[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 (433)

Search Parameters:
Keywords = TRMM

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 10381 KiB  
Article
Modeling and Application of Drought Monitoring with Adaptive Spatial Heterogeneity Using Eco–Geographic Zoning: A Case Study of Drought Monitoring in Yunnan Province, China
by Quanli Xu, Shan Li, Junhua Yi and Xiao Wang
Water 2024, 16(17), 2500; https://doi.org/10.3390/w16172500 - 3 Sep 2024
Viewed by 483
Abstract
Drought, characterized by frequent occurrences, an extended duration, and a wide range of destruction, has become one of the natural disasters posing a significant threat to both socioeconomic progress and agricultural livelihoods. Large-scale geographical environments often exhibit obvious spatial heterogeneity, leading to significant [...] Read more.
Drought, characterized by frequent occurrences, an extended duration, and a wide range of destruction, has become one of the natural disasters posing a significant threat to both socioeconomic progress and agricultural livelihoods. Large-scale geographical environments often exhibit obvious spatial heterogeneity, leading to significant spatial differences in drought’s development and outcomes. However, traditional drought monitoring models have not taken into account the impact of regional spatial heterogeneity on drought, resulting in evaluation results that do not match the actual situation. In response to the above-mentioned issues, this study proposes the establishment of ecological–geographic zoning to adapt to the spatially stratified heterogeneous characteristics of large-scale drought monitoring. First, based on the principles of ecological and geographical zoning, an appropriate index system was selected to carry out ecological and geographical zoning for Yunnan Province. Second, based on the zoning results and using data from NASA’s Moderate Resolution Imaging Spectroradiometer (MODIS) and the Tropical Rainfall Measuring Mission (TRMM) 3B43, the vegetation condition index (VCI), the temperature condition index (TCI), the precipitation condition index (TRCI), and three topographic factors including the digital elevation model (DEM), slope (SLOPE), and aspect (ASPECT) were selected as model parameters. Multiple linear regression models were then used to establish integrated drought monitoring frameworks at different eco–geographical zoning scales. Finally, the standardized precipitation evapotranspiration index (SPEI) was used to evaluate the monitoring effects of the model, and the spatiotemporal variation patterns and characteristics of winter and spring droughts in Yunnan Province from 2008–2019 were further analyzed. The results show that (1) compared to the traditional non-zonal models, the drought monitoring model constructed based on ecological–geographic zoning has a higher correlation and greater accuracy with the SPEI and (2) Yunnan Province experiences periodic and seasonal drought patterns, with spring being the peak period of drought occurrence and moderate drought and light drought being the main types of drought in Yunnan Province. Therefore, we believe that ecological–geographic zoning can better adapt to geographical spatial heterogeneity characteristics, and the zonal drought monitoring model constructed can more effectively identify the actual occurrence of drought in large regions. This research finding can provide reference for the formulation of drought response policies in large-scale regions. Full article
(This article belongs to the Special Issue Drought Risk Assessment and Human Vulnerability in the 21st Century)
Show Figures

Figure 1

Figure 1
<p>Overall research flowchart.</p>
Full article ">Figure 2
<p>Study area.</p>
Full article ">Figure 3
<p>Eco–geographical zoning map of Yunnan Province.</p>
Full article ">Figure 4
<p>Comparison of drought monitoring accuracy in Yunnan Province before and after zoning.</p>
Full article ">Figure 5
<p>CDI values variation in winter and spring seasons from 2008 to 2019 for each ecological–geographic regions.</p>
Full article ">Figure 6
<p>CDI value changes of different ecological and geographical zones during winter and spring seasons from 2008 to 2013.</p>
Full article ">Figure 7
<p>CDI value changes of different ecological and geographical zones during winter and spring seasons from 2014 to 2019.</p>
Full article ">Figure 8
<p>Proportions of different levels of drought area in different ecological–geographic regions.</p>
Full article ">
13 pages, 8386 KiB  
Article
Nocturnal Extreme Rainfall over the Central Yungui Plateau under Cold and Warm Upper-Level Anomaly Backgrounds during Warm Seasons in 1980–2020
by Weihua Yuan and Zhi Li
Atmosphere 2024, 15(9), 1057; https://doi.org/10.3390/atmos15091057 - 1 Sep 2024
Viewed by 283
Abstract
The spatiotemporal and cloud features of the extreme rainfall under the warm and cold upper-level anomalies over the central Yungui Plateau (YGP) were investigated using the hourly rain gauge records, ERA5 reanalysis data, TRMM, and Fengyun satellite data, aiming to refine the understanding [...] Read more.
The spatiotemporal and cloud features of the extreme rainfall under the warm and cold upper-level anomalies over the central Yungui Plateau (YGP) were investigated using the hourly rain gauge records, ERA5 reanalysis data, TRMM, and Fengyun satellite data, aiming to refine the understanding of different types of extreme rainfall. Extreme rainfall under an upper-level negative temperature anomaly (cold events) presents stronger convective cloud features when compared with the positive temperature anomaly (warm events). The maximum rainfall intensity and duration in cold events is much larger than that of warm events, while the brightness temperature of the cloud top is lower, and the ratio of convective rainfall is higher. In cold events, the middle-to-upper troposphere is dominated by a cold anomaly, and an unstable configuration with upper (lower) cold (warm) anomalies is observed around the central YGP. Although the upper-level temperature anomaly is positive, the anomalous divergence and convergence of southerly and northerly winds, as well as the strong moisture center and upward motions, are also found over the central YGP in warm events. The stronger atmospheric instability and higher convective energy under the upper-level cold anomalous circulation are closely associated with the rainfall features over the central YGP. The results indicate that the upper tropospheric temperature has significant influences on extreme rainfall, and thus more attention should be paid to the upper tropospheric temperature in future analyses. Full article
(This article belongs to the Special Issue Characteristics of Extreme Climate Events over China)
Show Figures

