[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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

Search Results (99)

Search Parameters:
Keywords = annual maxima

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 7833 KiB  
Article
Spatial and Temporal Characterization of Near Space Temperature and Humidity and Their Driving Influences
by Wenhui Luo, Jinji Ma, Miao Li, Haifeng Xu, Cheng Wan and Zhengqiang Li
Remote Sens. 2024, 16(22), 4307; https://doi.org/10.3390/rs16224307 - 19 Nov 2024
Viewed by 463
Abstract
Near space refers to the atmospheric region 20–100 km above Earth’s surface, encompassing the stratosphere, mesosphere, and part of the thermosphere. This region is susceptible to surface and upper atmospheric disturbances, and the atmospheric temperature and humidity profiles can finely characterize its complex [...] Read more.
Near space refers to the atmospheric region 20–100 km above Earth’s surface, encompassing the stratosphere, mesosphere, and part of the thermosphere. This region is susceptible to surface and upper atmospheric disturbances, and the atmospheric temperature and humidity profiles can finely characterize its complex environment. To analyze the relationship between changes in temperature and humidity profiles and natural activities, this study utilizes 18 years of temperature and water vapor data from the TIMED/SABER and AURA/MLS instruments to investigate the variations in temperature and humidity with altitude, time, and spatial distribution. In addition, multiple linear regression analysis is used to examine the impact mechanisms of solar activity, the El Niño–Southern Oscillation (ENSO), and the Quasi-Biennial Oscillation (QBO) on temperature and humidity. The results show that in the mid- and low-latitude regions, temperature and water vapor reach their maxima at an altitude of 50 km, with values of 265 K and 8–9 × 10⁻⁶ ppmv, respectively; the variation characteristics differ across latitudes and altitudes, with a clear annual cycle; the feedback effects of solar activity and the ENSO index on temperature and humidity in the 20–40 km atmospheric layer are significantly different. Among these factors, solar activity is the most significant influence on temperature and water vapor, with response coefficients of −0.2 to −0.16 K/sfu and 0.8 to 4 × 10⁻⁶ ppmv/sfu, respectively. Secondly, in the low-latitude stratospheric region, the temperature response to ENSO is approximately −1.5 K/MEI, while in the high-latitude region, a positive response of 3 K/MEI is observed. The response of water vapor to ENSO varies between −1 × 10⁻⁷ and −4 × 10⁷ ppmv/sfu. In the low-latitude stratospheric region, the temperature and humidity responses to the QBO index exhibit significant differences, ranging from −1.8 to −0.6 K/10 m/s. Additionally, there are substantial differences in responses between the polar regions and the low-latitude equatorial region. Finally, a three-dimensional model coefficient was constructed to illustrate the influence of solar activity, ENSO, and QBO on temperature and humidity in the near space. The findings of this study contribute to a deeper understanding of the temperature and humidity variation characteristics in near space and provide valuable data and model references for predicting three-dimensional parameters of temperature and humidity in this region. Full article
Show Figures

Figure 1

Figure 1
<p>TIMED/SABER temperature data.</p>
Full article ">Figure 2
<p>Aura/MLS temperature data.</p>
Full article ">Figure 3
<p>Time series of solar, ENSO, and QBO activity data from January 2005 to December 2022. (<b>a</b>) Time series of solar activity data from January 2005 to December 2022; (<b>b</b>) Time series of ENSO-MEI data from January 2005 to December 2022; (<b>c</b>) Time series of QBO data from January 2005 to December 2022.</p>
Full article ">Figure 3 Cont.
<p>Time series of solar, ENSO, and QBO activity data from January 2005 to December 2022. (<b>a</b>) Time series of solar activity data from January 2005 to December 2022; (<b>b</b>) Time series of ENSO-MEI data from January 2005 to December 2022; (<b>c</b>) Time series of QBO data from January 2005 to December 2022.</p>
Full article ">Figure 4
<p>Monthly data for temperature and water vapor concentration. (<b>a</b>) Monthly data of temperature data in the mid-latitude region of the Northern Hemisphere from 2005 to 2022; (<b>b</b>) Monthly data of water vapor data in the mid-latitude region of the Northern Hemisphere from 2005 to 2022.</p>
Full article ">Figure 5
<p>Range of temperature and humidity profile extrema. (<b>a</b>) Range of temperature profile extrema; (<b>b</b>) Range of humidity profile extrema.</p>
Full article ">Figure 6
<p>Temperature and water vapor response to solar activity at 180°E. (<b>a</b>) temperature response to solar activity at 180°E; (<b>b</b>) water vapor response to solar activity at 180°E.</p>
Full article ">Figure 7
<p>Temperature and water vapor response to ENSO at 180°E. (<b>a</b>) temperature response to ENSO at 180°E; (<b>b</b>) water vapor response to ENSO at 180°E.</p>
Full article ">Figure 8
<p>Temperature and water vapor response to QBO at 180° E. (<b>a</b>) temperature response to the QBO; (<b>b</b>) water vapor response to the QBO.</p>
Full article ">Figure 9
<p>Box plot of coefficients at different altitudes. (<b>a</b>) Box plot of TMP-solar coefficients at different altitudes; (<b>b</b>) Box plot of TMP-ENSO coefficients at different altitudes; (<b>c</b>) Box plot of TMP-QBO coefficients at different altitudes; (<b>d</b>) Box plot of H<sub>2</sub>O-solar coefficients at different altitudes; (<b>e</b>) Box plot of H<sub>2</sub>O-ENSO coefficients at different altitudes; (<b>f</b>) Box plot of H<sub>2</sub>O-QBO coefficients at different altitudes.</p>
Full article ">Figure 10
<p>3D scatter plot of coefficients. (<b>a</b>) 3D scatter plot of temperature response coefficients to solar activity; (<b>b</b>) 3D scatter plot of temperature response coefficients to ENSO; (<b>c</b>) 3D scatter plot of temperature response coefficients to QBO; (<b>d</b>) 3D scatter plot of water vapor response coefficients to solar activity; (<b>e</b>) 3D scatter plot of water vapor response coefficients to ENSO; (<b>f</b>) 3D scatter plot of water vapor response coefficients to QBO.</p>
Full article ">
23 pages, 3848 KiB  
Article
Evaluation of Tidal Asymmetry and Its Effect on Tidal Energy Resources in the Great Island Region of the Gulf of California
by Anahí Bermúdez-Romero, Vanesa Magar, Manuel López-Mariscal and Jonas D. De Basabe
J. Mar. Sci. Eng. 2024, 12(10), 1740; https://doi.org/10.3390/jmse12101740 - 2 Oct 2024
Viewed by 770
Abstract
Hydrokinetic tidal energy is one of the few marine renewable energy resources with sufficiently mature technology for commercial exploitation. However, several parameters affect its exploitability, such as the minimum speed threshold, ambient turbulence levels, or tidal asymmetry, to name but a few. These [...] Read more.
Hydrokinetic tidal energy is one of the few marine renewable energy resources with sufficiently mature technology for commercial exploitation. However, several parameters affect its exploitability, such as the minimum speed threshold, ambient turbulence levels, or tidal asymmetry, to name but a few. These parameters are particularly important in regions with lower mean speeds than those in first-generation sites, such as the North Sea. The Gulf of California is one of those regions. In this paper, a Delft3D Flexible Mesh Suite (Delft3D FM) model in barotropic configuration is set up over the Gulf of California using a flexible mesh with resolution varying from O (500 m) in the deep regions to O (10 m) in the coastal regions. A simulation is run over the year of 2020, with a tidal forcing of 75 components. The model is validated at four tidal gauge locations and four Acoustic Doppler Current profiler (ADCP) locations. The speed, U, and tidal power density (TPD) indicators used for the validation were the annual means, the annual means for speeds above the 0.5 m s−1 threshold, the annual means of the spring tide maxima, and the annual maxima. The contour maps of the annual means, that is, the annual means for speeds above the 0.5 m s−1 threshold, allow us to identify tidal energy hot spots throughout the Gulf of California, particularly in the Great Island region (GIR). In this region, these hot spots have higher U and TPD values, in agreement with previous studies. The patterns of circulation around Tiburón Island and San Esteban Island on the East, and Ángel de la Guarda Island and San Lorenzo Island on the West, the four islands in the region with the highest tidal energy potential, are also discussed while recognizing that Tiburón Channel, between Tiburón Island and San Esteban Island, has proved to be the best siting location, based on the technical results obtained so far. The hot spots sites are further characterized by computing the tidal asymmetry in these small regions, showing the locations of the sites with smallest asymmetry, which would be the best for tidal energy exploitation. The hot spots around San Esteban Island are particularly important because they have the largest TPD in the GIR, with the model predicting a TPD on the order of 500–1000 W m−2. Here, complementary field measurements obtained with two ADCPs, close to San Esteban Island, one at 15 m depth, SEs (shallow region), and the other at 60 m depth, SEd (deep region), produced TPDs of 1200 W m−2 and 400 W m−2, respectively. The analysis of the vertical profiles and the tidal asymmetry over the vertical shows the importance of developing 3D models in future investigations. Full article
(This article belongs to the Special Issue Advances in Marine Computational Fluid Dynamics)
Show Figures

