[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

Journals

Article Types

Countries / Regions

Search Results (30)

Search Parameters:
Keywords = rainfall temporal downscaling

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
26 pages, 29445 KiB  
Article
Weather Research and Forecasting Model (WRF) Sensitivity to Choice of Parameterization Options over Ethiopia
by Andualem Shiferaw, Tsegaye Tadesse, Clinton Rowe and Robert Oglesby
Atmosphere 2024, 15(8), 974; https://doi.org/10.3390/atmos15080974 - 14 Aug 2024
Viewed by 710
Abstract
Downscaling seasonal climate forecasts using regional climate models (RCMs) became an emerging area during the last decade owing to RCMs’ more comprehensive representation of the important physical processes at a finer resolution. However, it is crucial to test RCMs for the most appropriate [...] Read more.
Downscaling seasonal climate forecasts using regional climate models (RCMs) became an emerging area during the last decade owing to RCMs’ more comprehensive representation of the important physical processes at a finer resolution. However, it is crucial to test RCMs for the most appropriate model setup for a particular purpose over a given region through numerical experiments. Thus, this sensitivity study was aimed at identifying an optimum configuration in the Weather, Research, and Forecasting (WRF) model over Ethiopia. A total of 35 WRF simulations with different combinations of parameterization schemes for cumulus (CU), planetary boundary layer (PBL), cloud microphysics (MP), longwave (LW), and shortwave (SW) radiation were tested during the summer (June to August, JJA) season of 2002. The WRF simulations used a two-domain configuration with a 12 km nested domain covering Ethiopia. The initial and boundary forcing data for WRF were from the Climate Forecast System Reanalysis (CFSR). The simulations were compared with station and gridded observations to evaluate their ability to reproduce different aspects of JJA rainfall. An objective ranking method using an aggregate score of several statistics was used to select the best-performing model configuration. The JJA rainfall was found to be most sensitive to the choice of cumulus parameterization and least sensitive to cloud microphysics. All the simulations captured the spatial distribution of JJA rainfall with the pattern correlation coefficient (PCC) ranging from 0.89 to 0.94. However, all the simulations overestimated the JJA rainfall amount and the number of rainy days. Out of the 35 simulations, one that used the Grell CU, ACM2 PBL, LIN MP, RRTM LW, and Dudhia SW schemes performed the best in reproducing the amount and spatio-temporal distribution of JJA rainfall and was selected for downscaling the CFSv2 operational forecast. Full article
(This article belongs to the Special Issue Climate Change and Regional Sustainability in Arid Lands)
Show Figures

