of 0.55, making water infiltration into the soil difficult and, consequently, favoring flood occurrence. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the Rio Grande do Sul (RS) and Santa Catarina (SC) States in Brazil (solid black line); Santiago weather radar (RS, dashed grey circle); natural disasters: flooding (green triangle), flood (white circle), landslide (yellow triangle), flash flood (pink star), rockfall (red cross), tornado (red diamond); and CEMADEN stations (open red circles). (<b>b</b>) Location of the National Center for Monitoring and Early Warning of Natural Disasters (CEMADEN) rain-gauge stations (open red circles) and accumulated precipitation (mm) from 26 April to 5 May 2024.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>f</b>) Problems caused by the precipitation and floods in Rio Grande do Sul State from 26 April to 5 May 2024. Sources: (<b>a</b>) Correio Braziliense [<a href="#B24-atmosphere-15-01123" class="html-bibr">24</a>], (<b>b</b>) Agência Brasil [<a href="#B25-atmosphere-15-01123" class="html-bibr">25</a>], (<b>c</b>) Jovem Pan [<a href="#B26-atmosphere-15-01123" class="html-bibr">26</a>], (<b>d</b>) O Globo [<a href="#B27-atmosphere-15-01123" class="html-bibr">27</a>], (<b>e</b>) UOL [<a href="#B28-atmosphere-15-01123" class="html-bibr">28</a>], and (<b>f</b>) UOL [<a href="#B29-atmosphere-15-01123" class="html-bibr">29</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Monthly precipitation recorded by 27 CEMADEN stations (black line) from January to May 2024 and precipitation climatology from INMET for the 1991–2020 period. (<b>b</b>) Daily precipitation from April to May 2024 was recorded by 27 CEMADEN stations (black line). (<b>c</b>) Daily moisture from April to May 2024 from SMAP for top layer soil moisture (0–5 cm, black line), root zone soil moisture (0–100 cm, blue line), and total profile soil moisture (0 to model bedrock depth, red line). Figures (<b>a</b>) and (<b>b</b>) indicate the period with the heaviest precipitation with grey and red boxes.</p>
Full article ">Figure 4
<p>Flooding area (dark blue) of ARIA/OPERA derived from DSWx-HLS and political mesoregions of Rio Grande do Sul State. In the upper-right figure, the red square highlights the RS state, that has been zoomed in and displayed in the left figure.</p>
Full article ">Figure 5
<p>Flooding area (dark blue) of ARIA/OPERA derived from DSWx-HLS over the metropolitan region of Porto Alegre and the residential areas. Number of houses impacted by flood is in RMPOA: 335.552, Southeast: 49.865, Central-East: 48.293, Central: 1376, Northeast: 1.519, Northwest: 5.185, and Southwest: 3.813. Residential areas are indicated in red on the map. In the upper-right figure, the red square highlights the area that has been zoomed in and displayed in the left figure.</p>
Full article ">Figure 6
<p>April 2024 climate indices and global SST anomalies (°C) based on the OISSTv2 dataset.</p>
Full article ">Figure 7
<p>April 2024 OLR anomalies (W m<sup>−2</sup>) based on NOAA dataset.</p>
Full article ">Figure 8
<p>Zoomed-in view of the April 2024 anomalies in the Indian Ocean: (<b>a</b>) OLR anomalies (W m<sup>−2</sup>) and (<b>b</b>) SST anomalies (°C). The black boxes indicate the areas used to compute IOD.</p>
Full article ">Figure 9
<p>Hovmöller diagram for OLR anomalies (shaded change at 16 W m<sup>−2</sup>) averaged between 5° N and 5° S and tropical disturbances (in colored lines). The x-axis represents longitudes, and the y-axis represents time. Source: NOAA [<a href="#B56-atmosphere-15-01123" class="html-bibr">56</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) Velocity potential anomaly (10<sup>6</sup> m<sup>2</sup> s<sup>−1</sup>) (shaded) and divergent wind component (m s<sup>−1</sup>; arrows) at 250 hPa (right panel) and (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>) stream function anomaly (10<sup>6</sup> m<sup>2</sup> s<sup>−1</sup>) at 250 hPa. Letters L and H indicate low- and high-pressure anomalous centers, respectively.</p>
Full article ">Figure 11
<p>Comparison between (<b>a</b>) a conceptual model of the Rossby wave dispersion at upper levels associated with tropical convection [<a href="#B58-atmosphere-15-01123" class="html-bibr">58</a>] and (<b>b</b>) the case of April 2024: 250 hPa stream function anomaly (10<sup>6</sup> m<sup>2</sup> s<sup>−1</sup>; (contour), velocity potential anomaly (10<sup>6</sup> m<sup>2</sup> s<sup>−1</sup>; shaded), and divergent wind component (m s<sup>−1</sup>; arrows). Letter H indicates high-pressure anomalous centers.</p>
Full article ">Figure 12
<p>Anomaly of wind magnitude (m s<sup>−1</sup>, shaded) and streamlines at (<b>a</b>) 250 hPa and (<b>b</b>) 850 hPa. The anomaly is obtained by the difference between the average from 26 April and 5 May and the average for April and May from 1980 to 2023. Green continuous lines indicate (<b>a</b>) divergence higher than 0.5 × 10<sup>−5</sup> s<sup>−1</sup> and (<b>b</b>) convergence lower than −0.3 × 10<sup>−5</sup> s<sup>−1</sup>, both for the event average.</p>