Figure 1

Figure 1
<p>The topography (shading, the color bar on the right, units: m) and the colored dots represent diurnal peaks of extreme rainfall amount (&gt;26 mm h<sup>−1</sup>) averaged over 1980–2020 (the color bar at the bottom, in local solar time).</p>
Full article ">Figure 2
<p>(<b>a</b>) The 1st EOF mode of the temperature anomaly averaged in 2100–0400 LST during 462 extreme event days over central YGP and (<b>b</b>) the principal component. The composite 300 hPa temperature anomalies (unit: K) of (<b>c</b>) cold and (<b>d</b>) warm events. Blue dots mark the focal stations.</p>
Full article ">Figure 3
<p>Anomalous circulations averaged during 2100–0400 LST on cold events. The shading represents the anomalous temperature (unit: K) averaged at (<b>a</b>) 200–400 hPa, (<b>b</b>) 400–600 hPa, (<b>c</b>) 600–700 hPa, and (<b>d</b>) 700–800 hPa. White contours represent the geopotential height (units: gpm) at (<b>a</b>) 150 hPa (interval 100), (<b>c</b>) 500 hPa (interval 50), (<b>d</b>) 800 hPa (interval 20), and (<b>b</b>) specific humidity (interval 0.2, unit: g kg<sup>−1</sup>) averaged at 200–900 hPa. The streamlines are the anomalous wind fields at (<b>a</b>) 350 hPa, (<b>b</b>) 500 hPa (<b>c</b>) 700 hPa, and (<b>d</b>) 800 hPa. The purple contours are the divergence (interval 1, unit: 10<sup>−6</sup> s<sup>−1</sup>) at (<b>a</b>) 150 hPa, (<b>c</b>) are the vertical velocity at 500 hPa (interval 1, unit: 10<sup>−1</sup> Pa s<sup>−1</sup>) and (<b>d</b>) averaged over 700–800 hPa. Blue dots represent the focal stations.</p>
Full article ">Figure 4
<p>Same as the caption in <a href="#atmosphere-15-01057-f003" class="html-fig">Figure 3</a>a–d, but for warm events.</p>
Full article ">Figure 5
<p>Temperature anomalies averaged over 300–400 hPa (shaded, unit: K), anomalous divergence (purple contours, unit: 10<sup>−6</sup> s<sup>−1</sup>, starting from 1, interval 1), anomalous geopotential height (white contours, unit: gpm, starting from −800, interval 100), and winds (streamlines) averaged over 150–200 hPa. The gray contour indicates the elevations of 1000, 1500, and 3000 m. Red dots mark the focal stations. The number on the top center of each panel represents the time, the number before (after) the slash is the day (hour in LST) relative to the cold events, and a negative (positive) value indicates the day before (after) the cold events.</p>
Full article ">Figure 6
<p>Same as the caption in <a href="#atmosphere-15-01057-f005" class="html-fig">Figure 5</a> but for temperature anomalies (shaded, unit: K), winds (black vectors, unit: m s<sup>−1</sup>), and divergence (blue contours, unit: 10<sup>−6</sup> s<sup>−1</sup>, starting from −2, interval 1) averaged over 750–875 hPa. White dots represent the focal stations.</p>
Full article ">Figure 7
<p>Same as the caption in <a href="#atmosphere-15-01057-f005" class="html-fig">Figure 5</a> but for warm events.</p>
Full article ">Figure 8
<p>Same as the caption in <a href="#atmosphere-15-01057-f006" class="html-fig">Figure 6</a> but for warm events.</p>
Full article ">Figure 9
<p>(<b>a</b>) The occurrences of warm (red lines) and cold (blue lines) events in different months and (<b>b</b>) the diurnal peaks in different hours.</p>
Full article ">Figure 10
<p>The mean TBB (units: K) derived from Fengyun satellite measurements for (<b>a</b>) cold and (<b>b</b>) warm events.</p>
Full article ">Figure 11
<p>The maximum, 75%, mean, 25%, and minimum values of (<b>a</b>) durations (units: hours) and (<b>b</b>) maximum hourly rainfall intensity (units: mm h<sup>−1</sup>) of cold (blue) and warm (red) events.</p>
Full article ">Figure 12
<p>Temperature anomaly (shaded, unit: K) and relative humidity (contours, unit: %) at 2100–0400 LST on (<b>a</b>) cold and (<b>b</b>) warm events averaged over 24–26° N.</p>
Full article ">
20 pages, 5597 KiB  
Article
Downscaling TRMM Monthly Precipitation in Cloudy and Rainy Regions and Analyzing Spatiotemporal Variations: A Case Study in the Dongting Lake Basin
by Haonan Xia, Huanhua Peng, Jun Zhai, Haifeng Gao, Diandian Jin and Sijia Xiao
Remote Sens. 2024, 16(16), 2959; https://doi.org/10.3390/rs16162959 - 12 Aug 2024
Viewed by 589
Abstract
High-resolution and accurate precipitation data are essential for hydrological, meteorological, and ecological research at the watershed scale. However, in regions with complex terrain and significant rainfall variability, the limited number of rain gauge stations (RGS) is insufficient, and the spatial resolution of existing [...] Read more.
High-resolution and accurate precipitation data are essential for hydrological, meteorological, and ecological research at the watershed scale. However, in regions with complex terrain and significant rainfall variability, the limited number of rain gauge stations (RGS) is insufficient, and the spatial resolution of existing satellite precipitation data is too low to capture detailed precipitation patterns at the watershed scale. To address this issue, the downscaling of satellite precipitation products has become an effective method to obtain high-resolution precipitation data. This study proposes a monthly downscaling method based on a random forest model, aiming to improve the resolution of precipitation data in cloudy and rainy regions at mid-to-low latitudes. We combined the Google Earth Engine (GEE) platform with a local Python environment, introducing cloud cover characteristics into traditional downscaling variables (latitude, longitude, topography, and vegetation index). The TRMM data were downscaled from 25 km to 1 km, generating high-resolution monthly precipitation data for the Dongting Lake Basin from 2001 to 2019. Furthermore, we analyzed the spatiotemporal variation characteristics of precipitation in the study area. The results show the following: (1) In cloudy and rainy regions, our method improves resolution and detail while maintaining the accuracy of precipitation data; (2) The response of monthly precipitation to environmental variables varies, with cloud cover characteristics contributing more to the downscaling model than vegetation characteristics, helping to overcome the lag effect of vegetation characteristics; and (3) Over the past 20 years, there have been significant seasonal trends in precipitation changes in the study area, with a decreasing trend in winter and spring (January–May) and an increasing trend in summer and autumn (June–December). These results indicate that the proposed method is suitable for downscaling monthly precipitation data in cloudy and rainy regions of the Dongting Lake Basin. Full article
(This article belongs to the Section Ecological Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Location of the Dongting Lake Basin and the distribution of meteorological stations.</p>
Full article ">Figure 2
<p>Flowchart of the TRMM downscaling framework integrating Google Earth Engine and Python-based native machine learning methods.</p>
Full article ">Figure 3
<p>Monthly average importance in distribution of features (<b>a</b>) and annual variation in precipitation and environmental features (<b>b</b>).</p>
Full article ">Figure 4
<p>Model accuracy validation results of five features and their combinations.</p>
Full article ">Figure 5
<p>Validation results of original TRMM precipitation dataset with the RGS.</p>
Full article ">Figure 6
<p>Validation results of two precipitation datasets with 30 rain gauge stations from 2001 to 2019: (<b>a</b>) Annual monthly strategy downscaled data (<b>b</b>) Multi-year monthly strategy downscaled data.</p>
Full article ">Figure 7
<p>Monthly precipitation variations of three datasets from 2001 to 2019.</p>
Full article ">Figure 8
<p>Monthly accuracy validation results of downscaled precipitation data: Correlation Coefficient (CC) (<b>a</b>), Root Mean Square Error (RMSE) (<b>b</b>), and Bias (<b>c</b>).</p>
Full article ">Figure 9
<p>Monthly spatial distribution of downscaled precipitation from TRMM data from 2001 to 2019.</p>
Full article ">Figure 10
<p>Area proportion of average precipitation change trends in different months from 2001 to 2019.</p>
Full article ">Figure 11
<p>Spatial distribution of monthly average precipitation change trends in the Dongting Lake Basin.</p>
Full article ">Figure 12
<p>Proportion of precipitation area with significant variation trends at different altitudes.</p>
Full article ">
16 pages, 4230 KiB  
Article
Water Resources Monitoring in a Remote Region: Earth Observation-Based Study of Endorheic Lakes
by Jeremie Garnier, Rejane E. Cicerelli, Tati de Almeida, Julia C. R. Belo, Julia Curto, Ana Paula M. Ramos, Larissa V. Valadão, Frederic Satge and Marie-Paule Bonnet
Remote Sens. 2024, 16(15), 2790; https://doi.org/10.3390/rs16152790 - 30 Jul 2024
Viewed by 492
Abstract
In the western Andes, climate changes have led to drastic ecological changes during the Pleistocene and Holocene. Given the debate surrounding precipitation pattern changes and the lack of research on lakes in the Chilean Altiplano, this study aims to assess recent climate changes. [...] Read more.
In the western Andes, climate changes have led to drastic ecological changes during the Pleistocene and Holocene. Given the debate surrounding precipitation pattern changes and the lack of research on lakes in the Chilean Altiplano, this study aims to assess recent climate changes. The paper presents an innovative methodology based on Google Earth Engine (GEE), utilizing fluctuations in water levels in endorheic lakes as natural precipitation indicators. Three lakes (Chungará, Miscanti, and Miniques) in isolated drainage systems were studied, where changes in water levels directly reflect rainfall variations. Data from Landsat-OLI 8, Landsat-ETM+, Landsat-TM 5, and MODIS spanning 31 years were processed using the Google Earth Engine platform. The shapes of the water bodies were extracted using hue saturation value (HSV) composites. The surface areas of the lakes were compared with precipitation data from national meteorological stations and the Tropical Rainfall Measuring Mission (TRMM) using linear regression analyses. Both lake area and rainfall volume showed a decrease over time, with varying trends depending on environmental conditions. However, the analysis consistently indicates a reduction in the area and volume of Chilean lakes corresponding to observed rainfall patterns over the past three decades. Full article
Show Figures

Figure 1

Figure 1
<p>Localization of the study areas and metereological stations near the lakes; data source: Biblioteca del Congresso Nacional del Chile (BCN); Infraestructura de Datos Geoespaciales Chile (IDE Chile); Dirección General de Aguas (DGA).</p>
Full article ">Figure 2
<p>Methodological flowchart of the procedures.</p>
Full article ">Figure 3
<p>Dispersion graph between Miscanti Lake area and Miniques Lake area: green—2010–2017; orange—2000–2009; purple—1990–1999; red—1980.</p>
Full article ">Figure 4
<p>Chungará extent (red diamond) compared to the annual rainfall from 1987 to 2019.</p>
Full article ">Figure 5
<p>Miscanti and Miniques lakes area (red diamond) after the rainy period calculated from 1986 to 2017 compared to their annual rainfall.</p>
Full article ">Figure 6
<p>Miscanti and Miniques Lakes extend over time after the last day of the rainy season in 2015.</p>
Full article ">Figure 7
<p>Dispersion graphs of annual rainfall vs. Chungará Lake area at the end of the wet season for the years.</p>
Full article ">Figure 7 Cont.
<p>Dispersion graphs of annual rainfall vs. Chungará Lake area at the end of the wet season for the years.</p>
Full article ">Figure 8
<p>Linear regression analysis and dispersion graphs of annual rainfall (<span class="html-italic">X</span> axis) versus Miscanti and Miniques Lakes’ area (<span class="html-italic">Y</span> axis) at the end of the wet season for the years.</p>
Full article ">
16 pages, 3948 KiB  
Article
A Downscaling Method of TRMM Satellite Precipitation Based on Geographically Neural Network Weighted Regression: A Case Study in Sichuan Province, China
by Ge Zheng, Nan Zhang, Laifu Zhang, Yijun Chen and Sensen Wu
Atmosphere 2024, 15(7), 792; https://doi.org/10.3390/atmos15070792 - 30 Jun 2024
Viewed by 582
Abstract
Spatial downscaling is an effective way to improve the spatial resolution of precipitation products. However, the existing methods often fail to adequately consider the spatial heterogeneity and complex nonlinearity between precipitation and surface parameters, resulting in poor downscaling performance and inaccurate expression of [...] Read more.
Spatial downscaling is an effective way to improve the spatial resolution of precipitation products. However, the existing methods often fail to adequately consider the spatial heterogeneity and complex nonlinearity between precipitation and surface parameters, resulting in poor downscaling performance and inaccurate expression of regional details. In this study, we propose a precipitation downscaling model based on geographically neural network weighted regression (GNNWR), which integrates normalized difference vegetation index, digital elevation model, land surface temperature, and slope data to address spatial heterogeneity and complex nonlinearity. We explored the spatiotemporal trends of precipitation in the Sichuan region over the past two decades. The results show that the GNNWR model outperforms common methods in downscaling precipitation for the four distinct seasons, achieving a maximum R2 of 0.972 and a minimum RMSE of 3.551 mm. Overall, precipitation in Sichuan Province exhibits a significant increasing trend from 2001 to 2019, with a spatial distribution pattern of low in the northwest and high in the southeast. The GNNWR downscaled results exhibit the strongest correlation with observed data and provide a more accurate representation of precipitation spatial patterns. Our findings suggest that GNNWR is a practical method for precipitation downscaling considering its high accuracy and model performance. Full article
(This article belongs to the Special Issue Regional Climate Predictions and Impacts)
Show Figures