Figure 1

Figure 1
<p>Model domain and topography (<b>left</b>) and location of weather stations used for verification (<b>right</b>).</p>
Full article ">Figure 2
<p>Experimental set up. Each row contains experiments with KF, BMJ, and Grell CU schemes. First 3 columns: YSU PBL; columns 4–6: experiments with MYJ PBL scheme; and columns 7–9: ACM2 PBL scheme. Columns 1, 4, 7: WSM6 MP scheme; columns 2, 5, 8: LIN MP scheme; and columns 3, 6, 9: Morrison MP scheme. All 27 experiments utilize RRTM LW and Dudhia SW radiation schemes.</p>
Full article ">Figure 3
<p>Mean summer (JJA) rainfall (mm/day) in 2002 for (<b>a</b>) CHIRPS, (<b>b</b>) ENACTS, and WRF simulation. Simulations (<b>c</b>–<b>k</b>), (<b>l</b>–<b>t</b>), and (<b>u</b>–<b>ad</b>) used KF, BMJ, and Grell CU schemes, respectively. Rows 1–3: YSU; rows 4–6: MYJ; and rows 7–9: ACM2 PBL schemes. Rows 1, 4, and 7: WSM6; rows 2, 5, and 8: LIN; and rows 3, 6, and 9: Morrison MP scheme (<a href="#atmosphere-15-00974-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 4
<p>Pattern correlation coefficient (PCC) for JJA mean rainfall between 27 WRF simulations and (<b>a</b>) CHIRPS and (<b>b</b>) ENACTS.</p>
Full article ">Figure 5
<p>Observed number of rainy days over weather stations (<b>top</b>) and bias in WRF-simulated number of rainy days (<b>bottom</b>) during JJA season of 2002 compared to weather stations. For rainy day bias, simulations E1–E9, E10–E18, and E19–E27 used KF, BMJ, and Grell CU schemes, respectively. Columns 1–3: YSU; columns 4–6: MYJ; and columns 7–9: ACM2 PBL schemes. Columns 1, 4, and 7: WSM6; columns 2, 5, and 8: LIN; and columns 3, 6, and 9: Morrison MP scheme.</p>
Full article ">Figure 6
<p>Observed mean rainfall intensity (total rainfall/number of rainy days) (top) and bias in WRF-simulated mean rainfall intensity (E1–E27) compared to weather stations during JJA season of 2002 for 27 experiments. For intensity bias (E1–E27) day bias, simulations E1–E9, E10–E18, and E19–E27 used KF, BMJ, and Grell CU schemes, respectively. Columns 1–3: YSU; columns 4–6: MYJ; and columns 7–9: ACM2 PBL schemes. Columns 1, 4, and 7: WSM6; columns 2, 5, and 8: LIN; and columns 3, 6, and 9: Morrison MP scheme.</p>
Full article ">Figure 7
<p>Comparison of observed and simulated daily precipitation (mm) during the summer of 2002 at selected weather stations: (<b>a</b>) Addis Ababa (Bole) (central Ethiopia: 8.985<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> N, 38.786<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> E), (<b>b</b>) Dangila (northwestern Ethiopia: 11.434<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> N, 36.846<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> E), (<b>c</b>) Mekele (northern Ethiopia: 13.467<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> N, 39.533<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> E), (<b>d</b>) Dire Dawa (eastern Ethiopia: 9.625<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> N, 41.854<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> E), (<b>e</b>) Asaita (northeastern Ethiopia: 11.53<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> N, 41.53<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> E), and (<b>f</b>) Bure (western Ethiopia: 8.233<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> N, 35.1<math display="inline"><semantics> <msup> <mo> </mo> <mo>∘</mo> </msup> </semantics></math> E). Observed precipitation is shown in black, with simulation using KF CU (E1–E9) in blue, BMJ CU (E10–E18) in green, and Grell CU in red).</p>
Full article ">Figure 8
<p>Pearson correlation coefficient between observed and simulated daily precipitation during the summer of 2002 for 27 WRF simulations. Y-axis labels show cumulus, PBL, and microphysics schemes for each of the 27 experiments. Red dots indicate the location of weather stations and yellow boxes indicate the weather stations in each geographic group shown as G1 to G7 on the x-axis.</p>
Full article ">Figure 9
<p>Mean JJA rainfall (mm/day): (<b>a</b>) observed (weather station), (<b>b</b>) E1 (LW/SW: RRTM/Dudhai), (<b>c</b>) E28 (LW/SW: RRTMG/Dudhia); (<b>d</b>) E29 (LW/SW: RRTMG/RRTM); and (<b>e</b>) E30 (LW/SW: RRTMG/RRTMG).</p>
Full article ">Figure 10
<p>Mean JJA rainfall (mm/day): (<b>a</b>) CHIRPS, (<b>b</b>) ENACT, (<b>c</b>) E1 (LW/SW: RRTM/Dudhai), (<b>d</b>) E28 (LW/SW: RRTMG/Dudhia); (<b>e</b>) E29 (LW/SW: RRTMG/RRTM); and (<b>f</b>) E30 (LW/SW: RRTMG/RRTMG).</p>
Full article ">Figure 11
<p>Rank category based on aggregate score (AS).</p>
Full article ">Figure A1
<p>(<b>a</b>) Annual rainfall climatology: percent contribution of (<b>b</b>) March–May; (<b>c</b>) June–August; (<b>d</b>) September–November; and (<b>e</b>) December–February seasons to annual rainfall. Climatology based on 35 years of CHIRPS data.</p>
Full article ">Figure A2
<p>Bias in WRF-simulated mean summer (JJA) rainfall (mm/day) in 2002. Simulations in rows 1–3 used KF, BMJ, and Grell CU schemes, respectively. Columns 1–3, 4–6, and 7–9 used YSU, MYJ, and ACM2 PBL schemes, respectively. Columns 1, 4, and 7 used the WSM6 scheme; 2, 5, and 8 the LIN scheme; and 3, 6, and 9 the Morrison MP scheme (<a href="#atmosphere-15-00974-f002" class="html-fig">Figure 2</a>).</p>
Full article ">
20 pages, 1014 KiB  
Article
Temporal Distribution of Extreme Precipitation in Barcelona (Spain) under Multi-Fractal n-Index with Breaking Point
by Benoît Gacon, David Santuy and Darío Redolat
Atmosphere 2024, 15(7), 804; https://doi.org/10.3390/atmos15070804 - 4 Jul 2024
Viewed by 569
Abstract
Rainfall regimes are experiencing variations due to climate change, and these variations are adequately simulated by Earth System Models at a daily scale for most regions. However, there are not enough raw outputs to study extreme and sub-daily precipitation patterns on a local [...] Read more.
Rainfall regimes are experiencing variations due to climate change, and these variations are adequately simulated by Earth System Models at a daily scale for most regions. However, there are not enough raw outputs to study extreme and sub-daily precipitation patterns on a local scale. To address this challenge, Monjo developed the n-index by characterizing the intensity and concentration of precipitation based on mono-fractal theory. In this study, we explore the use of a multi-fractal approach to establish a more accurate method of time scaling useful to study extreme precipitation events at a finer temporal resolution. This study was carried out on the reference station of Barcelona (Spain) and its surroundings in order to be representative of the Mediterranean climate. For return periods between 2 and 50 years, two variables were analyzed: the n-index and the reference intensity I0. Moreover, a new parameter, the so-called “breaking point”, was designed here to describe the reference intensity I0, which is predominant for low time ranges. The results showed that both parameters are dependent on the time steps and the return period, and the scores confirmed the validity of our approach. Finally, the n-index was projected under downscaled CMIP6 climate scenarios by 2100, showing a sustained increase of up to +10%. Full article
(This article belongs to the Special Issue Geometry in Meteorology and Climatology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Example of mono-fractal IDF curves for Barcelona (Fabra Observatory), as considered in the ICARIA project [<a href="#B14-atmosphere-15-00804" class="html-bibr">14</a>]. Each colored line corresponds to a different return period (5, 10, 20, 30, 40, 50 and 100 years). The shared area indicates a 25–75th confidence level, calculated by an ensemble of 10 downscaled CMIP6 models for the historical period.</p>
Full article ">Figure 2
<p>Study area of the present work: (<b>a</b>) Location of the four observatories analyzed within the Barcelona Metropolitan Area (AMB; purple), Fabra Observatory is represented by a triangle and the other three by circles (yellow). (<b>b</b>) Position of Catalonia (yellow) and AMB (purple) in Europe.</p>
Full article ">Figure 3
<p>Example of multi-fractal <span class="html-italic">n</span>-index obtained from different time resolutions: (<b>a</b>) Hyetograph of a precipitation event of two hours gauged with four time steps (5 min, 10 min, 15 min and 20 min, sorted from the top to the bottom); (<b>b</b>) Maximum cumulative precipitation <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for a given duration <span class="html-italic">t</span> compared to ideal curves of <span class="html-italic">n</span>-index (gray dashed lines); (<b>c</b>) maximum averaged intensities (MAI) related to the previous <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> according to Equation (<a href="#FD1-atmosphere-15-00804" class="html-disp-formula">1</a>) and fitted curves of <span class="html-italic">n</span>-index (Equation (<a href="#FD2-atmosphere-15-00804" class="html-disp-formula">2</a>), red lines); (<b>d</b>) variation of the <span class="html-italic">n</span>-index as a function of the time step, compared to the two-parameter model (Equation (<a href="#FD6-atmosphere-15-00804" class="html-disp-formula">6</a>), green line) and the three-parameter model (Equation (<a href="#FD7-atmosphere-15-00804" class="html-disp-formula">7</a>), cyan line). Confidence level at 95% is represented by dashed lines.</p>
Full article ">Figure 4
<p>Regressions of the <span class="html-italic">n</span>-index according to the time step <span class="html-italic">t</span> and the return period <span class="html-italic">p</span> (color). The value of each parameter, the <span class="html-italic">p</span>-value for the linear method and the 95% confidence interval are detailed in <a href="#secAdot1-atmosphere-15-00804" class="html-sec">Appendix A.1</a>: (<b>a</b>) Linear Regression using the formula <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>∼</mo> <mn>1</mn> <mo>/</mo> <mo form="prefix">log</mo> <mfenced open="(" close=")"> <mi>t</mi> </mfenced> </mrow> </semantics></math> (Equation (<a href="#FD6-atmosphere-15-00804" class="html-disp-formula">6</a>)) on every time step; (<b>b</b>) Three-parametric regression model (Equation (<a href="#FD7-atmosphere-15-00804" class="html-disp-formula">7</a>)) fitted to the first four values before the breaking point (60 min); (<b>c</b>) Three-parametric regression model (Equation (<a href="#FD7-atmosphere-15-00804" class="html-disp-formula">7</a>)) fitted to all the values.</p>
Full article ">Figure 5
<p>Bias between the three regression models and the observed values for the <span class="html-italic">n</span>-index where colors represent return periods (as in <a href="#atmosphere-15-00804-f004" class="html-fig">Figure 4</a>): (<b>a</b>) Two-parameter <span class="html-italic">n</span>-index regression on the full period; (<b>b</b>) Two-parameter <span class="html-italic">n</span>-index regression for time steps lower than one hour; (<b>c</b>) Three-parameter <span class="html-italic">n</span>-index regression on the full period.</p>
Full article ">Figure 6
<p>Two different regressions to approximate the relation between time steps and <math display="inline"><semantics> <msub> <mi>I</mi> <mn>0</mn> </msub> </semantics></math> according to the return period. The value of each parameter, the <span class="html-italic">p</span>-value for the linear method and the 95% confidence intervals for the non-linear model are detailed in <a href="#secAdot2-atmosphere-15-00804" class="html-sec">Appendix A.2</a>: (<b>a</b>) Linear regression following the formula <math display="inline"><semantics> <mrow> <mo form="prefix">log</mo> <mfenced separators="" open="(" close=")"> <msub> <mi>I</mi> <mn>0</mn> </msub> </mfenced> <mo>∼</mo> <mo form="prefix">log</mo> <mfenced open="(" close=")"> <mi>t</mi> </mfenced> </mrow> </semantics></math> as Equation (<a href="#FD8-atmosphere-15-00804" class="html-disp-formula">8</a>); (<b>b</b>) Regression of the parameters <math display="inline"><semantics> <msub> <mi>I</mi> <mn>0</mn> </msub> </semantics></math> by using three parameters with Equation (<a href="#FD9-atmosphere-15-00804" class="html-disp-formula">9</a>); and their respective bias (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 7
<p>Historical evolution and climate projections for different climate change scenarios (colors) of annual mean <span class="html-italic">n</span>-index (lines) and percentiles p(30,70) (coloured areas). Projections are calculated as the ten-year moving averages of the outputs obtained through the application of the downscalling method to the models mentioned in <a href="#atmosphere-15-00804-t002" class="html-table">Table 2</a> on the observatories UF, UG and VT according to <a href="#atmosphere-15-00804-t001" class="html-table">Table 1</a>.</p>
Full article ">
26 pages, 8941 KiB  
Article
Impact of Climate Change on the Water Balance of the Akaki Catchment
by Alemayehu Kabeta Guyasa, Yiqing Guan and Danrong Zhang
Water 2024, 16(1), 54; https://doi.org/10.3390/w16010054 - 22 Dec 2023
Cited by 3 | Viewed by 1765
Abstract
Climate change has an impact on water resources. Estimations of the variations in water balance under climate change variables are essential for managing and developing the water resource of a catchment. The current investigation evaluated the magnitude of the change in the water [...] Read more.
Climate change has an impact on water resources. Estimations of the variations in water balance under climate change variables are essential for managing and developing the water resource of a catchment. The current investigation evaluated the magnitude of the change in the water balance component of the Akaki catchment, Ethiopia, using the semi-distributed hydrological model, the Soil and Water Assessment Tool (SWAT), with the integration of the Coordinated Regional Downscaling Experiment of Africa under RCP4.5 and 8.5. The SWAT model was developed using spatial and temporal data; it was calibrated (1991–2001) and validated (2002–2004) using SWAT-CUP. The statistical monthly SWAT model performance values of the NSE, PBIAS (%), and R2 showed good agreement between calibration and validation. On an annual basis, projected rainfall is expected to increase by 14.96%, 4.13%, 8.39%, and 10.39% in the 2040s under RCP4.5 and 8.5 and in the 2060s under RCP4.5 and 8.5, respectively, with inconsistent change on a monthly projections basis for each scenario. The projected monthly and yearly temperatures are expected to increase under different climate change scenarios. Annual evapotranspiration and potential evapotranspiration increased under both RCPs, whereas surface runoff, lateral flow, and water yield declined under the climate scenarios of each RCP. Monthly projected water yield showed a non-uniform change in the first 30 years and in the second years under the RCP4.5 and RCP8.5 scenarios. These results show that the catchment is highly vulnerable to hydrological and agricultural drought due to water availability. These research findings provide valuable evidence on the role of climate change in water balance, which will help decision makers to achieve better water resource management. Full article
(This article belongs to the Section Water and Climate Change)
Show Figures