Full article ">Figure 13
<p>(<b>a</b>–<b>l</b>) Brightness temperature (°C) images from the infrared channel (CH13, 10.35 µm) provided by the ABI sensor aboard the GOES-16 satellite from 26 April 2024 at 1600 UTC to 4 May 2024 at 1930 UTC.</p>
Full article ">Figure 14
<p>(<b>a</b>) Anomaly and (<b>b</b>) average of the sea surface temperature (°C) from 26 April to 5 May 2024, and (<b>c</b>) anomalies of the vertically integrated moisture flux divergence (shaded; 10<sup>−5</sup> kg m<sup>−2</sup> s<sup>−1</sup>) and vertically integrated moisture flux vectors (kg m<sup>−1</sup> s<sup>−1</sup>) between 1000 hPa and 200 hPa. Anomalies were obtained by subtracting the climatology averaged over April and May from 1980 to 2023 from the average for 26 April to 5 May 2024.</p>
Full article ">Figure 15
<p>(<b>a</b>,<b>b</b>) Accumulated precipitation (mm) estimated from MERGE and CHIRPS, (<b>c</b>) maximum soil moisture (volume fraction) from SMAP, and (<b>d</b>) accumulated total lightning flash (flashes/km<sup>2</sup>) from GLM aboard GOES-16 satellite from 26 April to 5 May 2024.</p>
Full article ">Figure 16
<p>(<b>a</b>–<b>l</b>) Constant Plan Position Indicator (CAPPI) at 2 km height of reflectivity from Santiago weather radar (RS) from 27 April at 2000 UTC to 2 May at 0100 UTC.</p>
Full article ">Figure 17
<p>CAPPI at 2 km height of reflectivity and vertical transversal section of a thunderstorm from Santiago radar on 27 April at (<b>a</b>,<b>b</b>) 1950, (<b>c</b>,<b>d</b>) 2000, and (<b>e</b>,<b>f</b>) 2020 UTC. The dashed line in figures (<b>a</b>,<b>c</b>,<b>e</b>) represents the vertical transversal section, and the red diamond is the location of the São Martinho da Serra tornado.</p>
Full article ">
22 pages, 6778 KiB  
Article
Improving Atmospheric Temperature and Relative Humidity Profiles Retrieval Based on Ground-Based Multichannel Microwave Radiometer and Millimeter-Wave Cloud Radar
by Longwei Zhang, Yingying Ma, Lianfa Lei, Yujie Wang, Shikuan Jin and Wei Gong
Atmosphere 2024, 15(9), 1064; https://doi.org/10.3390/atmos15091064 - 3 Sep 2024
Viewed by 526
Abstract
Obtaining temperature and humidity profiles with high vertical resolution is essential for describing and predicting atmospheric motion, and, in particular, for understanding the evolution of medium- and small-scale weather processes, making short-range and near-term weather forecasting, and implementing weather modifications (artificial rainfall, artificial [...] Read more.
Obtaining temperature and humidity profiles with high vertical resolution is essential for describing and predicting atmospheric motion, and, in particular, for understanding the evolution of medium- and small-scale weather processes, making short-range and near-term weather forecasting, and implementing weather modifications (artificial rainfall, artificial rain elimination, etc.). Ground-based microwave radiometers can acquire vertical tropospheric atmospheric data with high temporal and spatial resolution. However, the accuracy of temperature and relative humidity retrieval is still not as accurate as that of radiosonde data, especially in cloudy conditions. Therefore, improving the observation and retrieval accuracy is a major challenge in current research. The focus of this study was to further improve the accuracy of atmospheric temperature and humidity profile retrieval and investigate the specific effects of cloud information (cloud-base height and cloud thickness) on temperature and humidity profile retrieval. The observation data from the ground-based multichannel microwave radiometer (GMR) and the millimeter-wave cloud radar (MWCR) were incorporated into the retrieval process of the atmospheric temperature and relative humidity profiles. The retrieval was performed using the backpropagation neural network (BPNN). The retrieval results were quantified using the mean absolute error (MAE) and root mean square error (RMSE). The statistical results showed that the temperature profiles were less affected by the cloud information compared with the relative humidity profiles. Cloud thickness was the main factor affecting the retrieval of relative humidity profiles, and the retrieval with cloud information was the best retrieval method. Compared with the retrieval profiles without cloud information, the MAE and RMSE values of most of the altitude layers were reduced to different degrees after adding cloud information, and the relative humidity (RH) errors of some altitude layers were reduced by approximately 50%. The maximum reduction in the RMSE and MAE values for the retrieval of temperature profiles with cloud information was about 1.0 °C around 7.75 km, and the maximum reduction in RMSE and MAE values for the relative humidity profiles was about 10%, which was obtained around 2 km. Full article
(This article belongs to the Section Atmospheric Techniques, Instruments, and Modeling)
Show Figures

Figure 1

Figure 1
<p>The fitting graph of S-BT and GMR-BT for channels 6–14 under clear conditions.</p>