Figure 1

Figure 1
<p>Study region with bathymetry provided by the product GEBCO2019, and locations of islands and ADCP moorings used for validation in the midriff region. Island Acronyms—AGI: Angel de la Guarda Island. SLI: San Lorenzo Island. SEI: San Esteban Island. TI: Tiburon Island. ADCP Mooring Acronyms—BC: Ballenas Channel. DE: Delfin Sill. SL: San Lorenzo. SE: San Esteban. The insert shows the boundaries of the 24 subdomains generated to run the model. A coordinate system showing the velocity component <span class="html-italic">u</span> perpendicular to the GoC and the velocity component <span class="html-italic">v</span> along the GoC, is included for reference.</p>
Full article ">Figure 2
<p>Central computing subdomain with part of SEI, showing the grid refinement near the shoreline.</p>
Full article ">Figure 3
<p>Annual mean speed (m s<sup>−1</sup>) predicted by the model. The thin black lines show the model subdomains used for computational purposes, as described in the grid generation and analysis <a href="#sec2dot2-jmse-12-01740" class="html-sec">Section 2.2</a>.</p>
Full article ">Figure 4
<p>Time series of observed (blue continuous line) and modeled (red dashed line) flow velocity (<b>left panel</b>) and tidal levels (<b>right panel</b>) for two spring–neap cycles.</p>
Full article ">Figure 5
<p>Maps for the two largest islands in the GIR, AGI, and TI. (<b>Upper panels</b>): annual mean of threshold speed, <math display="inline"><semantics> <msub> <mover> <mi>U</mi> <mo>¯</mo> </mover> <mrow> <mi>t</mi> <mi>h</mi> </mrow> </msub> </semantics></math> [m s<sup>−1</sup>]. (<b>Lower panels</b>): annual mean threshold Tidal Power Density, <math display="inline"><semantics> <msub> <mover> <mrow> <mi>T</mi> <mi>P</mi> <mi>D</mi> </mrow> <mo>¯</mo> </mover> <mrow> <mi>t</mi> <mi>h</mi> </mrow> </msub> </semantics></math> [W m<sup>−2</sup>].</p>
Full article ">Figure 6
<p>Maps for SEI and SLI. (<b>Upper panels</b>): annual mean threshold speed, <math display="inline"><semantics> <msub> <mover> <mi>U</mi> <mo>¯</mo> </mover> <mrow> <mi>t</mi> <mi>h</mi> </mrow> </msub> </semantics></math> [m s<sup>−1</sup>]. (<b>Lower panels</b>): annual mean threshold Tidal Power Density, <math display="inline"><semantics> <msub> <mover> <mrow> <mi>T</mi> <mi>P</mi> <mi>D</mi> </mrow> <mo>¯</mo> </mover> <mrow> <mi>t</mi> <mi>h</mi> </mrow> </msub> </semantics></math> [W m<sup>−2</sup>].</p>
Full article ">Figure 7
<p><math display="inline"><semantics> <mrow> <mi>T</mi> <mi>F</mi> <mi>A</mi> </mrow> </semantics></math> maps around AGI (<b>left panel</b>) and TI (<b>right panel</b>), with contour lines of the <math display="inline"><semantics> <msub> <mi>U</mi> <mrow> <mi>t</mi> <mi>h</mi> </mrow> </msub> </semantics></math>. The white arrows represent a schematic of the direction of the residual circulation around each of the islands.</p>
Full article ">Figure 8
<p>Maps of <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>F</mi> <mi>A</mi> </mrow> </semantics></math> around SEI (<b>left panel</b>) and SLI (<b>right panel</b>) with contour lines of the <math display="inline"><semantics> <msub> <mi>U</mi> <mrow> <mi>t</mi> <mi>h</mi> </mrow> </msub> </semantics></math>. Schematics of general residual circulation around SEI and SLI are also shown.</p>
Full article ">Figure 9
<p>Current speed (<b>left panels</b>) and TPD (<b>right panels</b>) for S1 (flood dominant tide), S2 (ebb dominant tide), and S3 (symmetrical tide). The blue line indicates flood periods and red lines ebb periods.</p>
Full article ">Figure 10
<p>Speed profiles of a neap–spring tidal cycle for SEd (<b>upper panel</b>) and SEs (<b>low panel</b>). The white line shows <span class="html-italic">v</span>, the depth-averaged velocity component along the GoC.</p>
Full article ">Figure 11
<p>TPD profiles for a 15-day period at SEd (<b>upper panel</b>) and SEs (<b>low panel</b>).</p>
Full article ">Figure 12
<p>Vertical profile for SEd (<b>left panel</b>) and SEs (<b>right panel</b>). Vertical averages of speed for flood and ebb periods are shown on the top of the panels.</p>
Full article ">
12 pages, 2039 KiB  
Communication
Interaction of Chicken Heterophils and Eimeria tenella Results in Different Phenotypes of Heterophil Extracellular Traps (HETs)
by Zaida Rentería-Solís, Liliana M. R. Silva, Thomas Grochow, Runhui Zhang, Tran Nguyen-Ho-Bao, Arwid Daugschies, Anja Taubert, Iván Conejeros and Carlos Hermosilla
Poultry 2024, 3(3), 318-329; https://doi.org/10.3390/poultry3030024 - 9 Sep 2024
Viewed by 1063
Abstract
Chicken coccidiosis causes annual losses exceeding GBP 10 billion globally. The most pathogenic species for domestic fowls including Eimeria tenella, E. acervulina, and E. maxima, can lead to gastrointestinal issues ranging from mild to fatal. In this study, stages of [...] Read more.
Chicken coccidiosis causes annual losses exceeding GBP 10 billion globally. The most pathogenic species for domestic fowls including Eimeria tenella, E. acervulina, and E. maxima, can lead to gastrointestinal issues ranging from mild to fatal. In this study, stages of E. tenella and freshly isolated chicken heterophils were co-cultured for 180 min. These interactions were analyzed using live 3D holotomographic and confocal microscopy. We observed that E. tenella stages were entrapped by heterophils and heterophil extracellular traps (HETs). Notably, different HET phenotypes, specifically sprHETs and aggHETs, were induced regardless of the stage. Furthermore, the quantification of extracellular DNA release from co-cultures of heterophils and sporozoites (ratio 1:1) for 180 min demonstrated a significantly higher release (p = 0.04) compared to negative controls. In conclusion, research on the chicken innate immune system, particularly fowl-derived HETs, remains limited. More detailed investigations are needed, such as exploring the time-dependent triggering of HETs, to establish a standard incubation time for this pathogen defense mechanism. This will enhance our understanding of its role in parasite survival or death during HET confrontation. Full article
(This article belongs to the Special Issue Current Research and Key Issues in Poultry Immunology)
Show Figures