Figure 1

Figure 1
<p>Map of the study area.</p>
Full article ">Figure 2
<p>Land use/cover.</p>
Full article ">Figure 3
<p>Soil type.</p>
Full article ">Figure 4
<p>Flowchart framework of study.</p>
Full article ">Figure 5
<p>Observed versus simulated for calibration from 1991 to 2001.</p>
Full article ">Figure 6
<p>Observed versus simulated for validation from 2002 to 2004.</p>
Full article ">Figure 7
<p>Projected rainfall under different climate scenarios.</p>
Full article ">Figure 8
<p>Projected monthly maximum temperature under different climate scenarios.</p>
Full article ">Figure 9
<p>Projected monthly minimum temperature under different climate scenarios.</p>
Full article ">Figure 10
<p>Projected evapotranspiration under different climate scenarios.</p>
Full article ">Figure 11
<p>Projected potential evapotranspiration under different climate scenarios.</p>
Full article ">Figure 12
<p>Projected surface runoff under different climate scenarios.</p>
Full article ">Figure 13
<p>Projected water yield under different climate scenarios.</p>
Full article ">Figure 14
<p>Projected lateral flow under different climate scenarios.</p>
Full article ">Figure 15
<p>Projected monthly evapotranspiration under different climate scenarios.</p>
Full article ">Figure 16
<p>Projected monthly potential evapotranspiration under climate scenarios.</p>
Full article ">Figure 17
<p>Projected monthly surface runoff under different climate scenarios.</p>
Full article ">Figure 18
<p>Projected monthly water yield under different climate scenarios.</p>
Full article ">Figure 19
<p>Projected monthly lateral flow under different climate scenarios.</p>
Full article ">
12 pages, 3552 KiB  
Article
Climate Change Flood Risk Analysis: Application of Dynamical Downscaling and Hydrological Modeling
by Fernando Neves Lima, Ana Carolina Vasques Freitas and Josiano Silva
Atmosphere 2023, 14(7), 1069; https://doi.org/10.3390/atmos14071069 - 25 Jun 2023
Cited by 1 | Viewed by 1452
Abstract
Floods are a recurring natural phenomenon during the rainy season in many Brazilian municipalities. Nevertheless, shifts in weather patterns have contributed to an increased incidence of these events in urban areas, where their impact can be amplified by the way the surrounding catchment [...] Read more.
Floods are a recurring natural phenomenon during the rainy season in many Brazilian municipalities. Nevertheless, shifts in weather patterns have contributed to an increased incidence of these events in urban areas, where their impact can be amplified by the way the surrounding catchment is occupied. Hence, the present study sought to evaluate the susceptibility of the urban drainage infrastructure in João Monlevade, Brazil, to the effects of climate change by undertaking a comprehensive assessment of the Carneirinhos catchment, including its morphometric characteristics. For this purpose, we employed a hydrological model driven by regional rainfall projections from a high-resolution climate model (HadGEM2-ES downscaled to 5 km resolution) under the Representative Concentration Pathways (RCP8.5) scenario. Several combinations of rainfall occurrence were simulated, incorporating temporal aspects (different durations and return times), as well as spatial aspects (concentrated and distributed rainfall within the catchment). The results showed that the area of exposed soil in the Carneirinhos catchment experienced an increase of more than 140% from 2016 to 2019, favoring runoff surface and evaporation, which contributed to the increase in the number of flood events in the region. In addition, only 1 of the 56 heavy rainfall event simulations performed did not exceed the capacity of the macro drainage gallery. Full article
(This article belongs to the Special Issue Climate Extremes and Their Impacts)
Show Figures

Figure 1

Figure 1
<p>Carneirinhos catchment location (<b>right panel</b>) relative to Brazil (<b>top left</b>) and Piracicaba River Basin (<b>bottom left</b>).</p>
Full article ">Figure 2
<p>Carneirinhos creek before canalization (<b>A</b>) and channel beginning efforts in the early 1970s (<b>B</b>). Source: Water distribution department of João Monlevade dataset.</p>
Full article ">Figure 3
<p>Catchment subdivision proposed.</p>
Full article ">Figure 4
<p>Methodology flowchart.</p>
Full article ">Figure 5
<p>Catchment land use and occupation comparison.</p>
Full article ">Figure 6
<p>Quantile plot of observed annual maximum daily precipitation with Gamma distribution fit.</p>
Full article ">Figure 7
<p>Disaggregation and rearrangement for 5-year return period.</p>
Full article ">Figure 8
<p>Comparison between hydrographs generated considering: (<b>a</b>) rainfall located at all catchments with simulated rainfall from the Eta/HADGEM2-ES model in the RCP8.5 scenario; (<b>b</b>) rainfall located at all catchments with the observed rainfall data.</p>
Full article ">
8 pages, 805 KiB  
Proceeding Paper
Downscaling of Satellite Rainfall Data Using Remotely Sensed NDVI and Topographic Datasets
by Zarina Yasmeen, Muhammad Jehanzeb Masud Cheema, Saddam Hussain, Zainab Haroon, Sadaf Amin and Muhammad Sohail Waqas
Environ. Sci. Proc. 2022, 23(1), 40; https://doi.org/10.3390/environsciproc2022023040 - 6 Feb 2023
Cited by 1 | Viewed by 1119
Abstract
Rainfall is a key factor in hydrological, meteorological, and water management applications in restricted regions or basins, but its measurement remains difficult in mountainous or otherwise remote places due to a lack of readily available rain gauges. While satellite rainfall data offer a [...] Read more.
Rainfall is a key factor in hydrological, meteorological, and water management applications in restricted regions or basins, but its measurement remains difficult in mountainous or otherwise remote places due to a lack of readily available rain gauges. While satellite rainfall data offer a better temporal resolution than other sources, the majority of this data are only available at a coarse geographic resolution, which distorts the true picture of precipitation. Thus, researchers at the University of Agriculture in Faisalabad used the normalized difference vegetation index (NDVI) monthly data and 1 km topography data for the whole Indus Basin from 2002 to 2011 to reduce the TRMM’s spatial resolution from 25 km to 1 km. An approach to downscaling based on a regression model with residual correction was established in this study. First, we resampled the NDVI and TRMM datasets to a 25 km resolution and established a regression model connecting the two datasets. Precipitation was forecasted at a distance of 25 km. The TRMM 3B43 product was then adjusted downward by the projected precipitation to achieve the residual value. The IDW method was used to reduce the resolution of the residual image from 25 km to 1 km. Rainfall was predicted using a regression model applied to NDVI at a 1 km spatial resolution. The final downscaled precipitation was created by combining the modeled precipitation at 1 km resolution with the residual image. The result was double-checked by the post-processing steps of validation and calibration. Full article
Show Figures

Figure 1

Figure 1
<p>Relationship between NDVI and TRMM.</p>
Full article ">Figure 2
<p>TRMM rainfall at 1 km resolution.</p>
Full article ">Figure 3
<p>Correlation between downscaled TRMM base on NDVI and rain gauges.</p>
Full article ">
18 pages, 5006 KiB  
Article
Improved Daily Spatial Precipitation Estimation by Merging Multi-Source Precipitation Data Based on the Geographically Weighted Regression Method: A Case Study of Taihu Lake Basin, China
by Yi Pan, Qiqi Yuan, Jinsong Ma and Lachun Wang
Int. J. Environ. Res. Public Health 2022, 19(21), 13866; https://doi.org/10.3390/ijerph192113866 - 25 Oct 2022
Cited by 4 | Viewed by 1575
Abstract
Accurately estimating the spatial and temporal distribution of precipitation is crucial for hydrological modeling. However, precipitation products based on a single source have their advantages and disadvantages. How to effectively combine the advantages of different precipitation datasets has become an important topic in [...] Read more.
Accurately estimating the spatial and temporal distribution of precipitation is crucial for hydrological modeling. However, precipitation products based on a single source have their advantages and disadvantages. How to effectively combine the advantages of different precipitation datasets has become an important topic in developing high-quality precipitation products internationally in recent years. This paper uses the measured precipitation data of Multi-Source Weighted-Ensemble Precipitation (MSWEP) and in situ rainfall observation in the Taihu Lake Basin, as well as the longitude, latitude, elevation, slope, aspect, surface roughness, distance to the coastline, and land use and land cover data, and adopts a two-step method to achieve precipitation fusion: (1) downscaling the MSWEP source precipitation field using the bilinear interpolation method and (2) using the geographically weighted regression (GWR) method and tri-cube function weighting method to achieve fusion. Considering geographical and human activities factors, the spatial and temporal distribution of precipitation errors in MSWEP is detected. The fusion of MSWEP and gauge observation precipitation is realized. The results show that the method in this paper significantly improves the spatial resolution and accuracy of precipitation data in the Taihu Lake Basin. Full article
(This article belongs to the Special Issue Ecological Environment Assessment Based on Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Map of the study area and location of the rain gauge stations.</p>
Full article ">Figure 2
<p>(<bold>a</bold>) Elevation and (<bold>b</bold>) land use and land cover for the Taihu Lake Basin.</p>
Full article ">Figure 3
<p>Principle diagram of bilinear interpolation.</p>
Full article ">Figure 4
<p>Flowchart of the GWR-based two-step merging scheme proposed in this study.</p>
Full article ">Figure 5
<p>Variable importance plot.</p>
Full article ">Figure 6
<p>Cross-validation continuous statistics (<bold>a</bold>) CC, (<bold>b</bold>) rBIAS, and (<bold>c</bold>) RMSD on daily combined precipitation data using four weighting functions (Gaussian, exponential, bi-square, and tri-cube).</p>
Full article ">Figure 7
<p>Cross-validation categorical statistics (POD, FAR, FBI, and CSI) for daily combined precipitation data using two weighting functions (Gaussian and tri-cube) in the no-rain precipitation intensity class.</p>
Full article ">Figure 8
<p>KGE values for the fused precipitation product calculated at three spatial resolutions (0.1°, 0.01°, and 1 km).</p>
Full article ">Figure 9
<p>Scatter plots with color density showing rain gauge observation and the estimated precipitation by the GWR model for fusion in 2015.</p>
Full article ">Figure 10
<p>Time series of areal-average daily precipitation of gauge observation and estimated precipitation for 2015.</p>
Full article ">Figure 11
<p>Taylor diagrams for the daily precipitation of the gauge observation and the GWR model across the entire period and different seasons in 2015.</p>
Full article ">Figure 12
<p>KGE values for the daily precipitation of the GWR model at 70 gauge sites in 2015.</p>
Full article ">Figure 13
<p>Evaluation statistics for the categorical indices at five precipitation intensity classes (mm/d). From left to right and top to bottom: POD, FAR, FBI, and CSI. The dashed black line represents the optimal value of each index.</p>
Full article ">Figure 13 Cont.
<p>Evaluation statistics for the categorical indices at five precipitation intensity classes (mm/d). From left to right and top to bottom: POD, FAR, FBI, and CSI. The dashed black line represents the optimal value of each index.</p>
Full article ">
7 pages, 508 KiB  
Proceeding Paper
Stochastics Modelling of Rainfall Process in Asia Region: A Systematics Review
by Hilda Ayu Pratikasiwi, Elma Dwi Putri Sinaga, Hanny Nirwani, Milkah Royna, Perdinan and Akhmad Faqih
Environ. Sci. Proc. 2022, 19(1), 22; https://doi.org/10.3390/ecas2022-12816 - 14 Jul 2022
Viewed by 941
Abstract
In recent years, use of the stochastic model has been growing due to the high complexity and dynamics of the atmosphere, especially the rainfall process. Various concepts have been applied to rainfall modeling, ranging from simplistic approaches to more complex models. It is [...] Read more.
In recent years, use of the stochastic model has been growing due to the high complexity and dynamics of the atmosphere, especially the rainfall process. Various concepts have been applied to rainfall modeling, ranging from simplistic approaches to more complex models. It is important to understand different stochastic rainfall modeling approaches, as well as their advantages and limitations. This paper determines the development of the latest stochastic rainfall models in the Asia region, where different concepts of stochastic rainfall models were highlighted. It reviews the different methodologies used, including rainfall forecasting, spatio-temporal analysis, and extreme event. We select 30 articles from 1571 literature published between 2013–2022 from the Scopus database. The results show that the stochastic models often used in the literature consist of Markov chain, weather generator, probability distribution, ARIMA, and the Bayesian model. In the recent development in Asia, stochastic models in rainfall modeling research are widely used to generate the occurrence and amount of rainfall data, statistical downscaling, future rainfall trends, and estimation of extreme values. The difference in spatio-temporal, climate conditions, and the parameters model can cause the performance of each model to be different. Full article
(This article belongs to the Proceedings of The 5th International Electronic Conference on Atmospheric Sciences)
Show Figures