Full article ">Figure 2
<p>The fitting graph of C-BT and GMR-BT for channels 6–14 under clear conditions.</p>
Full article ">Figure 3
<p>A comparison of C-BT and GMR-BT in the K-band and V-band under clear conditions.</p>
Full article ">Figure 4
<p>A comparison of C-BT and GMR-BT in the K-band and V-band under cloudy conditions.</p>
Full article ">Figure 5
<p>The MAE comparison of C-BT and GMR-BT for Xi’an’s Jinghe Station in 2018.</p>
Full article ">Figure 6
<p>The impact of cloud information on the temperature (<b>a</b>,<b>b</b>) and relative humidity (<b>c</b>,<b>d</b>) profiles of low cloud conditions.</p>
Full article ">Figure 7
<p>The impact of cloud information on the temperature (<b>a</b>,<b>b</b>) and relative humidity (<b>c</b>,<b>d</b>) profiles of high cloud conditions.</p>
Full article ">Figure 8
<p>A comparison between the temperature profiles generated using the BPNN retrieval models, the GMR product, and the radiosonde under clear conditions at (<b>a</b>) 20:00 (UTC + 8) on 2 November 2018; (<b>b</b>) 20:00 (UTC + 8) on 12 December 2018; (<b>c</b>) 20:00 (UTC + 8) on 14 December 2018; and (<b>d</b>) 20:00 (UTC + 8) on 15 December 2018.</p>
Full article ">Figure 9
<p>The radiosonde data were used as reference measurements to compare the BPNN retrieval models and the GMR product under clear conditions for (<b>a</b>) the GMR and radiosonde data temperature profile scatter plots; (<b>b</b>) the BPNN and radiosonde data temperature profile scatter plots; (<b>c</b>) the RMSE of the temperature profiles; and (<b>d</b>) the MAE of the temperature profiles.</p>
Full article ">Figure 10
<p>A comparison between the relative humidity profiles generated using the BPNN retrieval model, the GMR product, and the radiosonde under clear conditions at (<b>a</b>) 20:00 (UTC + 8) on 2 November 2018; (<b>b</b>) 20:00 (UTC + 8) on 9 December 2018; (<b>c</b>) 20:00 (UTC + 8) on 14 December 2018; and (<b>d</b>) 20:00 (UTC + 8) on 16 December 2018.</p>
Full article ">Figure 11
<p>Using radiosonde data as reference measurements, a comparison of the BPNN retrieval models and the GMR product under clear conditions for (<b>a</b>) the GMR and radiosonde data relative humidity profile scatter plots; (<b>b</b>) the BPNN and radiosonde data relative humidity profile scatter plots; (<b>c</b>) the RMSE of the relative humidity profiles; and (<b>d</b>) the MAE of the relative humidity profiles.</p>
Full article ">Figure 12
<p>A comparison between the temperature profiles generated using the BPNN and BPNN (cloud) retrieval models, the GMR product, and the radiosonde under cloudy conditions at (<b>a</b>) 20:00 (UTC + 8) on 2 November 2018; (<b>b</b>) 20:00 (UTC + 8) on 18 November 2018; (<b>c</b>) 20:00 (UTC + 8) on 11 December 2018; and (<b>d</b>) 20:00 (UTC + 8) on 12 December 2018.</p>
Full article ">Figure 13
<p>Using radiosonde data as reference measurements, a comparison of the BPNN and BPNN (cloud) retrieval models under cloudy conditions for (<b>a</b>) the BPNN and radiosonde data temperature profile scatter plot; (<b>b</b>) the BPNN (cloud) and radiosonde data temperature profile scatter plots; (<b>c</b>) the RMSE of the temperature profiles; and (<b>d</b>) the MAE of the temperature profiles.</p>
Full article ">Figure 14
<p>A comparison between the relative humidity profiles generated using the BPNN and BPNN (cloud) retrieval models, the GMR product, and the radiosonde under cloudy conditions at (<b>a</b>) 20:00 (UTC + 8) on 24 November 2018 (single-layer cloud: cloud-base height: 5190 m; cloud thickness: 2100 m); (<b>b</b>) 20:00 (UTC + 8) on 7 December 2018 (double-layer cloud: cloud-base height: 2850(330) m; cloud thickness: 5970(1560) m); (<b>c</b>) 20:00 (UTC + 8) on 8 December 2018 (single-layer cloud: cloud-base height: 1170 m; cloud thickness: 540 m); and (<b>d</b>) 20:00 (UTC + 8) on 18 December 2018 (single-layer cloud: cloud-base height: 6240 m; cloud thickness: 3450 m).</p>
Full article ">Figure 15
<p>Using radiosonde data as reference measurements, a comparison of the BPNN and BPNN (cloud) retrieval models under cloudy conditions for (<b>a</b>) the BPNN and radiosonde data relative humidity profile scatter plots; (<b>b</b>) the BPNN (cloud) and radiosonde data relative humidity profile scatter plots; (<b>c</b>) the RMSE of the relative humidity profiles; and (<b>d</b>) the MAE of the relative humidity profiles.</p>
Full article ">
18 pages, 9930 KiB  
Article
A Comparative Study of Cloud Microphysics Schemes in Simulating a Quasi-Linear Convective Thunderstorm Case
by Juan Huo, Yongheng Bi, Hui Wang, Zhan Zhang, Qingping Song, Minzheng Duan and Congzheng Han
Remote Sens. 2024, 16(17), 3259; https://doi.org/10.3390/rs16173259 - 2 Sep 2024
Viewed by 595
Abstract
An investigation is undertaken to explore a sudden quasi-linear precipitation and gale event that transpired in the afternoon of 30 May 2024 over Beijing. It was situated at the southwestern periphery of a double-center low-vortex system, where a moisture-rich belt efficiently channeled abundant [...] Read more.