Figure 1

Figure 1
<p>Live cell 3D-holotomographic microscopy unveiling interactions of chicken heterophils with <span class="html-italic">E. tenella</span> sporozoites. For 180 min, vital <span class="html-italic">E. tenella</span> sporozoites were co-cultured with freshly isolated heterophils. At the end of the experiment, some cells were still viable (green arrow), and some of them were entrapping sporozoites (zoom, 3D rendering, entrapment: black and white arrows). Yellow arrow: sporozoite membrane, which was not visible at the entrapment point (white and black arrows). Scale bars: 20 and 5 µm.</p>
Full article ">Figure 2
<p><span class="html-italic">Eimeria tenella</span>-triggered different types of HETs. Confocal microscopy analysis of non-confronted heterophils ((<b>A1</b>–<b>A4</b>) control panels) and heterophils confronted with <span class="html-italic">E. tenella</span> sporozoites for 180 min ((<b>B1</b>–<b>B4</b>,<b>C1</b>–<b>C4</b>) <span class="html-italic">E. tenella</span> panels). Structures were stained for DNA (DAPI, blue), elastase (red), and histones (green). Co-localization of the stains proves the nature of avian HETs. (<b>B1</b>–<b>B4</b>) <span class="html-italic">ag</span>gHETs with the involvement of multiple heterophils and the entrapment of <span class="html-italic">E. tenella</span> sporozoites (white arrows). (<b>C1</b>–<b>C4</b>) Several short <span class="html-italic">spr</span>HETs are observed with ensnared <span class="html-italic">E. tenella</span> sporozoites (white arrows) in these thin web-like structures. Percentage of HETosis was 6.14% ± 1.35% (Mean ± standard deviation). Scale bar: 10 µm.</p>
Full article ">Figure 3
<p><span class="html-italic">Eimeria tenella</span>-triggered HETosis is a stage-independent process. <span class="html-italic">E. tenella</span> stages were co-cultured for 180 min (1:1). Sporozoites (light blue arrows), sporocysts (white arrows), and oocysts (yellow arrows) were entrapped within HET-derived filaments formed by DNA backbone (blue) decorated with histones (green) and elastase (red) granules. A sporozoite (light blue arrow) was firmly grabbed by a fine DNA web-like structure. Scale bar: 10 µm.</p>
Full article ">Figure 4
<p>Extracellular DNA quantification. (<b>a</b>) Confrontation of heterophils and <span class="html-italic">E. tenella</span> sporozoites (1:1) conducted to a higher extracellular DNA release than non-stimulated heterophils (<span class="html-italic">p</span> = 0.0399). Individual variation is observed: two animals responded highly to the stimulus, while the third animal reacted mildly to the confrontation of <span class="html-italic">E. tenella</span> sporozoites. (<b>b</b>) Extracellular DNA quantification showed no dose dependency.</p>
Full article ">
20 pages, 12373 KiB  
Article
Monitoring Grassland Variation in a Typical Area of the Qinghai Lake Basin Using 30 m Annual Maximum NDVI Data
by Meng Li, Guangjun Wang, Aohan Sun, Youkun Wang, Fang Li and Sihai Liang
Remote Sens. 2024, 16(7), 1222; https://doi.org/10.3390/rs16071222 - 30 Mar 2024
Cited by 2 | Viewed by 1092
Abstract
The normalized difference vegetation index (NDVI) can depict the status of vegetation growth and coverage in grasslands, whereas coarse spatial resolution, cloud cover, and vegetation phenology limit its applicability in fine-scale research, especially in areas covering various vegetation or in fragmented landscapes. In [...] Read more.
The normalized difference vegetation index (NDVI) can depict the status of vegetation growth and coverage in grasslands, whereas coarse spatial resolution, cloud cover, and vegetation phenology limit its applicability in fine-scale research, especially in areas covering various vegetation or in fragmented landscapes. In this study, a methodology was developed for obtaining the 30 m annual maximum NDVI to overcome these shortcomings. First, the Landsat NDVI was simulated by fusing Landsat and MODIS NDVI by using the enhanced spatial and temporal adaptive reflectance fusion model (ESTARFM), and then a single-peaked symmetric logistic model was employed to fit the Landsat NDVI data and derive the maximum NDVI in a year. The annual maximum NDVI was then used as a season-independent substitute to monitor grassland variation from 2001 to 2022 in a typical area covering the major vegetation types in the Qinghai Lake Basin. The major conclusions are as follows: (1) Our method for reconstructing the NDVI time series yielded higher accuracy than the existing dataset. The root mean square error (RMSE) for 91.8% of the pixels was less than 0.1. (2) The annual maximum NDVI from 2001 to 2022 exhibited spatial distribution characteristics, with higher values in the northern and southern regions and lower values in the central area. In addition, the earlier vegetation growth maximum dates were related to the vegetation type and accompanied by higher NDVI maxima in the study area. (3) The overall interannual variation showed a slight increasing trend from 2001 to 2022, and the degraded area was characterized as patches and was dominated by Alpine kobresia spp., Forb Meadow, whose change resulted from a combination of permafrost degradation, overgrazing, and rodent infestation and should be given more attention in the Qinghai Lake Basin. Full article
(This article belongs to the Special Issue Remote Sensing of Mountain and Plateau Vegetation)
Show Figures