Figure 1

Figure 1
<p>Flowchart outlining protocol of review using PRISMA.</p>
Full article ">
17 pages, 2986 KiB  
Article
Identifying Characteristics of Guam’s Extreme Rainfalls Prior to Climate Change Assessment
by Myeong-Ho Yeo, James Pangelinan and Romina King
Water 2022, 14(10), 1578; https://doi.org/10.3390/w14101578 - 14 May 2022
Viewed by 2028
Abstract
Extreme rainfall and its consequential flooding account for a devastating amount of damage to the Pacific Islands. Having an improved understanding of extreme rainfall patterns can better inform stormwater managers about current and future flooding scenarios, so they can minimize potential damages and [...] Read more.
Extreme rainfall and its consequential flooding account for a devastating amount of damage to the Pacific Islands. Having an improved understanding of extreme rainfall patterns can better inform stormwater managers about current and future flooding scenarios, so they can minimize potential damages and disruptions. In this study, the scaling invariant properties of annual maximum precipitations (AMPs) are used for describing the regional patterns of extreme rainfalls over Guam. AMPs are calculated at seven stations in Guam and exhibit distinct simple scaling behavior for two different time frames: (1) from 15 min to 45 min; and (2) from 45 min to 24 h. With these two different behaviors, the conventional estimation methods for sub-hourly durations overestimate the frequencies at a site in which breakpoints are clearly observed, while the proposed Scaling Generalized Extreme Value (GEV) method, based on the Scaling Three-NCM (S3NCM) method, provides comparable estimates. A new regional extreme rainfall analysis approach based on scaling exponents is introduced in this study. Results show distinct extreme rainfall patterns over Guam. Moreover, the numerical and graphical analyses identify that a tropical cyclone may increase daily AMPs by 3%, on average. Full article
(This article belongs to the Special Issue Urbanization, Climate Change and Flood Risk Management)
Show Figures

Figure 1

Figure 1
<p>Scheme of this study.</p>
Full article ">Figure 2
<p>Geographical information on the study area. Upper map shows the location of Guam in the Mariana region, and the bottom map shows the digital elevation map. Black points denote the rain gages used in this study. Sources for World Ocean Reference and Bases in the upper map are Esri, GEBCO, NOAA, National Geographic, Garmin, HERE, Geonames.org, and other contributors.</p>
Full article ">Figure 3
<p>Quantiles estimated by conventional parameter estimation method (L-moments) and the proposed NCMs. (<b>A</b>,<b>B</b>) are quantiles plot for 15-min and 1-h durations for Almagosa, and (<b>C</b>,<b>D</b>) are those for Mt. Chachao.</p>
Full article ">Figure 4
<p>Scaling properties of AMPs. (<b>A</b>) is the NCM plot for Almagosa, (<b>B</b>) for Fena Lake, (<b>C</b>) for Geomag, (<b>D</b>) Mt. Chachao, (<b>E</b>) for Mt. Santa Rosa, (<b>F</b>) for Umatac, and (<b>G</b>) for Windward Hills, respectively.</p>
Full article ">Figure 5
<p>Test of Simple scaling. (<b>A</b>) is a plot the scaling exponent against order of NCM of AMPs for Almagosa, (<b>B</b>) for Fena Lake, (<b>C</b>) for Geomag, (<b>D</b>) Mt. Chachao, (<b>E</b>) for Mt. Santa Rosa, (<b>F</b>) for Umatac, and (<b>G</b>) for Windward Hills, respectively.</p>
Full article ">Figure 6
<p>IDF curves for comparing the estimation method of sub-hourly frequencies. Two stations (Mt Chachao and Mt Santo Rosa) are selected. Breakpoints are clearly observed at Mt Chachao (GA4) but not clear at Mt. Santa Rosa (GA5). (<b>A</b>,<b>C</b>,<b>E</b>) are IDF curves for the 15-min, 30-min, and 45-min durations for Mt. Chachao station, and (<b>B</b>,<b>D</b>,<b>F</b>) are those for Mt. Santa Rosa, respectively.</p>
Full article ">Figure 6 Cont.
<p>IDF curves for comparing the estimation method of sub-hourly frequencies. Two stations (Mt Chachao and Mt Santo Rosa) are selected. Breakpoints are clearly observed at Mt Chachao (GA4) but not clear at Mt. Santa Rosa (GA5). (<b>A</b>,<b>C</b>,<b>E</b>) are IDF curves for the 15-min, 30-min, and 45-min durations for Mt. Chachao station, and (<b>B</b>,<b>D</b>,<b>F</b>) are those for Mt. Santa Rosa, respectively.</p>
Full article ">Figure 7
<p>Interpolated scaling exponents for both short duration (<b>A</b>) and long duration (<b>B</b>).</p>
Full article ">Figure 8
<p>Estimated IDF curves using three temporal downscaling approaches. The 1-h IDF curves for Almagosa (<b>A</b>), Fena Lake (<b>B</b>), Geomag (<b>C</b>), Mt. Chachao (<b>D</b>), Mt. Santa Rosa (<b>E</b>), Umatac (<b>F</b>), and Windward Hills (<b>G</b>), respectively.</p>
Full article ">Figure 8 Cont.
<p>Estimated IDF curves using three temporal downscaling approaches. The 1-h IDF curves for Almagosa (<b>A</b>), Fena Lake (<b>B</b>), Geomag (<b>C</b>), Mt. Chachao (<b>D</b>), Mt. Santa Rosa (<b>E</b>), Umatac (<b>F</b>), and Windward Hills (<b>G</b>), respectively.</p>
Full article ">Figure 9
<p>Comparison of IDF curves for accounting for typhoon effects on IDF curves. The estimated IDF curves for Almagosa (<b>A</b>), Fena Lake (<b>B</b>), Geomag (<b>C</b>), Mt. Chachao (<b>D</b>), Mt. Santa Rosa (<b>E</b>), Umatac (<b>F</b>), and Windward Hills (<b>G</b>), respectively.</p>
Full article ">
20 pages, 5534 KiB  
Article
Assessment of Sponge City Flood Control Capacity According to Rainfall Pattern Using a Numerical Model after Muti-Source Validation
by Haichao Li, Hiroshi Ishidaira, Yanqi Wei, Kazuyoshi Souma and Jun Magome
Water 2022, 14(5), 769; https://doi.org/10.3390/w14050769 - 28 Feb 2022
Cited by 3 | Viewed by 3699
Abstract
Urban floods are a common urban disaster that threaten the economy and development of cities. Sponge cities can improve flood resistance ability and reduce floods by setting low-impact development measures (LID). Evaluating flood reduction benefits is the basic link in the construction of [...] Read more.
Urban floods are a common urban disaster that threaten the economy and development of cities. Sponge cities can improve flood resistance ability and reduce floods by setting low-impact development measures (LID). Evaluating flood reduction benefits is the basic link in the construction of sponge cities. Therefore, it is of great significance to evaluate the benefits of sponge cities from the perspective of different rain patterns. In this study, we investigated the urban runoff of various rainfall patterns in Mianyang city using the Strom Water Management Model (SWMM). We employed 2–100-year return periods and three different temporal rainfall downscaling methods to evaluate rain patterns and simulate urban runoff in Mianyang, with and without the implementation of sponge city measures. After calibration, model performance was validated using multi-source data concerning flood peaks and inter-annual variations in flood magnitude. Notably, the effects of peak rainfall patterns on historical floods were generally greater than the effects of synthetic rainfalls generated by temporal downscaling. Compared to the rainfall patterns of historical flood events, the flood protection capacities of sponge cities can be easily overestimated when using the synthetic rainfall patterns generated by temporal downscaling. Overall, an earlier flood peak was associated with better flood sponge city protection capacity. In this context, the results obtained in this study provide useful reference information about the impact of rainfall pattern on urban flood control by LID, and can be used for sponge city design in other part of China. Full article
(This article belongs to the Section Urban Water Management)
Show Figures