An investigation is undertaken to explore a sudden quasi-linear precipitation and gale event that transpired in the afternoon of 30 May 2024 over Beijing. It was situated at the southwestern periphery of a double-center low-vortex system, where a moisture-rich belt efficiently channeled abundant warm, humid air northward from the south. The interplay between dynamical lifting, convergent airflow-induced uplift, and the amplifying effects of the northern mountainous terrain’s topography creates favorable conditions that support the development and persistence of quasi-linear convective precipitation, accompanied by gale-force winds at the surface. The study also analyzes the impacts of five microphysics schemes (Lin, WSM6, Goddard, Morrison, and WDM6) employed in a weather research and forecasting (WRF) numerical model, with which the simulated rainfall and radar reflectivity are compared against ground-based rain gauge network and weather radar observations, respectively. Simulations with the five microphysics schemes demonstrate commendable skills in replicating the macroscopic quasi-linear pattern of the event. Among the schemes assessed, the WSM6 scheme exhibits its superior agreement with radar observations. The Morrison scheme demonstrates superior performance in predicting cumulative rainfall. Nevertheless, five microphysics schemes exhibit limitations in predicting the rainfall amount, the rainfall duration, and the rainfall area, with a discernible lag of approximately 30 min in predicting precipitation onset, indicating a tendency to forecast peak rainfall events slightly posterior to their true occurrence. Furthermore, substantial disparities emerge in the simulation of the vertical distribution of hydrometeors, underscoring the intricacies of microphysical processes. Full article
Show Figures

Figure 1

Figure 1
<p>Satellite images of the weather event observed by Himawari-8, obtained from observations at three visible bands (blue: 0.47 micron; green: 0.51 micron; red: 0.64 micron): (<b>a</b>–<b>i</b>) 1 h intervals from local time 10:00 to 18:00. The red dots represent the location of the Beijing urban area.</p>
Full article ">Figure 2
<p>Sites of all automatic weather stations and the measured rainfalls and wind speeds in Beijing. (<b>a</b>) Distribution of all AWS sites, the five selected representative sites are presented by green, black, blue, red and purple solid dots. (<b>b</b>) Variations of rainfall observed every minute from 14:00 to 15:00 at the five sites, the value 0 at the <span class="html-italic">X</span>-axis means 14:00 and 120 means 15:01. (<b>c</b>) Distribution of rainfall observed at the time of 14:55; (<b>d</b>) Variations of the wind speed observed every minute from 14:00 to 15:00 at the five sites. Unit of rainfall is mm and unit of the wind speed is ms<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Distribution of 500 hPa (<b>a</b>–<b>c</b>), 750 hPa (<b>d</b>–<b>f</b>), and 950 hPa (<b>g</b>–<b>i</b>) geopotential heights (contours, unit in dagpm), positive vorticity (filled color, unit in 10<sup>−5</sup> s<sup>−1</sup>), and winds during the event. Left column: at 09:00; middle column: at 12:00; right column: at 15:00. The two red areas in (<b>e</b>) denote the low-vortex centers.</p>
Full article ">Figure 4
<p>Distribution of relative humidity (filled color, unit: %), and wind (blue arrows; unit: ms<sup>−1</sup>) at 500 hPa (<b>a</b>–<b>c</b>), 750 hPa (<b>d</b>–<b>f</b>), and 950 hPa (<b>g</b>–<b>i</b>) at different moments. Left column: 09:00; middle column: 12:00; right column: 15:00. Blue circle denotes water vapor transport path at 750 hPa.</p>
Full article ">Figure 5
<p>Topographic map with the red solid dots marking the location of Beijing.</p>
Full article ">Figure 6
<p>Comparisons of radar reflectivity observed at 14:30 on 30 May 2024 simulated with five cloud microphysics schemes. (<b>a</b>) Observed, (<b>b</b>) Lin scheme, (<b>c</b>) WSM6 scheme, (<b>d</b>) Goddard scheme, (<b>e</b>) Morrison scheme, and (<b>f</b>) WDM6 scheme. Unit: dBZ.</p>
Full article ">Figure 7
<p>Rainfall accumulation of the event on 30 May 2024 (accumulated 13:00~16:00, unit mm) observed by AWS and simulated by model. (<b>a</b>) Observed, (<b>b</b>) Lin scheme, (<b>c</b>) WSM6 scheme, (<b>d</b>) Goddard scheme, (<b>e</b>) Morrison scheme, and (<b>f</b>) WDM6 scheme.</p>
Full article ">Figure 8
<p>Quantitative contrast of the rainfall accumulation between simulations and observations: (<b>a</b>) MAE is the mean absolute error; (<b>b</b>) CSI is the critical success index.</p>