Figure 1

Figure 1
<p>Location and vegetation types of the study area.</p>
Full article ">Figure 2
<p>Comparison of the NDVI images on 8 September 2022. (<b>a</b>) Landsat reference NDVI image, (<b>b</b>) reconstructed NDVI image.</p>
Full article ">Figure 3
<p>Spatial distribution and share of classes of RMSE in 2022.</p>
Full article ">Figure 4
<p>Landsat NDVI fitting curves of typical vegetation types. (<b>a</b>) <span class="html-italic">Alpine kobresia</span> spp., <span class="html-italic">Forb Meadow</span>. (<b>b</b>) <span class="html-italic">Subalpine broadleaf deciduous scrub</span>. (<b>c</b>) <span class="html-italic">Alpine grass</span>, <span class="html-italic">Carex Steppe</span>. (<b>d</b>) <span class="html-italic">Cultivated vegetation</span>. (<b>e</b>) <span class="html-italic">Temperate deciduous scrub</span>. (<b>f</b>) <span class="html-italic">Temperate needlegrass arid steppe</span>.</p>
Full article ">Figure 5
<p>Maxday for all grasslands in 2022. (<b>a</b>) Spatial distribution. (<b>b</b>) Frequency distribution.</p>
Full article ">Figure 6
<p>Spatial distribution of Landsat NDVI maxima from 2001 to 2022 and the DEM.</p>
Full article ">Figure 7
<p>Spatial distribution of the NDVI change from 2001 to 2022. (<b>a</b>) Slope. (<b>b</b>) CV.</p>
Full article ">Figure 8
<p>Degraded <span class="html-italic">Alpine kobresia</span> spp., <span class="html-italic">Forb Meadow</span> in the southern region of the study area.</p>
Full article ">Figure 9
<p>Degraded frozen soil zone in the southern region of the study area.</p>
Full article ">Figure 10
<p>Dense clusters of pika burrows in the degraded meadow areas (99°24′E, 37°04′N).</p>
Full article ">
20 pages, 14650 KiB  
Article
Towards Improved Satellite Data Utilization in China: Insights from an Integrated Evaluation of GSMaP-GNRT6 in Rainfall Patterns
by Zunya Wang and Qingquan Li
Remote Sens. 2024, 16(5), 755; https://doi.org/10.3390/rs16050755 - 21 Feb 2024
Cited by 1 | Viewed by 941
Abstract
To improve the utilization of satellite-based data and promote their development, this analysis comprehensively evaluates the performance of GSMaP Near-real-time Gauge-adjusted Rainfall Product version 6 (GSMaP_GNRT6) data in depicting precipitation over China from 2001 to 2020 by comparing four precipitation indices—accumulated precipitation, number [...] Read more.
To improve the utilization of satellite-based data and promote their development, this analysis comprehensively evaluates the performance of GSMaP Near-real-time Gauge-adjusted Rainfall Product version 6 (GSMaP_GNRT6) data in depicting precipitation over China from 2001 to 2020 by comparing four precipitation indices—accumulated precipitation, number of rainy days and rainstorm days, and precipitation maxima—with daily precipitation data from 2419 stations across China on monthly and annual time scales. The results show that the GSMaP-GNRT6 data effectively capture the overall spatial pattern of the four precipitation indices, but biases in the spatial distribution of the number of rainy days from July to September and the precipitation maxima during the wintertime are evident. A general underestimation of GSMaP-GNRT6 data is observed in the average for China. The annual precipitation amount, the number of rainy days and rainstorm days, and the precipitation maxima based on the GSMaP-GNRT6 data are 17.6%, 35.5%, 31.6% and 11.8% below the station observations, respectively. The GSMaP-GNRT6 data better depict the precipitation in eastern China, with the errors almost halved. And obvious overestimation of the number of rainstorm days and precipitation maxima occurs in western China, reaching up to 60%. Regarding the accumulated precipitation, the number of rainstorm days and the precipitation maxima, the GSMaP-GNRT6 data show an almost consistent interannual variation and increasing trends that are consistent with the station observations. However, the magnitude of the increasing trend based on the GSMaP-GNRT6 data is larger, especially at the beginning of the 21st century. Conversely, a considerable discrepancy in the annual variation and an almost opposite trend can be observed in the number of rainy days between the GSMaP-GNRT6 data and the station observations. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of observatory stations in China used in this analysis (black dots), shadings denoting the elevation of the terrain.</p>
Full article ">Figure 2
<p>Distribution of annual precipitation in China based on station observations ((<b>a</b>), unit: mm) and GSMaP-GNRT6 data ((<b>b</b>), unit: mm), as well as biases ((<b>c</b>), unit: mm), absolute mean errors ((<b>d</b>), unit: mm), relative errors ((<b>e</b>), unit: %) and root mean square errors ((<b>f</b>), unit: mm) of GSMaP-GNRT6 data compared to station observations.</p>
Full article ">Figure 3
<p>Spatial correlation coefficients of monthly climatic precipitation (<b>a</b>), number of rainy days (<b>b</b>), number of rainstorm days (<b>c</b>) and precipitation maxima (<b>d</b>) in China (blue lines, 72.05°–122.55°E/17.55°–39.95°N) and eastern China (orange lines, 105.05°–122.55°E/17.55°–39.95°N) between station observations and GSMaP-GNRT6 data.</p>
Full article ">Figure 4
<p>Climatological monthly precipitation averaged over China ((<b>a</b>), unit: mm) and eastern China ((<b>b</b>), unit: mm) based on station observations and GSMaP-GNRT6 data.</p>
Full article ">Figure 5
<p>Distribution of the annual number of rainy days in China based on station observations ((<b>a</b>), unit: days) and GSMaP-GNRT6 data ((<b>b</b>), unit: days), as well as biases ((<b>c</b>), unit: days), absolute mean errors ((<b>d</b>), unit: days), relative errors ((<b>e</b>), unit: %) and root mean square errors ((<b>f</b>), unit: days) of GSMaP-GNRT6 data compared to station observations.</p>