Figure 1

Figure 1
<p>Distribution map of meteorological stations in Sichuan Province.</p>
Full article ">Figure 2
<p>Definition of GNNWR downscaling model for precipitation.</p>
Full article ">Figure 3
<p>GNNWR neural network architecture and implementation strategy.</p>
Full article ">Figure 4
<p>The implementation process of downscaling algorithm for the GNNWR model.</p>
Full article ">Figure 5
<p>Comparison of downscaling images of different models. (1–4) represents spring, summer, autumn, and winter, respectively; (<b>a</b>) Raw TRMM data (~27.5 km); (<b>b</b>) RF downscaling data; (<b>c</b>) GWR downscaling data; (<b>d</b>) MGWR downscaling data; (<b>e</b>) GNNWR downscaling data.</p>
Full article ">Figure 6
<p>Time and monthly variation trends of annual total precipitation in Sichuan Province from 2001 to 2019. (<b>a</b>) Time variation trend of annual total precipitation; (<b>b</b>) Monthly precipitation variation trend.</p>
Full article ">Figure 7
<p>Spatial distribution of monthly average precipitation in Sichuan Province from 2001 to 2019.</p>
Full article ">
18 pages, 4992 KiB  
Article
Assessment of Satellite Products in Estimating Tropical Cyclone Remote Precipitation over the Yangtze River Delta Region
by Xinyue Wu, Yebing Liu, Shulan Liu, Yubing Jin and Huiyan Xu
Atmosphere 2024, 15(6), 667; https://doi.org/10.3390/atmos15060667 - 31 May 2024
Viewed by 420
Abstract
Satellite products have shown great potential in estimating torrential rainfall due to their wide and consistent global coverage. This study assessed the monitoring capabilities of satellite products for the tropical cyclone remote precipitation (TRP) over the Yangtze River Delta region (YRDR) associated with [...] Read more.
Satellite products have shown great potential in estimating torrential rainfall due to their wide and consistent global coverage. This study assessed the monitoring capabilities of satellite products for the tropical cyclone remote precipitation (TRP) over the Yangtze River Delta region (YRDR) associated with severe typhoon Khanun (2017) and super-typhoon Mangkhut (2018). The satellite products include the CPC MORPHing technique (CMORPH) data, Tropical Rainfall Measuring Mission 3B42 Version 7 (TRMM 3B42), and Integrated Multi-satellite Retrievals for the Global Precipitation Measurement Mission (GPM IMERG). Eight precision evaluation indexes and statistical methods were used to analyze and evaluate the monitoring capabilities of CMORPH, TRMM 3B42, and GPM IMERG satellite precipitation products. The results indicated that the monitoring capability of TRMM satellite precipitation products was superior in capturing the spatial distribution, and GPM products captured the temporal distributions and different category precipitation observed from gauge stations. In contrast, the CMORPH products performed moderately during two heavy rainfall events, often underestimating or overestimating precipitation amounts and inaccurately detecting precipitation peaks. Overall, the three satellite precipitation products showed low POD, high FAR, low TS, and high FBIAS for heavy rainfall events, and the differences in monitoring torrential TRP may be related to satellite retrieval algorithms. Full article
(This article belongs to the Section Meteorology)
Show Figures

Figure 1

Figure 1
<p>Map of (<b>a</b>) southeastern regions of China, and (<b>b</b>) the Yangtze River Delta region (YRDR) study area and locations of meteorological stations. The red points represent the locations of meteorological stations. JS, SH, AH, and ZJ denote Jiangsu, Shanghai, Anhui, and Zhejiang provinces, respectively.</p>
Full article ">Figure 2
<p>Twenty-four-hour cumulative rainfall from observations and satellite data: (<b>a</b>) observations (OBS means observations), (<b>b</b>) CMORPH, (<b>c</b>) TRMM, and (<b>d</b>) GPM from 0000 UTC, 15 October to 0000 UTC, 16 October 2017; the unit is mm. The stars of JS, SH, ZJ denote Jiangsu, Shanghai, and Zhejiang, respectively.</p>
Full article ">Figure 3
<p>Twenty-four-hour cumulative rainfall from observations and satellite datasets: (<b>a</b>) observations (OBS means observations), (<b>b</b>) CMORPH, (<b>c</b>) TRMM, and (<b>d</b>) GPM from 1200 UTC, 16 September to 1200 UTC, 16 September 2018; the unit is mm. The stars of YC, HZ, NB denote Yancheng of Jiangsu province, Hangzhou, and Ningbo of Zhejiang province, respectively.</p>
Full article ">Figure 4
<p>Scatter chart of 24 h accumulated precipitation at weather stations and (<b>a</b>) CMORPH, (<b>b</b>) GPM, and (<b>c</b>) TRMM 24 h accumulated precipitation during the period from 0000 UTC, 15 October 2017 to 0000 UTC, 16 October 2017; the unit is mm.</p>
Full article ">Figure 5
<p>Scatter plots of 24 h accumulated precipitation at different meteorological stations and (<b>a</b>) CMORPH, (<b>b</b>) GPM, and (<b>c</b>) TRMM of 24 h accumulated precipitation during the period from 1200 UTC, 16 September to 1200 UTC, 16 September 2018; the unit is mm.</p>
Full article ">Figure 6
<p>(<b>a</b>) Three-hour average precipitation from observations and satellite datasets, (<b>b</b>) RMSE, (<b>c</b>) BIAS, (<b>d</b>) MAE between observations and rainfall from CMORPH, TRMM, and GPM from 00:00 UTC, 15 October to 00:00 UTC, 16 October 2017; the unit is mm.</p>
Full article ">Figure 7
<p>(<b>a</b>) Three-hour average precipitation data from observations, (<b>b</b>) RMSE, (<b>c</b>) BIAS, (<b>d</b>) MAE between observations and satellite rainfall from CMORPH, TRMM, and GPM from 12:00 UTC, 16 September to 12:00 UTC, 17 September 2018; the unit is mm.</p>
Full article ">Figure 8
<p>(<b>a</b>) POD, (<b>b</b>) FAR, (<b>c</b>) TS, and (<b>d</b>) FBIAS for CMORPH, TRMM, and GPM based on different thresholds from 0000 UTC, 15 October 2017 to 0000 UTC, 16 October 2017.</p>
Full article ">Figure 9
<p>(<b>a</b>) POD, (<b>b</b>) FAR, (<b>c</b>) TS, and (<b>d</b>) FBIAS for CMORPH, TRMM, and GPM based on various thresholds from 1200 UTC, 16 September to 1200 UTC, 16 September 2018.</p>
Full article ">
19 pages, 7654 KiB  
Article
An Innovative Correction–Fusion Approach for Multi-Satellite Precipitation Products Conditioned by Gauge Background Fields over the Lancang River Basin
by Linjiang Nan, Mingxiang Yang, Hao Wang, Hejia Wang and Ningpeng Dong
Remote Sens. 2024, 16(11), 1824; https://doi.org/10.3390/rs16111824 - 21 May 2024
Cited by 1 | Viewed by 664
Abstract
Satellite precipitation products can help improve precipitation estimates where ground-based observations are lacking; however, their relative accuracy and applicability in data-scarce areas remain unclear. Here, we evaluated the accuracy of different satellite precipitation datasets for the Lancang River Basin, Western China, including the [...] Read more.
Satellite precipitation products can help improve precipitation estimates where ground-based observations are lacking; however, their relative accuracy and applicability in data-scarce areas remain unclear. Here, we evaluated the accuracy of different satellite precipitation datasets for the Lancang River Basin, Western China, including the Tropical Rainfall Measuring Mission (TRMM) 3B42RT, the Global Precipitation Measurement Integrated Multi-satellitE Retrievals (GPM IMERG), and Fengyun 2G (FY-2G) datasets. The results showed that GPM IMERG and FY-2G are superior to TRMM 3B42RT for meeting local research needs. A subsequent bias correction on these two datasets significantly increased the correlation coefficient and probability of detection of the products and reduced error indices such as the root mean square error and mean absolute error. To further improve data quality, we proposed a novel correction–fusion method based on window sliding data correction and Bayesian data fusion. Specifically, the corrected FY-2G dataset was merged with GPM IMERG Early, Late, and Final Runs. The resulting FY-Early, FY-Late, and FY-Final fusion datasets showed high correlation coefficients, strong detection performances, and few observation errors, thereby effectively extending local precipitation data sources. The results of this study provide a scientific basis for the rational use of satellite precipitation products in data-scarce areas, as well as reliable data support for precipitation forecasting and water resource management in the Lancang River Basin. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Topographic map of the study area.</p>
Full article ">Figure 2
<p>Flow chart of the proposed window sliding data correction method.</p>
Full article ">Figure 3
<p>Box charts showing the evaluation results for five satellite precipitation products: (<b>a</b>) the results of CC at daily scale; (<b>b</b>) the results of RMSE at daily scale; (<b>c</b>) the results of CC at monthly scale; (<b>d</b>) the results of RB at monthly scale; (<b>e</b>) the results of CC at annual scale; (<b>f</b>) the results of RB at annual scale. The different boxes in the figure represent different satellite precipitation products. CC: correlation coefficient; RMSE: root mean square error; RB: relative bias.</p>