Full article ">Figure 9
<p>The vertical distribution of mixing ratios of the cloud droplet, rain, ice, snow and graupel at different stages simulated with five cloud microphysics schemes. Left column: at 14:00, middle column: 14:30, and right column: 15:00. (<b>a</b>–<b>c</b>) Lin scheme, (<b>d</b>–<b>f</b>) WSM6 scheme, (<b>g</b>–<b>i</b>) Goddard scheme, (<b>j</b>–<b>l</b>) Morrison scheme, (<b>m</b>–<b>o</b>) WDM6 scheme. Unit of the mixing ratio in g kg<sup>−1</sup>.</p>
Full article ">
22 pages, 75910 KiB  
Article
Identification and Deformation Characteristics of Active Landslides at Large Hydropower Stations at the Early Impoundment Stage: A Case Study of the Lianghekou Reservoir Area in Sichuan Province, Southwest China
by Xueqing Li, Weile Li, Zhanglei Wu, Qiang Xu, Da Zheng, Xiujun Dong, Huiyan Lu, Yunfeng Shan, Shengsen Zhou, Wenlong Yu and Xincheng Wang
Remote Sens. 2024, 16(17), 3175; https://doi.org/10.3390/rs16173175 - 28 Aug 2024
Viewed by 451
Abstract
Reservoir impoundment imposes a significant triggering effect on bank landslides. Studying the early identification of landslides and their stability concerning reservoir water levels and rainfall is vital for guaranteeing the safety of residents and infrastructure in reservoir regions. This study proposed a method [...] Read more.
Reservoir impoundment imposes a significant triggering effect on bank landslides. Studying the early identification of landslides and their stability concerning reservoir water levels and rainfall is vital for guaranteeing the safety of residents and infrastructure in reservoir regions. This study proposed a method for establishing a dynamic inventory of active landslides at large hydropower stations using integrated remote sensing techniques, demonstrated at Lianghekou Reservoir. We employed interferometric stacking synthetic aperture radar (stacking-InSAR) technology, small baseline subset interferometric synthetic aperture radar (SBAS-InSAR) technology, and optical satellite images to identify and catalogue active landslides. Moreover, we conducted field investigations to examine the deformation characteristics of landslides. Finally, Pearson’s correlation analysis was employed to evaluate the response between deformation values, reservoir water levels, and rainfall. The results revealed 75 active landslides, including 12 long-term active landslides before impoundment and 63 new landslides after impoundment, which were primarily concentrated in the Waduo and Yazho–Zatou regions. The correlation coefficient between landslide deformation values and the reservoir level was high (0.93), while the correlation coefficient with rainfall was low (0.57). The results of this research offer a crucial foundation for preventing and mitigating landslides in reservoir areas. Full article
(This article belongs to the Topic Landslides and Natural Resources)
Show Figures

Figure 1

Figure 1
<p>Study area overview. (<b>a</b>) Geographical position within Sichuan Province; (<b>b</b>) coverage of Sentinel-1A satellite imagery and historical earthquake distribution; (<b>c</b>) topographical features and reservoir inundation extent; (<b>d</b>) reservoir water level change curve; (<b>e</b>) stratigraphy and fault distribution.</p>
Full article ">Figure 2
<p>Flowchart of the technical methodology of this study.</p>
Full article ">Figure 3
<p>Reservoir water level change curve (<b>A</b>); classification of ascending and descending Sentinel-1A satellite imagery based on reservoir levels (<b>B</b>,<b>C</b>). Note: the number of SAR images utilized is denoted in parentheses.</p>
Full article ">Figure 4
<p>Cumulative phase maps generated using Stacking-InSAR for different stages: (<b>a</b>,<b>e</b>) R-0 Stage; (<b>b</b>,<b>f</b>) R-1 Stage; (<b>c</b>,<b>g</b>) R-2 Stage; (<b>d</b>,<b>h</b>) R-3 Stage. Note: the circled yellow areas indicate regions of abnormal deformation.</p>
Full article ">Figure 5
<p>Examples of deformation zones rejected by InSAR technology in the study area. (<b>a</b>,<b>b</b>) Deformation triggered by human engineering activities; (<b>c</b>,<b>d</b>) deformation caused by melting snow and ice; (<b>e</b>–<b>h</b>) Optical images corresponding to the deformation zones illustrated in (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 6
<p>(<b>a</b>) Deformation zones identified using stacking-InSAR; (<b>b</b>) Landslide boundaries delineated from optical imagery.</p>
Full article ">Figure 7
<p>Case study of landslide identification using optical satellite images: (<b>a</b>) Gebu landslide; (<b>b</b>) Maxi landslide.</p>