Full article ">Figure 6
<p>Climatological monthly number of rainy days averaged over China ((<b>a</b>), unit: days) and eastern China ((<b>b</b>), unit: days) based on station observations and GSMaP-GNRT6 data.</p>
Full article ">Figure 7
<p>Distribution of the annual number of rainstorm days in China based on station observations ((<b>a</b>), unit: days) and GSMaP-GNRT6 data ((<b>b</b>), unit: days), as well as biases ((<b>c</b>), unit: days), absolute mean errors ((<b>d</b>), unit: days), relative errors ((<b>e</b>), empty areas indicating no occurrence of rainstorm by station observations, unit: %) and root mean square errors ((<b>f</b>), unit: days) of GSMaP-GNRT6 data relative to station observations.</p>
Full article ">Figure 8
<p>Climatological monthly number of rainstorm days averaged over China ((<b>a</b>), unit: days) and eastern China ((<b>b</b>), unit: days) based on station observations and GSMaP-GNRT6 data.</p>
Full article ">Figure 9
<p>Distribution of the annual precipitation maxima in China based on station observations ((<b>a</b>), unit: mm) and GSMaP-GNRT6 data ((<b>b</b>), unit: mm), as well as biases ((<b>c</b>), unit: mm), absolute mean errors ((<b>d</b>), unit: mm), relative errors ((<b>e</b>), unit: %) and root mean square errors ((<b>f</b>), unit: mm) of GSMaP-GNRT6 data compared to station observations.</p>
Full article ">Figure 10
<p>Climatological monthly precipitation maxima averaged over China ((<b>a</b>), unit: mm) and eastern China ((<b>b</b>), unit: mm) based on station observations and GSMaP-GNRT6 data.</p>
Full article ">Figure 11
<p>Distribution of correlation coefficients of annual precipitation (<b>a</b>), number of rainy days (<b>b</b>), number of rainstorm days (<b>c</b>) and precipitation maxima (<b>d</b>) across China from 2001 to 2020 between station observations and GSMaP-GNRT6 data.</p>
Full article ">Figure 12
<p>Time series of annual precipitation ((<b>a</b>,<b>b</b>), unit: mm), number of rainy days ((<b>c</b>,<b>d</b>), unit: days), number of rainstorm days ((<b>e</b>,<b>f</b>), unit: days) and precipitation maxima ((<b>g</b>,<b>h</b>), unit: mm) across China (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) and eastern China (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>) from 2001 to 2020 based on station observations and GSMaP-GNRT6 data.</p>
Full article ">Figure 13
<p>Correlation coefficients of monthly precipitation ((<b>a</b>), unit: mm), number of rainy days ((<b>b</b>), unit: days), number of rainstorm days ((<b>c</b>), unit: days) and precipitation maxima ((<b>d</b>), unit: mm) averaged over China and eastern China between station observations and GSMaP-GNRT6 data.</p>
Full article ">Figure 14
<p>Linear trend coefficients of annual precipitation ((<b>a</b>,<b>b</b>), unit: mm/decade), number of rainy days ((<b>c</b>,<b>d</b>), unit: days/decade), number of rainstorm days ((<b>e</b>,<b>f</b>), unit: days/decade) and precipitation maxima ((<b>g</b>,<b>h</b>), unit: mm/decade) in China based on GSMaP-GNRT6 data (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) and station observations (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>), with the black dots indicating exceeding the confidence level of 95%.</p>
Full article ">
17 pages, 5126 KiB  
Article
Low-Flow Identification in Flood Frequency Analysis: A Case Study for Eastern Australia
by Laura Rima, Khaled Haddad and Ataur Rahman
Water 2024, 16(4), 535; https://doi.org/10.3390/w16040535 - 8 Feb 2024
Viewed by 1770
Abstract
Design flood estimation is an essential step in many water engineering design tasks such as the planning and design of infrastructure to reduce flood damage. Flood frequency analysis (FFA) is widely used in estimating design floods when the at-site flood data length is [...] Read more.
Design flood estimation is an essential step in many water engineering design tasks such as the planning and design of infrastructure to reduce flood damage. Flood frequency analysis (FFA) is widely used in estimating design floods when the at-site flood data length is adequate. One of the problems in FFA with an annual maxima (AM) modeling approach is deciding how to handle smaller discharge values (outliers) in the selected AM flood series at a given station. The objective of this paper is to explore how the practice of censoring (which involves adjusting for smaller discharge values in FFA) affects flood quantile estimates in FFA. In this regard, two commonly used probability distributions, log-Pearson type 3 (LP3) and generalized extreme value distribution (GEV), are used. The multiple Grubbs and Beck (MGB) test is used to identify low-flow outliers in the selected AM flood series at 582 Australian stream gauging stations. It is found that censoring is required for 71% of the selected stations in using the MGB test with the LP3 distribution. The differences in flood quantile estimates between LP3 (with MGB test and censoring) and GEV distribution (without censoring) increase as the return period reduces. A modest correlation is found (for South Australian catchments) between censoring and the selected catchment characteristics (correlation coefficient: 0.43), with statistically significant associations for the mean annual rainfall and catchment shape factor. The findings of this study will be useful to practicing hydrologists in Australia and other countries to estimate design floods using AM flood data by FFA. Moreover, it may assist in updating Australian Rainfall and Runoff (national guide). Full article
(This article belongs to the Section Hydrology)
Show Figures