Full article ">Figure 4
<p>Radar map of SPP detection results for different precipitation grades. POD: probability of detection; FAR: false alarm rate; CSI: critical success index; ETS: fair precursor score; FBI: Frequency Bias Index.</p>
Full article ">Figure 5
<p>Map of correction and validation sites in the Lancang River Basin.</p>
Full article ">Figure 6
<p>Correction effect of satellite precipitation products under different window side lengths.</p>
Full article ">Figure 7
<p>Evaluation of the effect of correction for different satellite precipitation products.</p>
Full article ">Figure 8
<p>Spatial distribution of the correction effects for different satellite precipitation products.</p>
Full article ">Figure 9
<p>Spatial distribution of CC values before and after fusion for different satellite precipitation products.</p>
Full article ">
25 pages, 19921 KiB  
Article
Evaluation of Daily and Hourly Performance of Multi-Source Satellite Precipitation Products in China’s Nine Water Resource Regions
by Hongji Gu, Dingtao Shen, Shuting Xiao, Chunxiao Zhang, Fengpeng Bai and Fei Yu
Remote Sens. 2024, 16(9), 1516; https://doi.org/10.3390/rs16091516 - 25 Apr 2024
Cited by 1 | Viewed by 785
Abstract
Satellite precipitation products (SPPs) are of great significance for water resource management and utilization in China; however, they suffer from considerable uncertainty. While numerous researchers have evaluated the accuracy of various SPPs, further investigation is needed to assess their performance across China’s nine [...] Read more.
Satellite precipitation products (SPPs) are of great significance for water resource management and utilization in China; however, they suffer from considerable uncertainty. While numerous researchers have evaluated the accuracy of various SPPs, further investigation is needed to assess their performance across China’s nine major water resource regions. This study used the latest precipitation dataset of the China Meteorological Administration’s Land Surface Data Assimilation System (CLDAS-V2.0) as the benchmark and evaluated the performance of six SPPs—GSMaP, PERSIANN, CMORPH, CHIRPS, GPM IMERG, and TRMM—using six indices: correlation coefficient (CC), root mean square error (RMSE), mean absolute error (MAE), probability of detection (POD), false alarm rate (FAR), and critical success index (CSI), at both daily and hourly scales across China’s nine water resource regions. The conclusions of this study are as follows: (1) The performance of the six SPPs was generally weaker in the west than in the east, with the Continental Basin (CB) exhibiting the poorest performance, followed by the Southwest Basin (SB). (2) At the hourly scale, the performance of the six SPPs was weaker compared to the daily scale, particularly in the high-altitude CB and the high-latitude Songhua and Liaohe River Basin (SLRB), where observing light precipitation and snowfall presents significant challenges. (3) GSMaP, CMORPH, and GPM IMERG demonstrated superior overall performance compared to CHIRPS, PERISANN, and TRMM. (4) CMORPH was found to be better suited for application in drought-prone areas, showcasing optimal performance in the CB and SB. GSMaP excelled in humid regions, displaying the best overall performance in the remaining seven basins. GPM IMERG serves as a complementary precipitation data source for the first two. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the nine major water resource regions in China.</p>
Full article ">Figure 2
<p>Spatial distribution of average precipitation in nine water resource regions for SPPs: (<b>a</b>) daily scale; (<b>b</b>) hourly scale.</p>
Full article ">Figure 2 Cont.
<p>Spatial distribution of average precipitation in nine water resource regions for SPPs: (<b>a</b>) daily scale; (<b>b</b>) hourly scale.</p>
Full article ">Figure 3
<p>Daily scale statistical indicator box plots for nine water resource regions: (<b>a</b>) CC; (<b>b</b>) RMSE; (<b>c</b>) MAE.</p>
Full article ">Figure 4
<p>Daily scale classification indicator box plots for nine water resource regions: (<b>a</b>) POD; (<b>b</b>) FAR; (<b>c</b>) CSI.</p>
Full article ">Figure 5
<p>Spatial distribution of daily scale statistical indicators for the nine water resource regions: (<b>a</b>) CC; (<b>b</b>) RMSE; (<b>c</b>) MAE.</p>
Full article ">Figure 6
<p>Spatial distribution of daily scale classification indicators for the nine water resource regions: (<b>a</b>) POD; (<b>b</b>) FAR; (<b>c</b>) CSI.</p>
Full article ">Figure 7
<p>Hourly scale statistical indicator box plots for nine water resource regions: (<b>a</b>) CC; (<b>b</b>) RMSE; (<b>c</b>) MAE.</p>
Full article ">Figure 8
<p>Hourly scale classification indicator box plots for nine water resource regions: (<b>a</b>) POD; (<b>b</b>) FAR; (<b>c</b>) CSI.</p>
Full article ">Figure 9
<p>Spatial distribution of hourly scale statistical indicators for the nine water resource regions: (<b>a</b>) CC; (<b>b</b>) RMSE; (<b>c</b>) MAE.</p>
Full article ">Figure 10
<p>Spatial distribution of hourly scale classification indicators for the nine water resource regions: (<b>a</b>) POD; (<b>b</b>) FAR; (<b>c</b>) CSI.</p>
Full article ">Figure 10 Cont.
<p>Spatial distribution of hourly scale classification indicators for the nine water resource regions: (<b>a</b>) POD; (<b>b</b>) FAR; (<b>c</b>) CSI.</p>
Full article ">Figure 11
<p>Daily scale indicators and elevation scatter plots of the nine water resource regions: (<b>a</b>) CC; (<b>b</b>) MAE; (<b>c</b>) CSI. In these figures, the color transitions from blue to yellow, representing an increase in the index.</p>
Full article ">Figure 11 Cont.
<p>Daily scale indicators and elevation scatter plots of the nine water resource regions: (<b>a</b>) CC; (<b>b</b>) MAE; (<b>c</b>) CSI. In these figures, the color transitions from blue to yellow, representing an increase in the index.</p>
Full article ">Figure 12
<p>Hourly scale indicators and elevation scatter plots of the nine water resource regions: (<b>a</b>) CC; (<b>b</b>) MAE; (<b>c</b>) CSI. In these figures, the color transitions from blue to yellow, representing an increase in the index.</p>
Full article ">Figure 12 Cont.
<p>Hourly scale indicators and elevation scatter plots of the nine water resource regions: (<b>a</b>) CC; (<b>b</b>) MAE; (<b>c</b>) CSI. In these figures, the color transitions from blue to yellow, representing an increase in the index.</p>
Full article ">Figure 13
<p>Scatter plots of winter daily scale snowfall monitoring data and satellite precipitation data in Harbin City, 2017–2022. In the figure, the straight line represents the fitted line of the scatter plot; the color transitions from red to blue, indicating an increasing point density.</p>
Full article ">Figure 14
<p>Scatter plots of winter hourly scale snowfall monitoring data and satellite precipitation data in Harbin City, 2017–2022. In the figure, the straight line represents the fitted line of the scatter plot; the color transitions from red to blue, indicating an increasing point density.</p>
Full article ">
23 pages, 7475 KiB  
Article
Time-Frequency Aliased Signal Identification Based on Multimodal Feature Fusion
by Hailong Zhang, Lichun Li, Hongyi Pan, Weinian Li and Siyao Tian
Sensors 2024, 24(8), 2558; https://doi.org/10.3390/s24082558 - 16 Apr 2024
Viewed by 616
Abstract
The identification of multi-source signals with time-frequency aliasing is a complex problem in wideband signal reception. The traditional method of first separation and identification especially fails due to the significant separation error under underdetermined conditions when the degree of time-frequency aliasing is high. [...] Read more.
The identification of multi-source signals with time-frequency aliasing is a complex problem in wideband signal reception. The traditional method of first separation and identification especially fails due to the significant separation error under underdetermined conditions when the degree of time-frequency aliasing is high. The single-mode recognition method does not need to be separated first. However, the single-mode features contain less signal information, making it challenging to identify time-frequency aliasing signals accurately. To solve the above problems, this article proposes a time-frequency aliasing signal recognition method based on multi-mode fusion (TRMM). This method uses the U-Net network to extract pixel-by-pixel features of the time-frequency and wave-frequency images and then performs weighted fusion. The multimodal feature scores are used as the classification basis to realize the recognition of the time-frequency aliasing signals. When the SNR is 0 dB, the recognition rate of the four-signal aliasing model can reach more than 97.3%. Full article
(This article belongs to the Special Issue Signal Processing and Machine Learning for Sensor Systems)
Show Figures