Full article ">Figure 8
<p>Active landslides before and after reservoir impoundment. (<b>a</b>) Changes in quantity; (<b>b</b>) distribution characteristics. Notes: A: long-term active landslides before impoundment; B: new active landslides after impoundment; B<sub>1</sub>: new landslides after one-phase impoundment; B<sub>2</sub>: new landslides after two-phase impoundment; B<sub>3</sub>: new landslides after three-phase impoundment. The number in parentheses is the number of active landslides.</p>
Full article ">Figure 9
<p>Visualization of Gaussian kernel density analysis results of active landslides.</p>
Full article ">Figure 10
<p>Integrated remote sensing analysis of the Boluzi landslide. (<b>a</b>) Landslide location; (<b>b</b>) optical image; (<b>c</b>) stacking-InSAR annual deformation phase result; (<b>d</b>) airborne LiDAR-based hill shade image.</p>
Full article ">Figure 11
<p>Airborne LiDAR interpretation results and field investigation of the Boluzi landslide. (<b>a</b>) Digital orthophoto; (<b>b</b>) localized collapse; (<b>c</b>) settlement cracking of the roadbed; (<b>d</b>) right boundary and leading edge of the large deformation zone; (<b>e</b>–<b>h</b>) localized collapse.</p>
Full article ">Figure 12
<p>(<b>a</b>) Annual average deformation rate of the Boluzi landslide (LOS); (<b>b</b>) cumulative deformation at different moments.</p>
Full article ">Figure 13
<p>Time series plot of cumulative deformation variables, reservoir water level, and rainfall at the three monitoring sites.</p>
Full article ">Figure 14
<p>Comprehensive remote sensing interpretation of the Waduo landslide. (<b>a</b>) Landslide location; (<b>b</b>) optical image; (<b>c</b>) stacking-InSAR annual deformation phase result; (<b>d</b>) airborne LiDAR-based hill shade image.</p>
Full article ">Figure 15
<p>Field investigation and airborne LiDAR interpretation results of the Waduo landslide. (<b>a</b>) Digital orthophoto; (<b>b</b>,<b>c</b>) boundary tension cracks; (<b>d</b>) H1 trailing edge cracks; (<b>e</b>) H1 leading edge cracks; (<b>f</b>) H2 roadbed settlement cracking; (<b>g</b>,<b>h</b>) H1 leading edge collapse; (<b>i</b>) H2 leading edge collapse.</p>
Full article ">Figure 16
<p>(<b>a</b>) Annual average deformation rate of the Waduo landslide (LOS); (<b>b</b>) cumulative deformation at different moments.</p>
Full article ">Figure 17
<p>Time series plot of cumulative deformation variables, reservoir water levels, and rainfall at three monitoring sites of H1.</p>
Full article ">Figure 18
<p>Time series plot of cumulative deformation variables, reservoir water levels, and rainfall at three monitoring sites of H2.</p>
Full article ">Figure 19
<p>Pearson’s correlation coefficients between deformation and reservoir water level.</p>
Full article ">Figure 20
<p>Pearson’s correlation coefficients between deformation and rainfall.</p>
Full article ">Figure 21
<p>Sentinel-1A visibility analysis of the study area. (<b>a</b>) Ascending orbit coverage distribution; (<b>b</b>) descending orbit coverage distribution; (<b>c</b>) joint ascending and descending orbit coverage. Pie charts in (<b>a</b>–<b>c</b>) illustrate the percentage of the Sentinel-1A data visibility area. Note: colors in the pie charts correspond to the legends in (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 22
<p>(<b>a</b>) Relationship between cumulative displacement, cumulative precipitation, and daily precipitation for the Boluzi landslide; (<b>b</b>) local enlargement of Step Zone I; (<b>c</b>) local enlargement of Step Zone III.</p>
Full article ">Figure 23
<p>Relationship between cumulative displacement, cumulative precipitation, and daily precipitation for the Waduo landslide.</p>
Full article ">
19 pages, 98931 KiB  
Article
Semi-Automatic Detection of Ground Displacement from Multi-Temporal Sentinel-1 Synthetic Aperture Radar Interferometry Analysis and Density-Based Spatial Clustering of Applications with Noise in Xining City, China
by Dianqiang Chen, Qichen Wu, Zhongjin Sun, Xuguo Shi, Shaocheng Zhang, Yi Zhang and Yunlong Wu
Remote Sens. 2024, 16(16), 3066; https://doi.org/10.3390/rs16163066 - 21 Aug 2024
Viewed by 745
Abstract
The China Loess Plateau (CLP) is the world’s most extensive and thickest region of loess deposits. The inherently loose structure of loess makes the CLP particularly vulnerable to geohazards such as landslides, collapses, and subsidence, resulting in substantial geological and environmental challenges. Xining [...] Read more.