Figure 1

Figure 1
<p>Geographical distribution of the selected 582 stream gauging stations in eastern Australia.</p>
Full article ">Figure 2
<p>Distribution of record length of the AM flood data and catchment areas of the selected 582 stations.</p>
Full article ">Figure 3
<p>Number of stations required censoring in this study from 582 stations.</p>
Full article ">Figure 4
<p>Fitting of GEV distribution to station 226222 with L-moments.</p>
Full article ">Figure 5
<p>Fitting of LP3 distribution to station 226222 without censoring.</p>
Full article ">Figure 6
<p>Fitting of LP3 distribution to station 226222 after censoring 13 PILFs.</p>
Full article ">Figure 7
<p>Location of stations that required censoring with LP3 distribution (415 stations).</p>
Full article ">Figure 8
<p>Percentage of AM flood data points that needed censoring.</p>
Full article ">Figure 9
<p>Location of annual maxima (AM) flood data stations that do not need censoring (167 stations).</p>
Full article ">Figure 10
<p>Distribution of new record length of the AM flood data after censoring.</p>
Full article ">Figure 11
<p>Variations in flood quantiles Q<sub>100</sub> (1% AEP) between GEV and LP3 distributions with censoring (total: 534 stations; QLD: 186 stations; VIC: 169 stations; NSW: 152 stations; and SA: 27 stations).</p>
Full article ">Figure 12
<p>Percentage variation between flood quantiles (Q10, Q20, Q50, and Q100) estimated by two methods: LP3 with censoring and GEV with L-moments (without censoring) for 534 stations.</p>
Full article ">Figure 13
<p>Absolute variation (%) in flood quantile estimates between GEV and LP3 distributions (with and without censoring) for the stations that required censoring.</p>
Full article ">
20 pages, 8835 KiB  
Article
Identifying Seismic Anomalies via Wavelet Maxima Analysis of Satellite Microwave Brightness Temperature Observations
by Haochen Wu, Pan Xiong, Jianghe Chen, Xuemin Zhang and Xing Yang
Remote Sens. 2024, 16(2), 303; https://doi.org/10.3390/rs16020303 - 11 Jan 2024
Viewed by 1074
Abstract
This study develops a wavelet maxima-based methodology to extract anomalous signals from microwave brightness temperature (MBT) observations for seismogenic activity. MBT, acquired via satellite microwave radiometry, enables subsurface characterization penetrating clouds. Five surface categories of the epicenter area were defined contingent on position [...] Read more.
This study develops a wavelet maxima-based methodology to extract anomalous signals from microwave brightness temperature (MBT) observations for seismogenic activity. MBT, acquired via satellite microwave radiometry, enables subsurface characterization penetrating clouds. Five surface categories of the epicenter area were defined contingent on position (oceanic/terrestrial) and ambient traits (soil hydration, vegetal covering). Continuous wavelet transform was applied to preprocess annualized MBT readings preceding and succeeding prototypical events of each grouping, utilizing optimized wavelet functions and orders tailored to individualized contexts. Wavelet maxima graphs visually portraying signal intensity variations facilitated the identification of aberrant phenomena, including pre-seismic accrual, co-seismic perturbation, and postseismic remission signatures. The casework found 10 GHz horizontal-polarized MBT optimally detected signals for aquatic and predominantly humid/vegetative settings, whereas 36 GHz horizontal-polarized performed best for arid, vegetated landmasses. Quantitative machine learning methods are warranted to statistically define selection standards and augment empirical forecasting leveraging lithospheric stress state inferences from sensitive MBT parametrization. Full article
(This article belongs to the Section Satellite Missions for Earth and Planetary Exploration)
Show Figures

Figure 1

Figure 1
<p>Flow chart of wavelet analysis method.</p>
Full article ">Figure 2
<p>Stratification of the study area into grid cells of the Nepal earthquake (April 2015).</p>
Full article ">Figure 3
<p>Illustrative depiction of the wavelet maxima.</p>
Full article ">Figure 4
<p>Wavelet maxima sequence diagram.</p>
Full article ">Figure 5
<p>Wavelet maxima from MBT data near the epicenter of the Sea of Okhotsk earthquake (24 May 2013).</p>
Full article ">Figure 6
<p>Seismic wavelet maxima sequence in the Ms 8.3 earthquake in the Sea of Okhotsk.</p>
Full article ">Figure 7
<p>Wavelet maxima from MBT data near the epicenter of the Peru earthquake (26 May 2019).</p>
Full article ">Figure 8
<p>Seismic wavelet maxima sequence in the Peru earthquake.</p>
Full article ">Figure 9
<p>Wavelet maxima from MBT data near the epicenter of the Turkey earthquake (6 February 2023).</p>
Full article ">Figure 10
<p>Seismic wavelet maxima sequence in the Turkey earthquake.</p>
Full article ">Figure 11
<p>Wavelet maxima from MBT data near the epicenter of the New Zealand earthquake (13 November 2016).</p>
Full article ">Figure 12
<p>Seismic wavelet maxima sequence in the New Zealand earthquake.</p>
Full article ">
13 pages, 5805 KiB  
Article
Seasonal Variability of Dissolved Methane in the Shallow Coastal Zone: The Case Study of the Golubaya Bay, Northeastern Part of the Black Sea
by Elena S. Izhitskaya, Alexander V. Egorov and Peter O. Zavialov
J. Mar. Sci. Eng. 2023, 11(12), 2284; https://doi.org/10.3390/jmse11122284 - 1 Dec 2023
Viewed by 1031
Abstract
The variability of the dissolved methane content in coastal zones is an important component of the biogeochemical cycle in the marine ecosystem. The objective of this study is to investigate the seasonal variability of dissolved methane distribution in the aerobic shallow coastal zone [...] Read more.
The variability of the dissolved methane content in coastal zones is an important component of the biogeochemical cycle in the marine ecosystem. The objective of this study is to investigate the seasonal variability of dissolved methane distribution in the aerobic shallow coastal zone through the example of the small bay in the northeastern Black Sea. This study is based on the direct observations carried out during a long-term monitoring program conducted in the bay from 1999 to 2016. The seasonal and inter-annual variability of the dissolved methane pattern is considered under the climatic conditions as well as under the influence of extreme flood. The seasonal range of the dissolved methane content variability in the shallow part of the northeastern Black Sea is 1–2 orders of magnitude higher compared with the areas remote from the coast. The dissolved methane content in Golubaya Bay in summer is an order of magnitude higher than the winter values. In particular, local methane maxima located near the river and stream mouths and in the central bottom part of the bay have a well-shown seasonal cycle. The extreme flood conditions observed in July 2012 resulted in high methane concentrations 2 months after the flood event, when the surface concentrations of the dissolved CH4 exceeded the equilibrium with the atmospheric values by a factor of 400. The obtained results provide a unique opportunity to estimate the scale of the biogeochemical processes in marine coastal environments under the influence of climatic and extreme conditions. Full article
(This article belongs to the Special Issue Phytoplankton Dynamics and Biogeochemistry of Marine Ecosystems)
Show Figures