Figure 1

Figure 1
<p>Signal mixing model.</p>
Full article ">Figure 2
<p>Example of time-frequency diagrams for different signals ((<b>a</b>): 2ASK; (<b>b</b>): 4FSK; (<b>c</b>): 8PSK; (<b>d</b>): EQFM).</p>
Full article ">Figure 2 Cont.
<p>Example of time-frequency diagrams for different signals ((<b>a</b>): 2ASK; (<b>b</b>): 4FSK; (<b>c</b>): 8PSK; (<b>d</b>): EQFM).</p>
Full article ">Figure 3
<p>Flowchart for generating wave-frequency diagrams.</p>
Full article ">Figure 4
<p>Wave-frequency diagrams of different signals ((<b>a</b>): LFM; (<b>b</b>): EQFM + BPSK + DQPSK + 8PSK).</p>
Full article ">Figure 5
<p>Flow chart of image preprocessing.</p>
Full article ">Figure 6
<p>Time-frequency diagram and corresponding spectrum of BPSK + DQPSK + 8PSK + LFM model ((<b>a</b>). time-frequency distribution, (<b>b</b>). spectrum waveform).</p>
Full article ">Figure 7
<p>Sharpening processing schematic.</p>
Full article ">Figure 8
<p>Comparison of image preprocessing effect.</p>
Full article ">Figure 9
<p>Basic flowchart of TRMM method processing.</p>
Full article ">Figure 10
<p>Time-frequency diagram of MPSK signal ((<b>a</b>): BPSK; (<b>b</b>): QDPSK; (<b>c</b>): 8PSK).</p>
Full article ">Figure 11
<p>Wave-frequency diagram of MPSK signal ((<b>a</b>): BPSK; (<b>b</b>): QPSK; (<b>c</b>): 8PSK).</p>
Full article ">Figure 12
<p>Comparison of pixel frequencies and category weights for each type of signal.</p>
Full article ">Figure 13
<p>Structure of the U-Net network.</p>
Full article ">Figure 14
<p>The image of the ReLU function.</p>
Full article ">Figure 15
<p>Schematic of maximum pooling.</p>
Full article ">Figure 16
<p>Graph of the U-Net network segmentation output.</p>
Full article ">Figure 17
<p>Trends in loss values for U-Net networks ((<b>a</b>). trend of time-frequency plot loss values with training rounds; (<b>b</b>). trend of wave-frequency plot loss values with training rounds).</p>
Full article ">Figure 18
<p>Trend of recognition accuracy of dual signals in multimodal mode with SNR at different aliasing degrees ((<b>a</b>). trend of recognition accuracy versus SNR for dual signals in multimodal mode with 25% overlap; (<b>b</b>). the trend of recognition accuracy versus SNR for dual signals in multimodal mode with 50% overlap; (<b>c</b>). the trend of recognition accuracy versus SNR for dual signals in multimodal mode with 75% overlap; (<b>d</b>). the trend of recognition accuracy versus SNR for dual signals in multimodal mode with 100% overlap).</p>
Full article ">Figure 19
<p>Examples of the recognition results of the TRMM method ((<b>a</b>): BPSK + DQPSK + 8PSK + EQFM; (<b>b</b>): 8PSK + 32QAM + 4FSK + EQFM).</p>
Full article ">Figure 20
<p>Comparison of recognition rates of four-signal aliasing models ((<b>a</b>). 8PSK + 32QAM + 4FSK + EQFM model recognition rate variation graph; (<b>b</b>). 16QAM + 2ASK + AM + LFM model recognition rate variation graph; (<b>c</b>). 16QAM + 4FSK + AM + LFM model recognition rate variation graph; (<b>d</b>). 16QAM + 32QAM + 4FSK + EQFM model recognition rate variation graph; (<b>e</b>). BPSK+ DQPSK + 8PSK + LFM model recognition rate variation graph; (<b>f</b>). BPSK + 4FSK + 2FSK + EQFM model recognition rate variation graph).</p>
Full article ">
22 pages, 18324 KiB  
Article
Spatial Downscaling of Precipitation Data in Arid Regions Based on the XGBoost-MGWR Model: A Case Study of the Turpan–Hami Region
by Huanhuan He, Jinjie Wang, Jianli Ding and Lei Wang
Land 2024, 13(4), 448; https://doi.org/10.3390/land13040448 - 31 Mar 2024
Cited by 2 | Viewed by 1025
Abstract
Accurate and reliable precipitation data are important for analyzing regional precipitation distribution, water resource management, and ecological environment construction. Due to the scarcity of meteorological stations in the Turpan–Hami region, precipitation observation conditions are limited, and it is difficult to obtain precipitation data. [...] Read more.
Accurate and reliable precipitation data are important for analyzing regional precipitation distribution, water resource management, and ecological environment construction. Due to the scarcity of meteorological stations in the Turpan–Hami region, precipitation observation conditions are limited, and it is difficult to obtain precipitation data. Firstly, the applicability of TRMM 3B43v7, GPM_3IMERGM 06, and CMORPH CDR satellite precipitation data for the Turpan–Hami Region was evaluated, and the products with better applicability were selected. Next, the Extreme Gradient Boosting Algorithm (XGBoost) and the Shapley Additive Explanations for Machine Learning (SHAP) model were combined to carry out a feature importance analysis on the climate factors affecting precipitation (mean temperature, actual evapotranspiration, wind speed, cloud cover), from which climate factors with a greater influence on precipitation were selected. Combined with climate factors, normalized difference vegetation index (NDVI), slope, aspect, and elevation as explanatory variables, a Multi-Scale Geographically Weighted Regression (MGWR) model was constructed to obtain the monthly precipitation data of 1 km spatial resolution in the Turpan–Hami area from 2001 to 2020. Finally, the spatiotemporal distribution characteristics and changing trend of precipitation in the Turpan–Hami region from 2001 to 2020 were analyzed. The results show that (1) GPM_3IMERGM 06 satellite precipitation data exhibits good applicability in the Turpan–Hami region. (2) The precision verification of the downscaling results from a monthly scale and an annual scale shows that the accuracy and spatial resolution of the data are improved after downscaling. (3) From 2001 to 2020, the precipitation in the Turpan–Hami region showed an insignificantly increasing trend. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study area; (<b>a</b>) elevation and distribution of meteorological stations in Turpan–Hami region; (<b>b</b>) land use type map in Turpan–Hami region in 2020.</p>
Full article ">Figure 2
<p>Precision evaluation index of satellite precipitation data: (<b>a</b>) correlation coefficients, (<b>b</b>) root-mean-square errors, and (<b>c</b>) relative bias.</p>
Full article ">Figure 3
<p>(<b>a</b>) Mean of the absolute values of the extent to which the explanatory variables affect precipitation; (<b>b</b>) SHAP values of the explanatory variables.</p>
Full article ">Figure 4
<p>(<b>a</b>) Evaluation of the accuracy of GPM monthly precipitation data. (<b>b</b>) Evaluation of the accuracy of GPM_M monthly precipitation data.</p>
Full article ">Figure 5
<p>(<b>a</b>) Evaluation of the accuracy of GPM annual precipitation data. (<b>b</b>) Evaluation of the accuracy of GPM_M annual precipitation data.</p>
Full article ">Figure 6
<p>(<b>a</b>) Multi-year mean distribution of GPM_M precipitation data, 2001–2020. (<b>b</b>) Multi-year average distribution of GPM precipitation data, 2001–2020.</p>
Full article ">Figure 7
<p>(<b>a</b>) Multi-year average monthly precipitation in the Turpan–Hami region, 2001–2020. (<b>b</b>) Multi-year average monthly precipitation in various districts and counties of the Turpan–Hami region, 2001–2020.</p>
Full article ">Figure 8
<p>Distribution of multi-year average monthly precipitation in the Turpan–Hami region, 2001–2020.</p>
Full article ">Figure 9
<p>(<b>a</b>) Seasonal precipitation in the Turpan–Hami region, 2001–2020. (<b>b</b>) Trends in seasonal precipitation in the Turpan–Hami region, 2001–2020.</p>
Full article ">Figure 10
<p>Multi-year seasonal average precipitation distribution in the Turpan–Hami region, 2001–2020 (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">Figure 11
<p>(<b>a</b>) Trends of annual precipitation in the districts and counties of the Turpan–Hami region, 2001–2020. (<b>b</b>) Annual precipitation in the Turpan–Hami region, 2001–2020. (<b>c</b>) Trends of annual precipitation in the Turpan–Hami region, 2001–2020.</p>
Full article ">Figure 12
<p>Centers of gravity of multi-year monthly mean precipitation and their migration trajectories for January–December 2001–2020.</p>
Full article ">Figure 13
<p>Map of the location and trajectory of the center of gravity of precipitation, 2001–2020.</p>
Full article ">Figure 14
<p>Changes in latitude and longitude coordinates of the center of gravity of precipitation, 2001–2020.</p>
Full article ">Figure 15
<p>(<b>a</b>) Sen’s slope estimation results; (<b>b</b>) precipitation changes in the Turpan–Hami region, 2001–2020. SD: significant decrease; NSD: non-significant decrease; NSI: non-significant increase; SI: significant increase.</p>