The China Loess Plateau (CLP) is the world’s most extensive and thickest region of loess deposits. The inherently loose structure of loess makes the CLP particularly vulnerable to geohazards such as landslides, collapses, and subsidence, resulting in substantial geological and environmental challenges. Xining City, situated at the northwest edge of the CLP, is especially prone to frequent geological hazards due to intensified human activities and natural forces. Synthetic Aperture Radar Interferometry (InSAR) has become a widely used tool for identifying landslide hazards and displacement monitoring because of its high accuracy, low cost, and wide coverage. In this study, we utilized the small baseline subset (SBAS) InSAR technique to derive the line of sight (LOS) displacements of Xining City using Sentinel-1 datasets from ascending and descending orbits between October 2014 and September 2022. By integrating LOS displacements from the two datasets, we retrieved the eastward and vertical displacements to characterize the kinematics of active slopes. To identify the active areas semi-automatically, we applied the Density-Based Spatial Clustering of Applications with Noise (DBSCAN) algorithm to cluster InSAR measurement points (IMPs). Forty-eight active slopes with areas ranging from 0.0049 to 0.5496 km2 and twenty-five subsidence-dominant areas ranging from 0.023 to 3.123 km2 were identified across Xining City. Kinematics analysis of the Jiujiawan landslide indicated that acceleration started in August 2016, likely triggered by rainfall, and continued until the landslide. The extreme rainfall in August 2022 may have pushed the Jiujiawan landslide beyond its critical threshold, leading to instability. Additionally, the study identified nine active slopes that threaten the normal operation of the Lanzhou–Xinjiang High-Speed Railway, with kinematic analysis suggesting rainfall-related accelerations. The influence of anthropogenic activities on ground displacements in loess areas was also confirmed through time series displacement analysis. Our results can be leveraged for geohazard prevention and management in Xining City. As SAR image data continue to accumulate, InSAR can serve as a regular tool for maintaining up-to-date landslide inventories, thereby contributing to more sustainable geohazard management. Full article
(This article belongs to the Special Issue Advances in Remote Sensing for Land Subsidence Monitoring)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of our study area. (<b>b</b>) Geological map of Xining City [<a href="#B49-remotesensing-16-03066" class="html-bibr">49</a>].</p>
Full article ">Figure 2
<p>Sentinel-1 InSAR image pairs of (<b>a</b>) ascending and (<b>b</b>) descending orbit datasets.</p>
Full article ">Figure 3
<p>Workflow of semi-automatic detection of ground displacement in this study.</p>
Full article ">Figure 4
<p>Displacement rate maps in the LOS directions of the (<b>a</b>) ascending and (<b>b</b>) descending Sentinel-1 datasets and in the (<b>c</b>) eastward and (<b>d</b>) vertical directions.</p>
Full article ">Figure 5
<p>DBSCAN maps generated from the (<b>a</b>) ascending and (<b>b</b>) descending displacement rate maps, (<b>c</b>) identifying the cluster results by combining (<b>a</b>,<b>b</b>), and (<b>d</b>) the enlarged map.</p>
Full article ">Figure 6
<p>(<b>a</b>) Eastward and (<b>b</b>) vertical displacement rate maps of the Jiujiawan landslide.</p>
Full article ">Figure 7
<p>Time series eastward (E) and vertical (V) cumulative displacements of P1 and P2 in the Jiujiawan landslide.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>c</b>) Eastward and (<b>b</b>,<b>d</b>) vertical displacement rate maps of five typical landslides along the LXHR.</p>
Full article ">Figure 9
<p>Time series (<b>a</b>) eastward and (<b>b</b>) vertical cumulative displacements of P3-P6 in <a href="#remotesensing-16-03066-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 10
<p>Subsidence rate maps (<b>a1</b>–<b>f1</b>) and corresponding Google Earth optical images (<b>a2</b>–<b>f2</b>) of six typical subsidence zones.</p>
Full article ">Figure 11
<p>Cumulative vertical displacements of P7 and P8.</p>
Full article ">Figure 12
<p>Displacement rate maps (<b>a1</b>–<b>c1</b>) and corresponding Google Earth optical images (<b>a2</b>–<b>c2</b>) of typical active slopes affected by anthropogenic activities.</p>
Full article ">
20 pages, 6885 KiB  
Review
A Review of Rainfall Estimation in Indonesia: Data Sources, Techniques, and Methods
by Maulana Putra, Mohammad Syamsu Rosid and Djati Handoko
Signals 2024, 5(3), 542-561; https://doi.org/10.3390/signals5030030 - 16 Aug 2024
Viewed by 717
Abstract
Rainfall information with high spatial and temporal resolution are essential in various fields. Heavy rainfall in a short period can cause problems and disasters that result in loss of life and damage to property. Conversely, the absence of rain for an extended period [...] Read more.