Figure 1

Figure 1
<p>Study region: location of Golubaya Bay in the northeastern part of the Black Sea shore and an approximate bathymetric map of the bay.</p>
Full article ">Figure 2
<p>Dissolved methane content in the waters of Golubaya Bay during the period of observations from 1999 to 2016.</p>
Full article ">Figure 3
<p>Dissolved methane distribution in the water layers of Golubaya Bay in September 1999 ((<b>a</b>)—surface), June 2011 ((<b>b</b>)—surface, (<b>d</b>)—bottom), February 2012 ((<b>c</b>)—surface, (<b>e</b>)—bottom), June 2012 ((<b>f</b>)—surface, (<b>i</b>)—bottom), September 2012 ((<b>g</b>)—surface, (<b>j</b>)—bottom) and January 2013 ((<b>h</b>)—surface). Water sampling stations are indicated by dark blue rhombuses. The mouths of the watercourses are marked with blue triangles.</p>
Full article ">Figure 4
<p>Dissolved methane distribution in the water layers of Golubaya Bay in June 2013 ((<b>a</b>)—surface), July 2013 ((<b>b</b>)—surface, (<b>d</b>)—bottom), June 2014 ((<b>c</b>)—surface, (<b>e</b>)—bottom), June 2015 ((<b>f</b>)—surface, (<b>i</b>)—bottom), October 2015 ((<b>g</b>)—surface, (<b>j</b>)—bottom), January 2016 ((<b>h</b>)—surface, (<b>k</b>)—bottom) and October 2016 ((<b>l</b>)—surface). Water sampling stations are indicated by dark blue rhombuses. The mouths of the watercourses are marked with blue triangles.</p>
Full article ">Figure 5
<p>Dissolved methane concentrations in the waters of Ashamba River (2001–2016) and at the section from the Ashamba River mouth to the open sea (June 2014).</p>
Full article ">Figure 6
<p>Mean concentrations of dissolved methane for each of the surveys (blue points) depending on the time (1999–2016) in Golubaya Bay. The error bars of the datasets are shown by dark-red lines.</p>
Full article ">
19 pages, 6732 KiB  
Article
Atmospheric CO2 Isotopic Variations, with Estimation of Ocean and Plant Source Contributions
by Tom Quirk and Michael Asten
Atmosphere 2023, 14(11), 1623; https://doi.org/10.3390/atmos14111623 - 29 Oct 2023
Viewed by 1538
Abstract
This analysis uses both atmospheric CO2 concentrations and the accompanying δ13C isotopic measurements of CO2 over 40 years from 1978 to 2015 observed at ten different latitudes from 90° S to 82 °N. Atmospheric CO2 is separated into [...] Read more.
This analysis uses both atmospheric CO2 concentrations and the accompanying δ13C isotopic measurements of CO2 over 40 years from 1978 to 2015 observed at ten different latitudes from 90° S to 82 °N. Atmospheric CO2 is separated into two components of CO2 attributable to deep ocean and to plant (including fossil fuel) sources. The isotopic values assigned to the two components are δ13C = 0‰ and −26‰, respectively. The latitude variations in residual source component CO2 show the ocean source component peaking at the equator. This contrasts with the residual plant source component that peaks in the Arctic Circle region. Seasonal comparisons show no change in the ocean component peaking at the equator and no significant changes in its variation with latitude, while the plant component shows seasonal changes of the order of 15 ppm at high latitudes. The ocean component shows clear anomalous behavior in the three years following the 1989 Pacific Ocean regime shift (a shift independently identified from the changed biological time series). By contrast, the residual plant component shows a correlation in the timing of maxima in its annual variations with the timing of El Nino events over the time span of 1985–2015. It also shows a discontinuity in annual variation coinciding with the 1995 AMO phase change. We conclude that the ocean and plant components of atmospheric CO2 relate to independent sources of atmospheric CO2 and have approximately equal magnitudes. The observations are consistent with a hypothesis that variations in the ocean components have an origin from upwelling water from deep ocean currents, and variations in plant components are dominated by a combination of fossil fuel CO2, phytoplankton productivity, and forest and peat fires, which primarily occur in the northern hemisphere. Full article
Show Figures