Full article ">
26 pages, 6287 KiB  
Article
Superiority of Dynamic Weights against Fixed Weights in Merging Multi-Satellite Precipitation Datasets over Pakistan
by Nuaman Ejaz, Aftab Haider Khan, Muhammad Shahid, Kifayat Zaman, Khaled S. Balkhair, Khalid Mohammed Alghamdi, Khalil Ur Rahman and Songhao Shang
Water 2024, 16(4), 597; https://doi.org/10.3390/w16040597 - 17 Feb 2024
Viewed by 1577
Abstract
Satellite precipitation products (SPPs) are undeniably subject to uncertainty due to retrieval algorithms and sampling issues. Many research efforts have concentrated on merging SPPs to create high-quality merged precipitation datasets (MPDs) in order to reduce these uncertainties. This study investigates the efficacy of [...] Read more.
Satellite precipitation products (SPPs) are undeniably subject to uncertainty due to retrieval algorithms and sampling issues. Many research efforts have concentrated on merging SPPs to create high-quality merged precipitation datasets (MPDs) in order to reduce these uncertainties. This study investigates the efficacy of dynamically weighted MPDs in contrast to those using static weights. The analysis focuses on comparing MPDs generated using the “dynamic clustered Bayesian averaging (DCBA)” approach with those utilizing the “regional principal component analysis (RPCA)” under fixed-weight conditions. These MPDs were merged from SPPs and reanalysis precipitation data, including TRMM (Tropical Rainfall Measurement Mission) Multi-satellite Precipitation Analysis (TMPA) 3B42V7, PERSIANN-CDR, CMORPH, and the ERA-Interim reanalysis precipitation data. The performance of these datasets was evaluated in Pakistan’s diverse climatic zones—glacial, humid, arid, and hyper-arid—employing data from 102 rain gauge stations. The effectiveness of the DCBA model was quantified using Theil’s U statistic, demonstrating its superiority over the RPCA model and other individual merging methods in the study area The comparative performances of DCBA and RPCA in these regions, as measured by Theil’s U, are 0.49 to 0.53, 0.38 to 0.45, 0.37 to 0.42, and 0.36 to 0.43 in glacial, humid, arid, and hyper-arid zones, respectively. The evaluation of DCBA and RPCA compared with SPPs at different elevations showed poorer performance at high altitudes (>4000 m). The comparison of MPDs with the best performance of SPP (i.e., TMPA) showed significant improvement of DCBA even at altitudes above 4000 m. The improvements are reported as 49.83% for mean absolute error (MAE), 42.31% for root-mean-square error (RMSE), 27.94% for correlation coefficient (CC), 40.15% for standard deviation (SD), and 13.21% for Theil’s U. Relatively smaller improvements are observed for RPCA at 13.04%, 1.56%, 10.91%, 1.67%, and 5.66% in the above indices, respectively. Overall, this study demonstrated the superiority of DCBA over RPCA with static weight. Therefore, it is strongly recommended to use dynamic variation of weights in the development of MPDs. Full article
(This article belongs to the Section Hydrology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Topographical elevation in Pakistan based on the Shuttle Radar Topography Model (SRTM), (<b>b</b>) classification of Pakistan into four climatic zones showing meteorological stations and their serial numbers in each zone (GMS, HMS, AMS, and HAMS represent meteorological stations in glacial, humid, arid, and hyper-arid zones, respectively), and (<b>c</b>) mean annual precipitation variation across Pakistan.</p>
Full article ">Figure 2
<p>Spatial distributions of temporally averaged DCBA weights for the four merging members during 2000–2015.</p>
Full article ">Figure 3
<p>Spatial distributions of RPCA weights for the four merging members during 2000–2015.</p>
Full article ">Figure 4
<p>Spatial distribution maps in the glacial zone: (<b>a</b>) MAE, (<b>b</b>) RMSE for the DCBA (<b>left</b> side); (<b>a</b>) MAE, (<b>b</b>) RMSE for the RCPA (<b>right</b> side).</p>
Full article ">Figure 5
<p>Spatial distribution maps in the glacial zone: (<b>a</b>) CC, (<b>b</b>) SD, and (<b>c</b>) Theil’s U for the DCBA (<b>left</b> side); (<b>a</b>) CC, (<b>b</b>) SD, and (<b>c</b>) Theil’s U for the RPCA (<b>right</b> side).</p>
Full article ">Figure 6
<p>Spatial distribution maps in the humid zone: (<b>a</b>) MAE, (<b>b</b>) RMSE for the DCBA (<b>left</b> side); (<b>a</b>) MAE, (<b>b</b>) RMSE for the RCPA (<b>right</b> side).</p>
Full article ">Figure 7
<p>Spatial distribution maps in the humid zone: (<b>a</b>) CC, (<b>b</b>) SD, and (<b>c</b>) Theil’s U for the DCBA (<b>left</b> side); (<b>a</b>) CC, (<b>b</b>) SD, and (<b>c</b>) Theil’s U for the RPCA (<b>right</b> side).</p>
Full article ">Figure 8
<p>Spatial distribution maps in the arid zone: (<b>a</b>) MAE, (<b>b</b>) RMSE for the DCBA (<b>left</b> side); (<b>a</b>) MAE, (<b>b</b>) RMSE for the RCPA (<b>right</b> side).</p>
Full article ">Figure 9
<p>Spatial distribution maps in the arid zone: (<b>a</b>) CC, (<b>b</b>) SD, and (<b>c</b>) Theil’s U for the DCBA (<b>left</b> side); (<b>a</b>) CC, (<b>b</b>) SD, and (<b>c</b>) Theil’s U for the RPCA (<b>right</b> side).</p>
Full article ">Figure 10
<p>Spatial distribution maps in the hyper-arid zone: (<b>a</b>) MAE, (<b>b</b>) RMSE for the DCBA (<b>left</b> side); (<b>a</b>) MAE, (<b>b</b>) RMSE for the RCPA (<b>right</b> side).</p>
Full article ">Figure 11
<p>Spatial distribution maps in the hyper-arid zone: (<b>a</b>) CC, (<b>b</b>) SD, and (<b>c</b>) Theil’s U for the DCBA (<b>left</b> side); (<b>a</b>) CC, (<b>b</b>) SD, and (<b>c</b>) Theil’s U for the RPCA (<b>right</b> side).</p>
Full article ">
21 pages, 4370 KiB  
Article
Spatiotemporal Distributions of the Thunderstorm and Lightning Structures over the Qinghai–Tibet Plateau
by Yangxingyi Du, Dong Zheng, Yijun Zhang, Wen Yao, Liangtao Xu and Xianggui Fang
Remote Sens. 2024, 16(3), 468; https://doi.org/10.3390/rs16030468 - 25 Jan 2024
Viewed by 1010
Abstract
Utilizing data from the Tropical Rainfall Measuring Mission (TRMM) satellite’s precipitation radar (PR) and lightning imaging sensor (LIS), this study explores the spatiotemporal distributions of thunderstorm and lightning structures over the Qinghai–Tibet Plateau (QTP), an aspect that has not been explored previously. The [...] Read more.
Utilizing data from the Tropical Rainfall Measuring Mission (TRMM) satellite’s precipitation radar (PR) and lightning imaging sensor (LIS), this study explores the spatiotemporal distributions of thunderstorm and lightning structures over the Qinghai–Tibet Plateau (QTP), an aspect that has not been explored previously. The structural aspects are crucial when considering the impact of thunderstorm and lightning activity in the atmospheric processes. Thunderstorms over the QTP show clear spatial variations in both vertical height and horizontal extension. In the southern region, the average heights of 20 dBZ and 30 dBZ echo tops typically exceed 11.2 and 9.3 km, respectively. Meanwhile, in the eastern part, the average coverage areas for reflectivity greater than 20 dBZ and 30 dBZ consistently surpass 1000 and 180 km2, respectively. The spatial distribution of thunderstorm vertical development height relative to the surface aligns more closely with the horizontal extension, indicating stronger convection in the eastern QTP. The thunderstorm flash rate shows an eastward and northward prevalence, while the thunderstorm flash density peaks in the western and northeastern QTP, with a minimum in the southeast. Furthermore, in the eastern QTP, lightning duration, spatial expansion, and radiance are more pronounced, with the average values typically exceeding 0.22 s, 14.5 km, and 0.50 J m−2 sr−1 μm−1, respectively. Monthly variations reveal heightened values during the summer season for thunderstorm vertical extension, areas with reflectivity greater than 30 dBZ, and lightning frequency. Diurnal variations highlight an afternoon increase in thunderstorm vertical and horizontal extension, lightning frequency, duration, and spatial scale. From a statistical perspective, under weak convective conditions, lightning length exhibits a positive correlation with thunderstorm convection intensity, contrasting with the opposite relationship suggested by previous studies. This article further analyzes and discusses the correlations between various thunderstorm and lightning structural parameters, enhancing our understanding of the distinctive features of thunderstorm and lightning activities in the QTP. Full article
Show Figures