Rainfall information with high spatial and temporal resolution are essential in various fields. Heavy rainfall in a short period can cause problems and disasters that result in loss of life and damage to property. Conversely, the absence of rain for an extended period can also have negative social and economic impacts. Data accuracy, wide spatial coverage, and high temporal resolution are challenges in obtaining rainfall information in Indonesia. This article presents information on data sources and methods for measuring rainfall and reviews the latest research regarding statistical algorithms and machine learning to estimate rainfall in Indonesia. Rainfall information in Indonesia was obtained from several sources. Firstly, the method of direct rainfall measurement conducted with both manual and automatic rain gauges was reviewed; however, this data source provided minimal results, with uneven spatial density. Secondly, the application of remote sensing estimation using both radar and weather satellites was reviewed. The estimated rainfall results obtained using remote sensing showed more comprehensive spatial coverage and higher temporal resolution. Finally, we reviewed rainfall products obtained from model calculations, using both statistical and machine learning by integrating measurement and remote sensing data. The results of the review demonstrated that rainfall estimation products applied in remote sensing using machine learning models have the potential to produce more accurate spatial and temporal data. However, the validation of rainfall data from direct measurements is required first. This research’s contribution can provide practitioners and researchers in Indonesia and the surrounding region with information on problems, challenges, and recommendations for optimizing rainfall measurement products using appropriate adaptive technology. Full article
(This article belongs to the Special Issue Rainfall Estimation Using Signals)
Show Figures

Figure 1

Figure 1
<p>The distribution of automatic rain gauges includes 1233 units (green triangles), and the integrated weather radar network consists of 42 units (yellow) operating in the Indonesian region.</p>
Full article ">Figure 2
<p>The most commonly operated rain gauges in Indonesia.</p>
Full article ">Figure 3
<p>Comparison of radar imagery products and illustrations of the PPI, CAPPI, and CMAX products: (<b>a</b>) radar imagery of PPI dBZ products at an elevation of 0.5°, (<b>b</b>) CAPPI dBZ products at 0.5 km, and (<b>c</b>) CMAX products (modified from [<a href="#B58-signals-05-00030" class="html-bibr">58</a>]). (<b>d</b>) PPI, (<b>e</b>) CAPPI, and (<b>f</b>) CMAX measurement mechanisms (modified from [<a href="#B59-signals-05-00030" class="html-bibr">59</a>,<a href="#B60-signals-05-00030" class="html-bibr">60</a>]).</p>
Full article ">Figure 4
<p>Satellite imagery depicting atmospheric conditions in one region of Indonesia during high rainfall intensity: (<b>a</b>) rainfall during floods in Sidoarjo City, (<b>b</b>) time series graph of cloud peak temperature, red box shows the time of occurrence during floods and (<b>c</b>) cloud phase distinction RGB satellite imagery product. The yellowish color signifies a thick and high cloud containing ice particles. Red circle shows the location of Sidoarjo City (modified from [<a href="#B67-signals-05-00030" class="html-bibr">67</a>]).</p>
Full article ">Figure 5
<p>Parallax correction (in km) of Himawari-8 satellite based on cloud-top height of 15 km for the entire Territory of Indonesia. Cloud-top height in convective clouds is an essential feature in extreme weather nowcasting performed by weather forecasters to represent the core location of the severe region of the convective cloud (modified from [<a href="#B88-signals-05-00030" class="html-bibr">88</a>]).</p>
Full article ">Figure 6
<p>Rainfall estimation approaches using (<b>a</b>) rain gauges and (<b>b</b>) radar or weather satellite methods (modified from [<a href="#B89-signals-05-00030" class="html-bibr">89</a>]).</p>
Full article ">Figure 7
<p>An illustration of the (<b>a</b>) TMM, (<b>b</b>) PMM, and (<b>c</b>) WPMM techniques (modified from [<a href="#B24-signals-05-00030" class="html-bibr">24</a>]).</p>
Full article ">Figure 8
<p>Conceptual diagram of the MLP algorithm for rainfall estimation using single-polarization radar (<b>a</b>) and the estimation results compared with the radar equation (<b>b</b>) (modified from [<a href="#B115-signals-05-00030" class="html-bibr">115</a>]).</p>
Full article ">
Back to TopTop