Figure 1

Figure 1
<p>Flowchart of assessment of Sponge City Flood Control capacity based on multi-source validation and depends on different rainfall patterns.</p>
Full article ">Figure 2
<p>Location map of study area: central part of Mianyang City.</p>
Full article ">Figure 3
<p>Schematics of water balance analysis [<a href="#B39-water-14-00769" class="html-bibr">39</a>].</p>
Full article ">Figure 4
<p>Chi-square distribution probability density function plot [<a href="#B33-water-14-00769" class="html-bibr">33</a>].</p>
Full article ">Figure 5
<p>Single peak Chicago design storm by different peak coefficient.</p>
Full article ">Figure 6
<p>Multi-peak Chicago design storm by different peak coefficient.</p>
Full article ">Figure 7
<p>Annual maximum flood events aNDFIm<sub>year</sub> and simulated discharge.</p>
Full article ">Figure 8
<p>Extreme and average rainfall patterns analysis.</p>
Full article ">Figure 9
<p>Extreme and average flood control analysis.</p>
Full article ">Figure 10
<p>Different peak timing rainfall patterns analysis.</p>
Full article ">Figure 11
<p>Different peak timing flood control analysis.</p>
Full article ">Figure 12
<p>Multi peak rainfall patterns analysis.</p>
Full article ">Figure 13
<p>Effect of multi peak flood control analysis.</p>
Full article ">
23 pages, 4959 KiB  
Article
Assessment of Satellite-Based Rainfall Products Using a X-Band Rain Radar Network in the Complex Terrain of the Ecuadorian Andes
by Nazli Turini, Boris Thies, Rütger Rollenbeck, Andreas Fries, Franz Pucha-Cofrep, Johanna Orellana-Alvear, Natalia Horna and Jörg Bendix
Atmosphere 2021, 12(12), 1678; https://doi.org/10.3390/atmos12121678 - 14 Dec 2021
Cited by 1 | Viewed by 2571
Abstract
Ground based rainfall information is hardly available in most high mountain areas of the world due to the remoteness and complex topography. Thus, proper understanding of spatio-temporal rainfall dynamics still remains a challenge in those areas. Satellite-based rainfall products may help if their [...] Read more.
Ground based rainfall information is hardly available in most high mountain areas of the world due to the remoteness and complex topography. Thus, proper understanding of spatio-temporal rainfall dynamics still remains a challenge in those areas. Satellite-based rainfall products may help if their rainfall assessment are of high quality. In this paper, microwave-based integrated multi-satellite retrieval for the Global Precipitation Measurement (GPM) (IMERG) (MW-based IMERG) was assessed along with the random-forest-based rainfall (RF-based rainfall) and infrared-only IMERG (IR-only IMERG) products against the quality-controlled rain radar network and meteorological stations of high temporal resolution over the Pacific coast and the Andes of Ecuador. The rain area delineation and rain estimation of each product were evaluated at a spatial resolution of 11 km2 and at the time of MW overpass from IMERG. The regionally calibrated RF-based rainfall at 2 km2 and 30 min was also investigated. The validation results indicate different essential aspects: (i) the best performance is provided by MW-based IMERG in the region at the time of MW overpass; (ii) RF-based rainfall shows better accuracy rather than the IR-only IMERG rainfall product. This confirms that applying multispectral IR data in retrieval can improve the estimation of rainfall compared with single-spectrum IR retrieval algorithms. (iii) All of the products are prone to low-intensity false alarms. (iv) The downscaling of higher-resolution products leads to lower product performance, despite regional calibration. The results show that more caution is needed when developing new algorithms for satellite-based, high-spatiotemporal-resolution rainfall products. The radar data validation shows better performance than meteorological stations because gauge data cannot correctly represent spatial rainfall in complex topography under convective rainfall environments. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The distribution of (<b>a</b>) meteorological stations (19 April 2017 to 28 February 2018) and spatial coverage of radars (GUAXX: 16 June 2017 to 1 February 2018; CAXX: 19th April 2017 to 1 July 2017) used in this study, (<b>b</b>) the radars in the study period (GUAXX: 16 June 2017 to 1 February 2018; CAXX: 19th April 2017 to 1 July 2017). For validation purposes, we excluded the radar data in the very near range (&lt;10 km distance from the radar site) to avoid contamination through noise. We also excluded the far range &gt;50 km due to possible attenuation errors. Nevertheless, we show the rainfall amount in the entire radar range for better illustration. The extent of study area is shown in windows (W)-1. (<b>c</b>) Spatial distribution of the elevation in the radar coverage area. W-2 and W-3 rectangles outline the extent of <a href="#atmosphere-12-01678-f002" class="html-fig">Figure 2</a>a,b.</p>
Full article ">Figure 2
<p>Location of pixels with a minimum number of three gauges for (<b>a</b>) the University of Cuenca gauge network and (<b>b</b>) the UTPL gauge network. In <a href="#atmosphere-12-01678-f001" class="html-fig">Figure 1</a>a, W-2 and W-3 rectangles outline the extent of (<b>a</b>) and (<b>b</b>), respectively.</p>
Full article ">Figure 3
<p>Schematic view of how H, M, and F were designated in the rain area validation. The dry pixels are shown in white, and the rainy pixels are shown in grey. The standard approach defines M (F) when a rainy pixel in the radar (satellite-based rainfall product) is related to a dry pixel in the satellite-based rainfall product (radar) at the same time. In the temporal event-based approach (fourth row), the M (F) in the vicinity time of hits are defined as a reduction (continuous) in the event duration. Thus, the terms Duration+ (Duration-) are described. True misses and true false alarms are the errors occurring simultaneously or in the same pixel, respectively [<a href="#B17-atmosphere-12-01678" class="html-bibr">17</a>].</p>
Full article ">Figure 4
<p>Schematic view of how hits, misses, and false alarms are designated in the rain area validation. The dry pixels are shown in white, and the rainy pixels are shown in grey. The standard approach defines M (F) when a rainy pixel in the radar (satellite-based rainfall product) is related to a dry pixel in the satellite-based rainfall product (radar) at the same time. In the spatial event-based approach (second row), the M (F) in the neighboring pixels are defined as a spatially drifted miss (false alarm) of the event. The errors simultaneously and in the same pixel are called true misses and false alarms, respectively.</p>
Full article ">Figure 5
<p>Standard cross-table approach for all available MW overpass times for the validation of rain area delineation for (<b>a</b>) IR-only IMERG, (<b>b</b>) RF-based rainfall, and (<b>c</b>) MW-based IMERG. Note that the correct negative fraction extends to 100%.</p>
Full article ">Figure 6
<p>Spatial distribution of the validation metrics for rain area delineation at the time of MW overpass. (<b>a</b>) POD, (<b>b</b>) FAR, and (<b>c</b>) HSS showing the matrics for MW. The variables were calculated for MW-based IMERG. (<b>d</b>) POD, (<b>e</b>) FAR, and (<b>f</b>) HSS illustrating the performance of IR-only IMERG. (<b>g</b>) POD, (<b>h</b>) FAR, and (<b>i</b>) HSS showing the RF-based rainfall performance. The variables were calculated for each grid point of the validation data set over the stated period. For better illustration, we show the results up to 75 km distance from the center of each radar.</p>
Full article ">Figure 7
<p>Boxplot of the validation metrics for rain area delineation over the MW overpass time. The performance of (<b>a</b>) MW-based IMERG, (<b>b</b>) IR-only IMERG, and (<b>c</b>) RF-based rainfall along elevation. Boxes show the 25th, 50th, and 75th percentiles. Whiskers extensions are to the maximum data value between the 75th and 25th percentiles. Diamonds indicate outliers.</p>
Full article ">Figure 8
<p>The rain area delineation performance over the MW overpass time and at 11 km<sup>2</sup> for different rainfall rates for (<b>a</b>) MW-based IMERG, (<b>b</b>) RF-based rainfall, and (<b>c</b>) IR-only IMERG.</p>
Full article ">Figure 9