Figure 1

Figure 1
<p>Study region in the Qinghai–Tibet Plateau (QTP). (<b>a</b>) Topographic map depicting the study area enclosed by the red line, with altitude represented by color shading. (<b>b</b>) Visualization of the study area selection. Gray dots represent samples of RPFs with lightning, brown solid lines indicate 3000 m contours, and red solid lines represent the adjusted analytical boundaries for the southern and western areas of the QTP.</p>
Full article ">Figure 2
<p>Geographic distributions of average (<b>a</b>) <span class="html-italic">H<sub>20dBZ</sub></span>, (<b>b</b>) <span class="html-italic">H<sub>30dBZ</sub></span>, (<b>c</b>) <span class="html-italic">RH<sub>20dBZ</sub></span>, and (<b>d</b>) <span class="html-italic">RH<sub>30dBZ</sub></span> of thunderstorms over the QTP.</p>
Full article ">Figure 3
<p>Geographic distributions of average (<b>a</b>) <span class="html-italic">A<sub>&gt;20dBZ</sub></span> and (<b>b</b>) <span class="html-italic">A<sub>&gt;30dBZ</sub></span> of thunderstorms over the QTP.</p>
Full article ">Figure 4
<p>Temporal variations in average vertical and horizontal expansion characteristics of QTP thunderstorms: (<b>a</b>) monthly variation and (<b>b</b>) diurnal variation. The corresponding standard deviations are shown in <a href="#app1-remotesensing-16-00468" class="html-app">Tables S1 and S2 in the Supplementary Materials</a>.</p>
Full article ">Figure 5
<p>Geographic distributions of average (<b>a</b>) flash rate per minute, (<b>b</b>) flash density relative to <span class="html-italic">A<sub>&gt;20dBZ</sub></span> per minute, and (<b>c</b>) flash density relative to <span class="html-italic">A<sub>&gt;30dBZ</sub></span> per minute for thunderstorms over the QTP.</p>
Full article ">Figure 6
<p>Temporal variations in average flash rate and flash density of thunderstorms: (<b>a</b>) monthly variation and (<b>b</b>) diurnal variation. The corresponding standard deviations are shown in <a href="#app1-remotesensing-16-00468" class="html-app">Tables S1 and S2 in the Supplementary Materials</a>.</p>
Full article ">Figure 7
<p>Geographic distributions of average (<b>a</b>) flash duration, (<b>b</b>) flash length, (<b>c</b>) flash footprint, and (<b>d</b>) flash radiance over the QTP.</p>
Full article ">Figure 8
<p>Monthly (<b>a</b>) and diurnal (<b>b</b>) variations in average lightning structural parameters. The corresponding standard deviations are shown in <a href="#app1-remotesensing-16-00468" class="html-app">Tables S1 and S2 in the Supplementary Materials</a>.</p>
Full article ">Figure 9
<p>The correlation between the mean value of flash length (<span class="html-italic">L<sub>F</sub></span>) and (<b>a</b>) <span class="html-italic">H<sub>20dBZ</sub></span>, (<b>b</b>) <span class="html-italic">RH<sub>20dBZ</sub></span>, (<b>c</b>) <span class="html-italic">H<sub>30dBZ</sub></span>, and (<b>d</b>) <span class="html-italic">RH<sub>30dBZ</sub></span> based on sample classification.</p>
Full article ">
31 pages, 8739 KiB  
Article
Evaluating and Correcting Temperature and Precipitation Grid Products in the Arid Region of Altay, China
by Liancheng Zhang, Guli Jiapaer, Tao Yu, Jeanine Umuhoza, Haiyang Tu, Bojian Chen, Hongwu Liang, Kaixiong Lin, Tongwei Ju, Philippe De Maeyer and Tim Van de Voorde
Remote Sens. 2024, 16(2), 283; https://doi.org/10.3390/rs16020283 - 10 Jan 2024
Cited by 2 | Viewed by 1270
Abstract
Temperature and precipitation are crucial indicators for investigating climate changes, necessitating precise measurements for rigorous scientific inquiry. While the Fifth Generation of European Centre for Medium-Range Weather Forecasts Atmospheric Reanalysis (ERA5), ERA5 of the Land Surface (ERA5-Land), and China Meteorological Forcing Dataset (CMFD) [...] Read more.
Temperature and precipitation are crucial indicators for investigating climate changes, necessitating precise measurements for rigorous scientific inquiry. While the Fifth Generation of European Centre for Medium-Range Weather Forecasts Atmospheric Reanalysis (ERA5), ERA5 of the Land Surface (ERA5-Land), and China Meteorological Forcing Dataset (CMFD) temperature and precipitation products are widely used worldwide, their suitability for the Altay region of arid and semi-arid areas has received limited attention. Here, we used the Altay region as the study area, utilizing meteorological station data and implementing the residual revision method for temperature and the coefficient revision method for precipitation to rectify inaccuracies in monthly temperature and precipitation records from ERA5-Land, ERA5, and CMFD. We evaluate the accuracy of these datasets before and after correction using bias, Taylor diagrams, and root-mean-square error (RMSE) metrics. Additionally, we employ Tropical Rainfall Measuring Mission satellite precipitation data (TRMM) as a benchmark to assess the performance of ERA5-Land, ERA5, and CMFD monthly precipitation before and after correction. The results revealed significant differences in the temperature and precipitation capture capabilities of ERA5-Land, ERA5, and CMFD in the Altay region. Overall, these data exhibit substantial errors and are not directly suitable for scientific research. However, we applied residual and coefficient revision methods. After this revision, ERA5-Land, ERA5, and CMFD showed significantly improved temperature and precipitation capture capabilities, especially for ERA5-Land. In terms of temperature, post-revision-CMFD (CMFDPR) demonstrated better temperature capture capabilities. All three datasets showed weaker performance in mountainous regions compared to plains. Notably, post-revision-ERA5 (ERA5PR) seemed unsuitable for capturing temperature in the Altay region. Concerning rain, CMFDPR, post-revision-ERA5-Land (ERA5-LandPR) and ERA5PR outperformed TRMM in capturing precipitation. CMFDPR and ERA5-LandPR both outperform ERA5PR. In summary, the revision datasets effectively compensated for the sparse distribution of meteorological stations in the Altay region, providing reliable data support for studying climate change in arid and semi-arid areas. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geographical distribution of the study area and spatial distribution of temperature and precipitation validation and revision points at meteorological observation sites (The standard map number is: GS (2019) 1823).</p>
Full article ">Figure 2
<p>Distribution of residuals for monthly temperature data of ERA5-Land, ERA5, and CMFD.</p>
Full article ">Figure 3
<p>Bottom box plot of bias for ERA5-Land, ERA5, and CMFD temperature (red dots: mean bias, green horizontal line: median bias, black dots: outliers; ERLT represents the ERA5-Land temperature data, ERAT represents the ERA5 temperature data, and CMFDT represents the CMFD temperature data).</p>
Full article ">Figure 4
<p>Taylor diagram of temperature data assessment compared with meteorological observations in the Altay region ((<b>a</b>) represents the ERA5-Land temperature before and after correction; (<b>b</b>) represents the ERA5 temperature before and after correction; (<b>c</b>) represents the CMFD temperature before and after correction. The numbers in the diagram are the corresponding months; i.e., “1” is January; REF represents the observed state).</p>
Full article ">Figure 5
<p>Spatial distribution of RMSE for ERA5-Land, ERA5, and CMFD temperature before and after correction (Using the 32 meteorological observation verification stations as benchmarks, calculating the RMSE for ERA5-Land, ERA5, and CMFD temperatures before and after the correction, employing the inverse distance weighting interpolation method to interpolate the RMSE for the 32 stations, resulting in spatial distribution of RMSE; R-ERLT represents post-revision-ERA5-Land temperature data, R-ERAT represents post-revision-ERA5 temperature data, R-CMFDT represents post-revision-CMFD temperature data).</p>
Full article ">Figure 6
<p>Spatial distribution of CMFD<sub>PR</sub>, ERA5-Land<sub>PR</sub>, and ERA5<sub>PR</sub> temperature in 2018.</p>
Full article ">Figure 7
<p>Bottom box plot of bias for the ERA5-Land<sub>PR</sub>, ERA5<sub>PR</sub>, and CMFD<sub>PR</sub> temperature (red dots: mean bias, green horizontal line: median bias, black dots: outliers).</p>
Full article ">Figure 8
<p>Bottom box plot of bias for ERA5-Land, ERA5, and CMFD precipitation from 1979 to 2018 (red dots: mean bias, green horizontal line: median bias, black dots: outliers; ERLP represents ERA5-Land precipitation, ERAP represents ERA5 precipitation, and CMFDP represents CMFD precipitation).</p>
Full article ">Figure 9
<p>Taylor diagram of precipitation data assessment compared with meteorological observations in the Altay region from 1979 to 2018 ((<b>a</b>) represents the ERA5-Land precipitation before and after correction; (<b>b</b>) represents the ERA5 precipitation before and after correction; and (<b>c</b>) represents the CMFD precipitation before and after correction. The numbers in the diagram are the corresponding months; i.e., “1” is January; R-ERLP represents post-revision-ERA5-Land precipitation, R-ERAP represents post-revision-ERA5 precipitation, and R-CMFDP represents post-revision-CMFD precipitation).</p>
Full article ">Figure 10
<p>Spatial distribution of monthly precipitation for the ERA5-Land<sub>PR</sub>, ERA5<sub>PR</sub>, and CMFD<sub>PR</sub> in 2018.</p>
Full article ">Figure 11
<p>Bottom box plot of bias for the ERA5-Land<sub>PR</sub>, ERA5<sub>PR</sub>, and CMFD<sub>PR</sub> precipitation from 1979 to 2018 (red dots: mean bias, green horizontal line: median bias, black dots: outliers).</p>
Full article ">Figure 12
<p>Bottom box plot of bias for the pre- and post-correction ERA5-Land, ERA5, CMFD and TRRM precipitation from 1998 to 2018 (red dots: mean bias, green horizontal line: median bias, black dots: outliers).</p>
Full article ">Figure 13
<p>Taylor diagram of precipitation data assessment compared with meteorological observations in the Altay region from 1998 to 2018 ((<b>a</b>) represents the TRMM precipitation; (<b>b</b>) represents the ERA5-Land precipitation before and after correction; (<b>c</b>) represents the ERA5 precipitation before and after correction; and (<b>d</b>) represents the CMFD precipitation before and after correction. The numbers in the diagram are the corresponding months, i.e., “1” is January).</p>
Full article ">
30 pages, 9017 KiB  
Article
Combining Hydrological Models and Remote Sensing to Characterize Snowpack Dynamics in High Mountains
by Jamal Hassan Ougahi and John S. Rowan
Remote Sens. 2024, 16(2), 264; https://doi.org/10.3390/rs16020264 - 9 Jan 2024
Cited by 2 | Viewed by 1578
Abstract
Seasonal snowpacks, characterized by their snow water equivalent (SWE), can play a major role in the hydrological cycle of montane environments with months of snow accretion followed by episodes of melt controlling flood risk and water resource availability downstream. Quantifying the temporal and [...] Read more.
Seasonal snowpacks, characterized by their snow water equivalent (SWE), can play a major role in the hydrological cycle of montane environments with months of snow accretion followed by episodes of melt controlling flood risk and water resource availability downstream. Quantifying the temporal and spatial patterns of snowpack accumulation and its subsequent melt and runoff is an internationally significant challenge, particularly within mountainous regions featuring complex terrain with limited or absent observational data. Here we report a new approach to snowpack characterization using open-source global satellite and modelled data products (precipitation and SWE) greatly enhancing the utility of the widely used Soil and Water Assessment Tool (SWAT). The paper focusses on the c. 23,000 km2 Chenab river basin (CRB) in the headwaters of the Indus Basin, globally important because of its large and growing population and increasing water insecurity due to climate change. We used five area-weighted averaged satellite, gridded and reanalysis precipitation datasets: ERA5-Land, CMORPH, TRMM, APHRODITE and CPC UPP. As well as comparison to local weather station data, these were used in SWAT to model streamflow for evaluation against observed streamflow at the basin outlet. ERA5-Land data provided the best streamflow match-ups and was used to infer snowpack (SWE) dynamics at basin and sub-basin scales. Snow reference data were derived from remote sensing and modelled SWE re-analysis products: ULCA-SWE and KRA-SWE, respectively. Beyond conventional auto-calibration and single-variable approaches we undertook multi-variable calibration using R-SWAT to manually adjust snow parameters alongside observed streamflow data. Characterization of basin-wide patterns of snowpack build-up and melt (SWE dynamics) were greatly strengthened using KRA-SWE data accompanied by improved streamflow simulation in sub-basins dominated by seasonal snow cover. UCLA-SWE data also improved SWE estimations using R-SWAT but weakened the performance of simulated streamflow due to under capture of seasonal runoff from permanent snow/ice fields in the CRB. This research highlights the utility and value of remote sensing and modelling data to drive better understanding of snowpack dynamics and their contribution to runoff in the absence of in situ snowpack data in high-altitude environments. An improved understanding of snow-bound water is vital in natural hazard risk assessment and in better managing worldwide water resources in the populous downstream regions of mountain-fed large rivers under threat from climate change. Full article
(This article belongs to the Special Issue Remote Sensing of Floods: Progress, Challenges and Opportunities)
Show Figures

Figure 1

Figure 1
<p>Study area map of the Chenab river basin (CRB) showing weather station, stream, catchment boundary and data points used to extract data from satellite and reanalysis products.</p>
Full article ">Figure 2
<p>Land use/cover map of the Chenab river basin (CRB).</p>
Full article ">Figure 3
<p>Snow water equivalent (SWE) products (<b>a</b>) UCLA-SWE (without masking non-seasonal snow/ice) (<b>b</b>) UCLA-SWE (masking non-seasonal snow/ice pixels) (<b>c</b>) KRA-SWE in the CRB. In legend, values less than 1 are considered as no data or masked area.</p>
Full article ">Figure 4
<p>Flow chart of methodology showing implementation of SWAT model for streamflow and snow water equivalent (SWE) simulation. The simulation of streamflow and the SWE was conducted with uncalibrated SWAT model (SWAT<sub>uncal</sub>) and calibrated SWAT with R-SWAT (R-SWAT<sub>cal</sub>). *Observed precipitation data of weather stations is monthly scale.</p>
Full article ">Figure 5
<p>Monthly comparison of satellite gridded and reanalysis precipitation products (TRMM, CPCUPP, APHRODITE, ERA5)with station data at Muzaffarabad and Garhi Dupatta.</p>
Full article ">Figure 6
<p>Inter-product comparison of spatially averaged the SWE at monthly scale (<b>a</b>) global SWE products and (<b>b</b>) MODIS snow cover area (SCA) in the CRB.</p>
Full article ">Figure 7
<p>Comparison of spatial distribution of the SWEs in each sub-basin (<b>a</b>) UCLA-SWE (<b>b</b>) KRA-SWE in the Chenab River Basin (CRB).</p>
Full article ">Figure 8
<p>Model performance with various precipitation datasets for daily streamflow during 2000–2015. Streamflow simulation was conducted with the default SWAT parameters by inputting data from the satellite and reanalysis precipitation products such as (<b>a</b>) ERA5-Land (<b>b</b>) CPCUPP (<b>c</b>) APHRODITE (<b>d</b>) TRMM (<b>e</b>) CMORPH.</p>
Full article ">Figure 8 Cont.
<p>Model performance with various precipitation datasets for daily streamflow during 2000–2015. Streamflow simulation was conducted with the default SWAT parameters by inputting data from the satellite and reanalysis precipitation products such as (<b>a</b>) ERA5-Land (<b>b</b>) CPCUPP (<b>c</b>) APHRODITE (<b>d</b>) TRMM (<b>e</b>) CMORPH.</p>
Full article ">Figure 8 Cont.
<p>Model performance with various precipitation datasets for daily streamflow during 2000–2015. Streamflow simulation was conducted with the default SWAT parameters by inputting data from the satellite and reanalysis precipitation products such as (<b>a</b>) ERA5-Land (<b>b</b>) CPCUPP (<b>c</b>) APHRODITE (<b>d</b>) TRMM (<b>e</b>) CMORPH.</p>
Full article ">Figure 9
<p>Daily snow water equivalent (SWE) simulation with the SWAT model run with precipitation data from APHRODITE, CPCUPP and ERA5-Land and comparison with SWE products (UCLA and KRA) in the CRB.</p>
Full article ">Figure 10
<p>Daily streamflow simulation using single variable calibration approach during calibration (2000 to 2011) and validation (2011–2015) separated by red dotted line.</p>
Full article ">Figure 11
<p>Comparison of the simulated SWE (SWAT-SWE) with UCLA-SWE and KRA-SWE in the CRB with single variable calibration approach.</p>
Full article ">Figure 12
<p>Daily streamflow simulation with multi-variable calibration approach using observed streamflow and snow reference data during calibration and validation period separated by red dotted line (<b>a</b>) UCLA-SWE and (<b>b</b>) KRA-SWE in the CRB.</p>
Full article ">Figure 13
<p>Comparison of the simulated SWE (SWAT-SWE) with UCLA-SWE and KRA-SWE in the CRB. In panel (<b>a</b>), simulated SWE using UCLA-SWE product during calibration and validation period while panel (<b>b</b>), simulated SWE using KRA SWE product during calibration and validation period are shown.</p>
Full article ">Figure 14
<p>Simulated SWAT-SWE using KRA-SWE and UCLA-SWE product in selected sub-basins for KRA-SWAT SWE (<b>a</b>–<b>f</b>) and UCLA-SWAT (<b>g</b>–<b>l</b>) in selected sub-basins (BSN01, BSN02, BSN05, BSN12, BSN13, and BSN17) of the CRB.</p>
Full article ">
20 pages, 7613 KiB  
Article
Application of TRMM for Spatio-Temporal Analysis of Precipitation in the Taiwan Strait and Its Adjacent Regions
by Yaozhao Zhong, Da Li, Lei Wang, Caiyun Zhang and Feng Zhang
J. Mar. Sci. Eng. 2023, 11(12), 2358; https://doi.org/10.3390/jmse11122358 - 14 Dec 2023
Viewed by 1109
Abstract
Precipitation patterns are highly valued in the fields of weather forecasting, water resource management, and estuary environment research. In this study, daily and monthly precipitation TRMM data from 1998 to 2019 were selected, and EOF analysis was employed to analyze the precipitation patterns [...] Read more.
Precipitation patterns are highly valued in the fields of weather forecasting, water resource management, and estuary environment research. In this study, daily and monthly precipitation TRMM data from 1998 to 2019 were selected, and EOF analysis was employed to analyze the precipitation patterns of the Taiwan Strait and its neighboring regions. We obtained the following results: (1) The rainy season (May–June) is the main contributor to precipitation in the study area. The EOF first mode reflected the overall consistency of the precipitation spatial distribution. However, within each river basin, the magnitude of precipitation variation is spatially different. The magnitude of precipitation variation is significant in the northwestern part of the Minjiang River basin, the southwestern part of the Jiulong River basin, and the southwestern corner of the Hanjiang River basin. These areas happen to correspond to the mountain areas, revealing that topographic precipitation plays a role in the spatial distribution of precipitation in the three river basins. (2) The spatial distributions of the EOF first mode and of precipitation during El Niño in the Minjiang River basin are consistent. This reveals that ENSO is probably the dominant factor in precipitation in the Minjiang River basin. The significant increase in precipitation during El Niño compared with a normal year in the Minjiang River basin confirms this point. (3) In all three strong El Niño years, 1998, 2010, and 2016, the Minjiang River basin experienced significant heavy precipitation in the fall and winter, whereas the Jiulong River and Hanjiang River basins did not (except in 2016). In other words, the Minjiang River basin is more affected by ENSO, while the Jiulong River and Hanjiang River basins are only limitedly impacted by ENSO. Full article
(This article belongs to the Section Physical Oceanography)
Show Figures

Figure 1

Figure 1
<p>Three major river basins on the west coast of the Taiwan Strait (the three areas surrounded by thick black solid lines) and 12 meteorological stations within the river basins (labeled by red pentagrams). The locations and approximate extents of mountains are shown as thick, solid orange lines. I: Wuyi Mountain Range. II: Shanling Mountain. III: Jiufeng Mountain Range. IV: Daiyun Mountain Range. V: Bopingling Mountain. VI: Lianhua Mountain. Base map data from ETOPO [<a href="#B18-jmse-11-02358" class="html-bibr">18</a>] downloaded at <a href="https://www.ncei.noaa.gov/products/etopo-global-relief-model" target="_blank">https://www.ncei.noaa.gov/products/etopo-global-relief-model</a> (accessed on 1 October 2023).</p>
Full article ">Figure 2
<p>(<b>a</b>) Comparison of cumulative probability distribution curves between 12 meteorological stations’ daily precipitation data and TRMM daily precipitation data in the study area. The blue solid lines in the panels represent data from meteorological stations, which were downloaded from <a href="https://www.ncei.noaa.gov/data/global-summary-of-the-day/archive/" target="_blank">https://www.ncei.noaa.gov/data/global-summary-of-the-day/archive/</a> (accessed on 15 October 2023). The red dashed lines represent TRMM data. The top sides of each panel are the names of the weather stations. (<b>b</b>) Daily precipitation comparison between meteorological station data (blue solid line) and TRMM data (red dashed line). The asterisks in the scatterplots present the spread of data relative to the fitted curves.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) Comparison of cumulative probability distribution curves between 12 meteorological stations’ daily precipitation data and TRMM daily precipitation data in the study area. The blue solid lines in the panels represent data from meteorological stations, which were downloaded from <a href="https://www.ncei.noaa.gov/data/global-summary-of-the-day/archive/" target="_blank">https://www.ncei.noaa.gov/data/global-summary-of-the-day/archive/</a> (accessed on 15 October 2023). The red dashed lines represent TRMM data. The top sides of each panel are the names of the weather stations. (<b>b</b>) Daily precipitation comparison between meteorological station data (blue solid line) and TRMM data (red dashed line). The asterisks in the scatterplots present the spread of data relative to the fitted curves.</p>
Full article ">Figure 3
<p>(<b>a</b>) Spatial distribution of multi-year average monthly precipitation in the Taiwan Strait and its neighboring regions for the 1998–2019 period. (<b>b</b>) Histograms of the 1998–2019 multi-year average monthly precipitation in the Taiwan Strait and its neighboring regions, the Minjiang River basin, the Jiulong River basin, and the Hanjiang River basin.</p>
Full article ">Figure 4
<p>(<b>a</b>) Histogram of annual precipitation in the Taiwan Strait and its neighboring regions, the Minjiang River basin, the Jiulong River basin, and the Hanjiang River basin, from 1998 to 2019. (<b>b</b>) Spatial distribution of annual precipitation in the Taiwan Strait and its neighboring regions during the 1998–2019 period.</p>
Full article ">Figure 4 Cont.
<p>(<b>a</b>) Histogram of annual precipitation in the Taiwan Strait and its neighboring regions, the Minjiang River basin, the Jiulong River basin, and the Hanjiang River basin, from 1998 to 2019. (<b>b</b>) Spatial distribution of annual precipitation in the Taiwan Strait and its neighboring regions during the 1998–2019 period.</p>
Full article ">Figure 5
<p>(<b>a</b>) Monthly precipitation anomaly in the Taiwan Strait and its neighboring regions and (<b>b</b>) MEI index, 1998–2019. The MEI value [<a href="#B20-jmse-11-02358" class="html-bibr">20</a>,<a href="#B21-jmse-11-02358" class="html-bibr">21</a>] was downloaded at <a href="https://psl.noaa.gov/enso/mei/" target="_blank">https://psl.noaa.gov/enso/mei/</a> (accessed on 15 October 2023). The red color represents positive value and blue color represents negative value.</p>
Full article ">Figure 6
<p>Results of EOF analysis of monthly precipitation anomalies in the Minjiang River basin from 1998 to 2019. (<b>a</b>,<b>c</b>,<b>e</b>) represent the spatial distributions of the first, second, and third modes, respectively; (<b>b</b>,<b>d</b>,<b>f</b>) represent the time coefficients of the first, second, and third modes, respectively.</p>
Full article ">Figure 7
<p>Results of EOF analysis of monthly precipitation anomalies in the Jiulong River basin from 1998 to 2019. (<b>a</b>,<b>c</b>,<b>e</b>) represent the spatial distributions of the first, second, and third modes, respectively; (<b>b</b>,<b>d</b>,<b>f</b>) represent the time coefficients of the first, second, and third modes, respectively.</p>
Full article ">Figure 8
<p>Results of EOF analysis of monthly precipitation anomalies in the Hanjiang River basin from 1998 to 2019. (<b>a</b>,<b>c</b>,<b>e</b>) represent the spatial distributions of the first, second, and third modes, respectively; (<b>b</b>,<b>d</b>,<b>f</b>) represent the time coefficients of the first, second, and third modes, respectively.</p>
Full article ">Figure 9
<p>Spatial distribution of precipitation for (<b>a</b>) the spring (March–April), (<b>b</b>) the rainy season (May–June), (<b>c</b>) the summer (July–September), and (<b>d</b>) the fall/winter (October–February, multi-year averages, 1998–2019.</p>
Full article ">Figure 10
<p>Spatial distribution of monthly precipitation in the Taiwan Strait and its neighboring regions during El Niño (<b>a</b>) and La Nina (<b>b</b>). Precipitation anomalies during El Niño (<b>c</b>) and La Nina (<b>d</b>).</p>
Full article ">
Back to TopTop