<p>Comparison of rainfall rates estimated by the radar and satellite-based products. (<b>a</b>,<b>d</b>,<b>g</b>) Scatter plot with radar rainfall rates (x-axis) and microwave-based IMERG, IR-only IMERG, and RF-based rainfall rates (y-axis), respectively. Only pixels with hits are considered. The parameters <span class="html-italic">n</span> show the total number of hits. (<b>b</b>,<b>e</b>,<b>h</b>) Quantile–quantile (Q–Q) plot of the radar (x-axis) and microwave-based IMERG (y-axis), IR-only IMERG (y-axis), and RF-based (y-axis) rainfall rates. The 10th, 50th, and 90th percentiles are illustrated. (<b>c</b>,<b>f</b>,<b>i</b>) The distribution of cumulative rainfall rate for the contingency table of each satellite-based product. The radar rain rate is displayed in black as a reference.</p>
Full article ">Figure 10
<p>Boxplot of the validation metrics for rain estimation at the MW overpass time. The performance of (<b>a</b>) MW-based IMERG, (<b>b</b>) IR-only IMERG, and (<b>c</b>) RF-based rainfall are shown along with elevation. Boxes show the 25th, 50th, and 75th percentiles. Whiskers extensions are to the maximum data value between the 75th and 25th percentiles. Diamonds indicate outliers.</p>
Full article ">Figure 11
<p>(<b>a</b>) Standard contingency table approach for all available RF-based rainfall products for both radars at 2 km<sup>2</sup> and 30 min. Note that the correct negative fraction extends to 100%. (<b>b</b>,<b>c</b>) The temporal event-based approach of the contingency table was evaluated in the M and F subsets, respectively. (<b>d</b>,<b>e</b>) The spatial event-based approach of the contingency table was evaluated in the M and F subsets, respectively. The numbers in the bars show the percentage.</p>
Full article ">Figure 12
<p>Comparison of rain rates estimated by the radar and RF-based rainfall at 2 km<sup>2</sup> and 30 min. (<b>a</b>) Scatter plot with radar rainfall (x-axis) and RF-based rainfall (y-axis). Only the pixels with hit are considered. (<b>b</b>) Q–Q plot of radar (x-axis) and RF-based rainfall rates (y-axis). The 10th, 50th, and 90th percentiles are illustrate. (<b>c</b>) The distribution of cumulative rainfall rate for the contingency table. (<b>d</b>) The distribution of cumulative rainfall rate based on the event-based (spatial and temporal) contingency table.</p>
Full article ">Figure 13
<p>Comparison of the validation metrics between the radar and RF-based rainfall at 2 km<sup>2</sup> and 30 min, 1 h, 3 h, and daily. The performance of RF-based (<b>a</b>) rain area delineation and (<b>b</b>) rain estimation is shown for different temporal resolutions.</p>
Full article ">
14 pages, 3224 KiB  
Article
Time Varying Spatial Downscaling of Satellite-Based Drought Index
by Hone-Jay Chu, Regita Faridatunisa Wijayanti, Lalu Muhamad Jaelani and Hui-Ping Tsai
Remote Sens. 2021, 13(18), 3693; https://doi.org/10.3390/rs13183693 - 15 Sep 2021
Cited by 7 | Viewed by 2271
Abstract
Drought monitoring is essential to detect the presence of drought, and the comprehensive change of drought conditions on a regional or global scale. This study used satellite precipitation data from the Tropical Rainfall Measuring Mission (TRMM), but refined the data for drought monitoring [...] Read more.
Drought monitoring is essential to detect the presence of drought, and the comprehensive change of drought conditions on a regional or global scale. This study used satellite precipitation data from the Tropical Rainfall Measuring Mission (TRMM), but refined the data for drought monitoring in Java, Indonesia. Firstly, drought analysis was conducted to establish the standardized precipitation index (SPI) of TRMM data for different durations. Time varying SPI spatial downscaling was conducted by selecting the environmental variables, normalized difference vegetation index (NDVI), and land surface temperature (LST) that were highly correlated with precipitation because meteorological drought was associated with vegetation and land drought. This study used time-dependent spatial regression to build the relation among original SPI, auxiliary variables, i.e., NDVI and LST. Results indicated that spatial downscaling was better than nonspatial downscaling (overall RMSEs: 0.25 and 0.46 in spatial and nonspatial downscaling). Spatial downscaling was more suitable for heterogeneous SPI, particularly in the transition time (R: 0.863 and 0.137 in June 2019 for spatial and nonspatial models). The fine resolution (1 km) SPI can be composed of the environmental data. The fine-resolution SPI captured a similar trend of the original SPI. Furthermore, the detailed SPI maps can be used to understand the spatio-temporal pattern of drought severity. Full article
(This article belongs to the Special Issue Drought Monitoring Using Satellite Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Processing workflow for time varying spatial downscaling SPI maps: (1) pre-processing step; (2) TRMM SPI generation using standardized drought analysis; (3) nonspatial and spatial downscaling; (4) validation and comparison.</p>
Full article ">Figure 2
<p>Regional SPI time series from 2015–2019 TRMM monthly data in Java: (<b>a</b>) SPI 3, (<b>b</b>) SPI 6, and (<b>c</b>) SPI 9.</p>
Full article ">Figure 3
<p>(<b>a</b>) RMSE boxplots comparing spatial and nonspatial downscaling of SPI 6 in 2018 and 2019 (during drought periods). (<b>b</b>) RMSE time series comparing spatial and nonspatial downscaling of SPI 6 in 2019.</p>
Full article ">Figure 4
<p>Estimation and observation plots of spatial (<b>a</b>,<b>c</b>) and nonspatial (<b>b</b>,<b>d</b>) downscaling of SPI 6 in June (<b>a</b>,<b>b</b>) and December (<b>c</b>,<b>d</b>) 2019.</p>
Full article ">Figure 5
<p>Spatial patterns of original SPI 6 (<b>a</b>,<b>d</b>), spatial downscaling (<b>b</b>,<b>e</b>) and nonspatial downscaling (<b>c</b>,<b>f</b>) results in June (<b>a</b>–<b>c</b>) and December (<b>d</b>–<b>f</b>) 2019.</p>
Full article ">Figure 6
<p>Average (<b>a</b>,<b>b</b>) and standard derivation (<b>c</b>,<b>d</b>) maps of SPI 6 in 2018–2019 when compared with original data (<b>a</b>,<b>c</b>) and spatial downscaling results (<b>b</b>,<b>d</b>).</p>
Full article ">Figure 7
<p>Monthly fine-resolution SPI 6 maps from January to December (<b>a</b>–<b>l</b>), 2019.</p>
Full article ">
21 pages, 6319 KiB  
Article
Assessing Future Rainfall Intensity–Duration–Frequency Characteristics across Taiwan Using the k-Nearest Neighbor Method
by Pei-Yuan Chen, Ching-Pin Tung, Jung-Hsuan Tsao and Chia-Jeng Chen
Water 2021, 13(11), 1521; https://doi.org/10.3390/w13111521 - 28 May 2021
Cited by 3 | Viewed by 3293
Abstract
This study analyzes the changes in rainfall intensities across Taiwan using the k-Nearest Neighbor method. Biases are corrected according to the identified discrepancy between the probability distribution of the model run and that of the observed data in the historical period. The [...] Read more.
This study analyzes the changes in rainfall intensities across Taiwan using the k-Nearest Neighbor method. Biases are corrected according to the identified discrepancy between the probability distribution of the model run and that of the observed data in the historical period. The projections of 21 weather stations in Taiwan under 10 (2RCP × 5GCM) scenarios for the near-(2021–2040) and far-future (2081–2100) are derived. The frequently occurred short-duration storm events in some regions decrease, but they are still vulnerable to flood considering the existing drainage capacities. More specifically, the land-subsidence region in the central, the landslide-sensitive mountainous region in the north and central, the pluvial- and fluvial-flood prone region in the north, and the eastern regions with vulnerable infrastructures should be especially aware of long-duration extreme events. Associations of the rainfall intensity with the different return period as well as the duration are further analyzed. The short-duration extreme events will become stronger, especially for 1-h events in the northern region and 1 or 2-h events in both the southern and eastern regions. In addition, places without experiences of long-lasting events may experience rainfall amounts exceeding 500 mm should be alert. Adaptation measures such as establishing distributed drainage system or renewing hydrological infrastructures in the eastern region are suggested considering the near future projection, and in the central and the southern regions for far future as well. Our findings can assist adaptation-related decision-making for more detailed stormwater/water resource management. Full article
(This article belongs to the Special Issue Water Resources Vulnerability and Resilience in a Changing Climate)
Show Figures

Figure 1

Figure 1
<p>Location of 21 weather stations across Taiwan for the northern, central, southern and eastern regions.</p>
Full article ">Figure 2
<p>Relation between the IDF values in the historical period (1986–2005) used to verify the reliability of the K-NN method, derived based on (i) the observed daily rainfall, (ii) the projected daily rainfall, and (iii) the observed hourly rainfall.</p>
Full article ">Figure 3
<p>Intensity-Duration-Frequency (IDF) curves from the k-NN method using both the simulated and observed daily precipitation of Tainan in the historical period (1986–2005).</p>
Full article ">Figure 4
<p>Rainfall intensities (<b>top</b>) and percentage changes (<b>bottom</b>) of events with 1-h duration and 2-yr return period (events<sub>short</sub>) for the near future (2021–2040) under RCP 2.6 and RCP 8.5 for five GCMs (CCSM4, CESM1-CAM5, GISS-E2-R, HadGEM2-AO, MIROC5).</p>
Full article ">Figure 5
<p>Rainfall intensities (<b>top</b>) and percentage of changes (<b>bottom</b>) of events with one-hour duration and two-year return period (events<sub>short</sub>) for far future (2081–2100) under RCP2.6 and RCP 8.5 for five GCMs (CCSM4, CESM1-CAM5, GISS-E2-R, HadGEM2-AO, MIROC5).</p>
Full article ">Figure 6
<p>Difference of rainfall intensities (<b>top</b>) and percentage changes (<b>bottom</b>) of events<sub>short</sub> between the near future (2021–2040) and far future (2081–2100).</p>
Full article ">Figure 7
<p>Rainfall intensities (<b>top</b>) and percentage changes (<b>bottom</b>) of events with 24-h duration and 25-yr return period (events<sub>long</sub>) for the near future (2021–2040) under RCP 2.6 and RCP 8.5 for five GCMs (CCSM4, CESM1-CAM5, GISS-E2-R, HadGEM2-AO, MIROC5).</p>
Full article ">Figure 8
<p>Rainfall intensities (<b>top</b>) and percentage of changes (<b>bottom</b>) of events with twenty-four-hour du-ration and twenty-five-year return period (events<sub>long</sub>) for far future (2081–2100) under RCP2.6 and RCP 8.5 for five GCMs (CCSM4, CESM1-CAM5, GISS-E2-R, HadGEM2-AO, MIROC5).</p>
Full article ">Figure 9
<p>Difference of rainfall intensities (<b>top</b>) and percentage changes (<b>bottom</b>) of events<sub>long</sub> between the near future (2021–2040) and far future (2081–2100).</p>
Full article ">Figure 10
<p>Rainfall intensities of the 1-h events (D1) in different return periods for the four regions and four scenarios.</p>
Full article ">Figure 11
<p>Rainfall intensities of events with twenty-four-hour duration (D24) changing with the return periods for the four regions and four scenarios.</p>
Full article ">Figure 12
<p>Rainfall intensities of 2-yr events (T2) of different durations for the four regions and four scenarios.</p>
Full article ">Figure 13
<p>Rainfall intensities of events with twenty-five-year return period (T25) changing with the durations for the four regions and four scenarios.</p>
Full article ">
19 pages, 3809 KiB  
Article
Evaluation of Sixteen Gridded Precipitation Datasets over the Caribbean Region Using Gauge Observations
by Abel Centella-Artola, Arnoldo Bezanilla-Morlot, Michael A. Taylor, Dimitris A. Herrera, Daniel Martinez-Castro, Isabelle Gouirand, Maibys Sierra-Lorenzo, Alejandro Vichot-Llano, Tannecia Stephenson, Cecilia Fonseca, Jayaka Campbell and Milena Alpizar
Atmosphere 2020, 11(12), 1334; https://doi.org/10.3390/atmos11121334 - 9 Dec 2020
Cited by 21 | Viewed by 3470
Abstract
The existence of several gridded precipitation products (GPP) has facilitated studies related to climate change, climate modeling, as well as a better understanding of the physical processes underpinning this key variable. Due to complexities in estimating rainfall, gridded datasets exhibit different levels of [...] Read more.
The existence of several gridded precipitation products (GPP) has facilitated studies related to climate change, climate modeling, as well as a better understanding of the physical processes underpinning this key variable. Due to complexities in estimating rainfall, gridded datasets exhibit different levels of accuracy across regions, even when they are developed at relatively high resolution or using sophisticated procedures. The performance of 16 GPP are evaluated over the Caribbean region, which includes the Caribbean Islands, and portions of Central South America. Monthly data for sixty weather stations are used as a reference for the period 1983–2010. The 16 GPP include six products based on station data only, two that combine ground station and satellite information, two merging station and reanalysis information, four based on reanalysis, and two using multisource information. The temporal resolution of the GPP ranged between daily and monthly and spatial resolution from 0.033° to 0.5°. The methodological approach employed combined a comparison of regional and sub-regional precipitation annual cycles, the Kling–Gupta efficiency (KGE) index, as well as several metrics derived from the standardized precipitation index (SPI). Overall, the best performances were obtained from GPCC025 and MSWEP2, likely reflecting the positive impact of the large number of station data utilized in their development. It is also demonstrated that a higher spatial resolution does not always mean better accuracy. There is a need for this kind of assessment when undertaking climate studies in regions like the Caribbean where resolution is a significant consideration. ERA5 performed best among the reanalyses analyzed and has the potential to be used to develop regionally based GPP by applying bias correction or downscaling techniques. The methodological approach employed provides a comprehensive and robust evaluation of the relative strengths and weaknesses of GPP in the Caribbean region. Full article
(This article belongs to the Special Issue Central America and Caribbean Hydrometeorology and Hydroclimate)
Show Figures

Figure 1

Figure 1
<p>Study area and the location of the rain-gauge network. (<b>a</b>) Caribbean region, (<b>b</b>) the western Caribbean sub-region, and (<b>c</b>) the eastern Caribbean sub-region.</p>
Full article ">Figure 2
<p>Annual cycle of observations and GPP, averaged over the region (<b>a</b>), western subregion (<b>b</b>) and eastern subregion (<b>c</b>) defined in <a href="#atmosphere-11-01334-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Boxplot of the <span class="html-italic">KGE</span> index and its components <span class="html-italic">B</span> (bias), <span class="html-italic">R</span> (correlation) and <span class="html-italic">V</span> (variability). The names of the GPP are colored according to data source considered in its development (red: stations; brown: stations and reanalysis; blue: stations and satellites; orange: reanalysis and green: multisources). The GPP have been sorted according to the median value of KGE index. In all the cases the optimal value is 1.0.</p>
Full article ">Figure 4
<p>As in <a href="#atmosphere-11-01334-f003" class="html-fig">Figure 3</a> but just for KGE index for dry (left) and wet (right) seasons.</p>
Full article ">Figure 5
<p>Spatial distribution of KGE scores for GPCC025, MSWEP2, GPCC05, MSWEP, and CHELSA. Colors for KGE range from violet to red, representing very poor to optimum skill respectively. Diagrams ordered according to median annual performance (see <a href="#atmosphere-11-01334-f003" class="html-fig">Figure 3</a>).</p>
Full article ">Figure 6
<p>Correlation between full data GPCCversion 8 and version 7 over the period 1983–2000. The products were used with a native spatial resolution of 1° × 1°.</p>
Full article ">Figure 7
<p>Differences between observation and the GPP for drought statistics derived from SPI-0.5 (left column) and SPI-1.5 (right column). Drought Frequency (DF) and Number of Drought Events (DE) are shown in panels (<b>a</b>,<b>b</b>), and (<b>c</b>,<b>d</b>), respectively. Bars show values for SPI3 (blue), SPI6 SPI6 (red) and SPI12 (yellow). In the case of DF (<b>a</b>,<b>b</b>) the differences are represented as percentage. (red) and SPI12 (yellow). In the case of DF (<b>a</b>,<b>b</b>) the differences are represented as percentage.</p>
Full article ">Figure 8
<p>Taylor diagrams for the areas with SPI lower than −0.5 (<b>a</b>), moderate and severe droughts (<b>b</b>,<b>c</b>, respectively). SPI greater than 0.5 as well as moderate and severe wet conditions are shown in (<b>d</b>–<b>f</b>), respectively.</p>
Full article ">
15 pages, 7417 KiB  
Article
An Ensemble Climate-Hydrology Modeling System for Long-Term Streamflow Assessment in a Cold-Arid Watershed
by Jie Sun, Yongping Li, Jiansen Wu and Hongyu Zhang
Water 2020, 12(8), 2293; https://doi.org/10.3390/w12082293 - 14 Aug 2020
Cited by 2 | Viewed by 2745
Abstract
Climate change can bring about substantial alternatives of temperature and precipitation in the spatial and temporal patterns. These alternatives would impact the hydrological cycle and cause flood or drought events. This study has developed an ensemble climate-hydrology modeling system (ECHMS) for long-term streamflow [...] Read more.
Climate change can bring about substantial alternatives of temperature and precipitation in the spatial and temporal patterns. These alternatives would impact the hydrological cycle and cause flood or drought events. This study has developed an ensemble climate-hydrology modeling system (ECHMS) for long-term streamflow assessment under changing climate. ECHMS consists of multiple climate scenarios (two global climate models (GCMs) and four representative concentration pathways (RCPs) emission scenarios), a stepwise-cluster downscaling method and semi-distributed land use-based runoff process (SLURP) model. ECHMS is able to reflect the uncertainties in climate scenarios, tackle the complex relationships (e.g., nonlinear/linear, discrete/continuous) between climate predictors and predictions without functional assumption, and capture the combination of snowmelt– and rainfall–runoff process with a simplicity of operation. Then, the developed ECHMS is applied to Kaidu watershed for analyzing the changes of streamflow during the 21st century. Results show that by 2099, the temperature increment in Kaidu watershed is mainly contributed by the warming in winter and spring. The precipitation will increase obviously in spring and autumn and decrease in winter. Multi-year average streamflow would range from 105.6 to 113.8 m3/s across all scenarios during the 21st century with an overall increasing trend. The maximum average increasing rate is 2.43 m3/s per decade in October and the minimum is 0.26 m3/s per decade in January. Streamflow change in spring is more sensitive to climate change due to its complex runoff generation process. The obtained results can effectively identify future streamflow changing trends and help manage water resources for decision makers. Full article
(This article belongs to the Special Issue Climate Change Impacts on Water Resources)
Show Figures

Figure 1

Figure 1
<p>The framework of the ensemble climate–hydrology modeling system (the flowchart of semi-distributed land use-based runoff process (SLURP) is derived from Sun et al. [<a href="#B24-water-12-02293" class="html-bibr">24</a>]).</p>
Full article ">Figure 2
<p>Kaidu watershed.</p>
Full article ">Figure 3
<p>Temperature under different scenarios [(<b>a1</b>)–(<b>a4</b>) are the annual temperature during the period 2010–2099 and (<b>b1</b>)–(<b>b4</b>) are the monthly average changes in 2080s compared with baseline period].</p>
Full article ">Figure 4
<p>Projected annual precipitations during the period 2010–2099 [(<b>a</b>,<b>b</b>) are under the models of HadGEM and MIROC in Bayanbulak station; (<b>c</b>,<b>d</b>) are under the models of HadGEM and MIROC in Dashankou station]</p>
Full article ">Figure 5
<p>Seasonal precipitation changes compared with the baseline.</p>
Full article ">Figure 6
<p>Projected annual streamflow [(<b>a</b>) time series of streamflow during the period 2010–2099 and (<b>b</b>) the associated average values and changing trends by Mann–Kendall test].</p>
Full article ">Figure 7
<p>Distributions of annual streamflow in different periods.</p>
Full article ">Figure 8
<p>Monthly streamflow during the period 2010–2099 (the orange color is the range of streamflow, the black line is the average streamflow, and the white dotted line is the trending line).</p>
Full article ">
13 pages, 1828 KiB  
Article
Diurnal Characteristics of Gravity Waves over the Tibetan Plateau in 2015 Summer Using 10-km Downscaled Simulations from WRF-EnKF Regional Reanalysis
by Tingting Qian, Fuqing Zhang, Junhong Wei, Jie He and Yinghui Lu
Atmosphere 2020, 11(6), 631; https://doi.org/10.3390/atmos11060631 - 14 Jun 2020
Cited by 7 | Viewed by 2669
Abstract
Diurnal variations of gravity waves over the Tibetan Plateau (TP) in summer 2015 were investigated based on high-resolution downscaled simulations from WRF-EnKF (Weather Research and Forecasting model and an ensemble Kalman filter) regional reanalysis data with particular emphasis on wave source, wave momentum [...] Read more.
Diurnal variations of gravity waves over the Tibetan Plateau (TP) in summer 2015 were investigated based on high-resolution downscaled simulations from WRF-EnKF (Weather Research and Forecasting model and an ensemble Kalman filter) regional reanalysis data with particular emphasis on wave source, wave momentum fluxes and wave energies. Strong diurnal precipitations, which mainly happen along the south slope of the TP, tend to excite upward-propagating gravity waves. The spatial and temporal distributions of the momentum fluxes of small-scale (10–200 km) and meso-scale (200–500 km) gravity waves agree well with the diurnal precipitation distributions. The power spectra of momentum fluxes also show that the small- and meso-scale atmospheric processes become important during the period of the strongest rainfall. Eastward momentum fluxes and northward momentum fluxes are dominant. Wave energies are described in terms of kinetic energy (KE), potential energy (PE) and vertical fluctuation energy (VE). The diurnal variation and spatial distribution of VE in the lower stratosphere correspond to the diurnal rainfall in the troposphere. Full article
(This article belongs to the Special Issue Gravity Waves in the Atmosphere)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Terrain height (unit: km; shading). The thin black lines are the province border. The red line represents the topography height at 4.5 km. The blue line represents the topography height at 0.3 km. (<b>b</b>) Area-averaged and 3-month-averaged horizontal momentum fluxes per unit volume (unit: Pa) and (<b>c</b>) common logarithm (base 10) of area-averaged and 3-month-averaged horizontal momentum fluxes per unit volume (unit: Pa) over the main TP (solid lines) and the south slope of the TP (dotted lines) from 6-h forecast results (blue lines) and 12-h forecast results (red lines). Legends in (<b>b</b>,<b>c</b>) are abbreviations of the simulation information. For example, f06 (12) refers to the 6(12)-h forecast. TP refers to the main TP region and S refers to the south slope of the TP.</p>
Full article ">Figure 2
<p>(<b>a</b>) Zonal wind speed at 200 hPa (shading: greater than 20 m/s; contour interval is 5 m/s) and geopotential height at 500 hPa (thick black lines). (<b>b</b>) Area and 3-month-averaged zonal wind (blue triangle line) varying with height.</p>
Full article ">Figure 3
<p>6-h accumulated rainfall (shading: shaded region shows rainfall greater than 3 mm/6 h; contour interval is 3 mm/6 h) averaged over the 3-month period at (<b>a</b>) 06 UTC, (<b>b</b>) 12 UTC, (<b>c</b>) 18 UTC, and (<b>d</b>) 00 UTC. Local time is UTC time plus 6 h.</p>
Full article ">Figure 4
<p>Geographic distribution of absolute horizontal momentum flux per unit volume (unit: Pa; shading, contour interval is 0.00109 Pa) at 30 hPa for 12-h forecast results at (<b>a</b>) 06 UTC (12 LT), (<b>b</b>) 12 UTC (18 LT), (<b>c</b>) 18 UTC (00 LT), and (<b>d</b>) 00 UTC (06 LT). The horizontal wavelength is taken from 10 km to 500 km. The red line represents the topography height at 4.5 km. The blue line represents the topography height at 0.3 km.</p>
Full article ">Figure 5
<p>Diurnal variation of (<b>a</b>) area-averaged momentum fluxes per unit volume (unit: Pa), (<b>b</b>) common logarithm (base 10) of area-averaged momentum fluxes per unit volume (unit: Pa) and (<b>c</b>) area-averaged momentum fluxes per unit mass (unit: J/kg) varying with height at 00 UTC (circle), 06 UTC (triangle), 12 UTC (star) and 18 UTC (square) above the main TP (red solid line) and south slope of TP (blue dotted lines). Legends are abbreviations of simulation information. For example, f06 (12) refers to 6(12)-h forecast. t00 (06,12,18) refers to UTC time. TP refers to the main TP region and S refers to the south slope of the TP.</p>
Full article ">Figure 6
<p>Time-averaged zonal momentum flux power spectra at 30 hPa in horizontal (kx, ky) wavenumber space for 12-h forecast results at (<b>a</b>) 06 UTC, (<b>b</b>) 12 UTC, (<b>c</b>) 18 UTC and (<b>d</b>) 00 UTC.</p>
Full article ">Figure 7
<p>Time-averaged meridional momentum flux power spectra at 30 hPa in horizontal (kx, ky) wavenumber space for 12-h forecast results at (<b>a</b>) 06 UTC, (<b>b</b>) 12 UTC, (<b>c</b>) 18 UTC and (<b>d</b>) 00 UTC.</p>
Full article ">Figure 8
<p>Power spectrum of (<b>a</b>) zonal and (<b>b</b>) meridional momentum fluxes per unit volume varying with wavelength at 30 hPa and at different times in a 1-day cycle.</p>
Full article ">Figure 9
<p>Time-averaged (<b>a1</b>–<b>a4</b>) KE (unit: J/kg, shading, c.i. = 10), (<b>b1</b>–<b>b4</b>) PE (unit: J/kg, shading, c.i. = 10), and (<b>c1</b>–<b>c4</b>) VE (unit: J/kg, shading, c.i. = 0.02; due to the large amplitude oscillation at different times, same red color uses for VE value greater than 0.2) at (<b>a1</b>–<b>c1</b>) 06 UTC, (<b>a2</b>–<b>c2</b>) 12 UTC, (<b>a3</b>–<b>c3</b>) 18 UTC, and (<b>a4</b>–<b>c4</b>) 00 UTC at 30 hPa.</p>
Full article ">
Back to TopTop