Figure 1

Figure 1
<p>Monthly atmospheric CO<sub>2</sub> at the South Pole 90° S, Christmas Island 2° N and Point Barrow 71° N, for 1978 to 2015. (<b>a</b>) CO<sub>2</sub> concentration in ppm. (<b>b</b>) CO<sub>2</sub> isotopic composition δ<sup>13</sup>C ‰.</p>
Full article ">Figure 2
<p>Monthly ocean component CO<sub>2</sub> concentrations at the South Pole, Christmas Island and Point Barrow for 1978 to 2015.</p>
Full article ">Figure 3
<p>Monthly plant component CO<sub>2</sub> concentrations at the South Pole, Christmas Island and Point Barrow for 1978 to 2015. An increase is evident around 1989 (the “bubble”).</p>
Full article ">Figure 4
<p>Residual annual variations in total CO<sub>2</sub> less the value at the South Pole, averaged for 1986 to 2015, for each of the ten stations listed in <a href="#atmosphere-14-01623-t001" class="html-table">Table 1</a>. Black vertical bars show standard deviations associated with blue data points.</p>
Full article ">Figure 5
<p>Residual annual variations in CO<sub>2</sub> source components, less the value at the South Pole, averaged for 1986 to 2015. (<b>a</b>) The ocean component and (<b>b</b>) the plant component.</p>
Full article ">Figure 6
<p>Residual seasonal variations in CO<sub>2</sub> source components, less the value at the South Pole, averaged for 1986 to 2015. (<b>a</b>) Ocean source for end boreal winter, average of February and March (<b>b</b>) plant source for end boreal winter, (<b>c</b>) ocean source for end boreal summer, average of August and September and (<b>d</b>) plant source for end boreal summer.</p>
Full article ">Figure 7
<p>Seasonal differences for end boreal winter less end boreal summer. Black vertical bars show standard deviations associated with blue data points. (<b>a</b>) The residual ocean component and (<b>b</b>) the residual plant component.</p>
Full article ">Figure 8
<p>Sensitivity of computed residual annual variations in CO<sub>2</sub> source components, less the value at the South Pole, averaged for 1986 to 2015. (<b>a</b>) Residual ocean component using δ<sup>13</sup>C<sub>ocean</sub> = 0‰ (as for <a href="#atmosphere-14-01623-f005" class="html-fig">Figure 5</a>a), plus computed values where δ<sup>13</sup>C<sub>ocean</sub> is perturbed by ±2. (<b>b</b>) Residual ocean component using δ<sup>13</sup>C<sub>plant</sub> = −26‰, and also perturbed by ±2. (<b>c</b>) Residual plant component using δ<sup>13</sup>C<sub>ocean</sub> = 0 (as for <a href="#atmosphere-14-01623-f005" class="html-fig">Figure 5</a>b), plus computed values where the δ<sup>13</sup>C<sub>ocean</sub> value of 0‰ is perturbed by ±2. (<b>d</b>) Residual plant component using the value δ<sup>13</sup>C<sub>plant</sub> = −26‰, and also perturbed by ±2.</p>
Full article ">Figure 9
<p>Annually averaged CO<sub>2</sub> measurements for the South Pole, Christmas Island, Mauna Loa and Point Barrow.</p>
Full article ">Figure 10
<p>Ocean component values obtained from annually averaged CO<sub>2</sub> observations for the South Pole, Christmas Island, Mauna Loa and Point Barrow. Note the straight-line fit (black line) for the South Pole from 1979 to 2000, which is projected to 2015.</p>
Full article ">Figure 11
<p>Annual increases in ocean component CO<sub>2</sub>, averaged for all stations. The increase is 0.73 ppm per year for 1986 to 2000 and 1.07 ppm per year for 2002 to 2015.</p>
Full article ">Figure 12
<p>Global lower troposphere anomaly from UAH satellite measurements [<a href="#B17-atmosphere-14-01623" class="html-bibr">17</a>]. Solid straight lines are least-squares fits for 1979 to 2000 and 2001 to 2015.</p>
Full article ">Figure 13
<p>Annual plant component CO<sub>2</sub> for the South Pole, Christmas Island, Mauna Loa and Point Barrow. The 1989 regime shift (“bubble”) in plant component atmospheric CO<sub>2</sub> is most apparent for the northernmost latitude, and for the years 1988 to 1992. The black straight line is a least-squares fit to the plant component CO<sub>2</sub> from 1978 to 2018 at the South Pole.</p>
Full article ">Figure 14
<p>Residual annual values from the straight-line fit to the South Pole measurements shown in <a href="#atmosphere-14-01623-f013" class="html-fig">Figure 13</a>. The vertical dashed line at 1995 denotes the time of the phase change in the AMO as given by [<a href="#B18-atmosphere-14-01623" class="html-bibr">18</a>]. Note also the presence of the “bubble” in the years of 1988 to 1992.</p>
Full article ">Figure 15
<p>Annual changes in atmospheric CO<sub>2</sub> concentrations with ocean and plant components of CO<sub>2</sub> in the atmosphere, averaged for 9 stations listed in <a href="#atmosphere-14-01623-t007" class="html-table">Table 7</a>. (<b>a</b>) Atmospheric CO<sub>2</sub> concentrations, (<b>b</b>) plant component CO<sub>2</sub>, (<b>c</b>) ocean component CO<sub>2</sub> and (<b>d</b>) Nino 3.4 Index. Red lines correspond with El Nino conditions (index &gt; 0.5).</p>
Full article ">
14 pages, 17573 KiB  
Article
The Variability of Maximum Daily Precipitation and the Underlying Circulation Conditions in Kraków, Southern Poland
by Robert Twardosz, Marta Cebulska and Izabela Guzik
Water 2023, 15(21), 3772; https://doi.org/10.3390/w15213772 - 28 Oct 2023
Cited by 2 | Viewed by 1406
Abstract
This article studies the intra-annual and long-term variability in the maximum daily precipitation totals and their association with atmospheric circulation in Kraków. It investigates daily precipitation maxima by year and by month. The research is based on daily precipitation totals in the years [...] Read more.
This article studies the intra-annual and long-term variability in the maximum daily precipitation totals and their association with atmospheric circulation in Kraków. It investigates daily precipitation maxima by year and by month. The research is based on daily precipitation totals in the years 1863–2021 and draws on the calendar of atmospheric circulation types by Niedźwiedź. It examines the frequency of precipitation maxima in individual months and their variation from one year to another. No statistically significant trend of change in precipitation over the study period has been found. All annual maximum daily precipitation totals in Kraków fall into the category of heavy precipitation (>10 mm), and almost 99% qualify as very heavy (>20 mm). In the summer months, these are about 3–4 times higher than in winter. The share of the daily precipitation maximum in the monthly total exceeds 30% in all months. The maximum daily precipitation occurring on 5 August 2021 was the highest in the period that extends from the start of instrumental measurements. The study period saw 12 cases of maximum precipitation that belong to ‘flood-inducing’ categories (over 70 mm/day). Such cases of the very heaviest precipitation occurred in cyclonic situations: Cc, Bc, Nc, NEc, Ec and SEc. Most spring and summer maxima were seen on days with a cyclonic circulation. The instances of high daily precipitation in the Kraków area led to the flooding of residential and historic buildings, as well as of municipal infrastructure. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of Kraków and of the city centre meteorological station on the map of Europe.</p>
Full article ">Figure 2
<p>Intra-day evolution of rainfall on 9 September 1963 and 5 August 2021.</p>
Full article ">Figure 3
<p>Share (%) of annual maximum daily precipitation (AMDP) in the annual total (1863–2021).</p>
Full article ">Figure 4
<p>Evolution of annual maximum daily precipitation (AMDP) and values smoothed by 5-year consecutive averages, and the trend line (1863–2021).</p>
Full article ">Figure 5
<p>Statistical characteristics of maximum daily precipitation (MMDP) by month (1863–2021).</p>
Full article ">Figure 6
<p>Absolute topography at 500 hp (dmgp, black lines), ground-level pressure (hPa, white lines) and relative topography 500–1000 hp (dmgp, colours) (<b>a</b>), distribution of the geopotential field and air temperatures at the 850 hPa (<b>b</b>) pressure levels, CAPE [J/kg] (<b>c</b>) and distribution of relative air humidity [%] at 700 hPa (<b>d</b>) on 9 September 1963 (12 UTC) [<a href="#B31-water-15-03772" class="html-bibr">31</a>]. Red square frame: Area of Poland.</p>
Full article ">Figure 7
<p>Surface synoptic map (IMGW-PIB) (<b>a</b>), distribution of the geopotential field and air temperatures at the isobaric levels of 500 hPa (<b>b</b>) and 850 hPa (<b>c</b>), distribution of relative air humidity [%] at 700 hPa (<b>d</b>) [<a href="#B31-water-15-03772" class="html-bibr">31</a>] on 5 August 2021 (12 UTC) and 6 August 2021 (00 UTC). Red square frame: Area of Poland.</p>
Full article ">
452 KiB  
Proceeding Paper
Development of Non-Stationary Rainfall Intensity–Duration–Frequency Curves for Calabar City, Nigeria
by Inyeneobong Cletus Odiong and Jonah C. Agunwamba
Eng. Proc. 2023, 56(1), 89; https://doi.org/10.3390/ASEC2023-15393 - 27 Oct 2023
Viewed by 712
Abstract
Rainfall intensity–duration–frequency (IDF) relationships are crucial in the design and management of hydraulic structures. At the core of the assumption for IDF development is that the statistics of past rainfall events will represent future rainfall events. It has been proven that climate change [...] Read more.
Rainfall intensity–duration–frequency (IDF) relationships are crucial in the design and management of hydraulic structures. At the core of the assumption for IDF development is that the statistics of past rainfall events will represent future rainfall events. It has been proven that climate change is a major trigger for non-stationarity; therefore, the assumption is untenable. This work is aimed at considering the impact of climate change in the development of IDF curves for a city. To account for this non-stationarity, an RCM was combined with measured data through a climate factor (CF) to develop a rainfall IDF for the coastal city of Calabar. The baseline and future climatic periods of the RCM were 1971–2010 and 2021–2060, respectively. The annual maxima series (AMS) were disaggregated and fitted to the Gumbel distribution. Results revealed that the magnitude of trend for the measured AMS and measured annual rainfall are −0.351 and +3.628, respectively. A CF value of 0.86 was obtained, and a generalized non-stationary rainfall IDF model was derived. When compared to models from similar studies, this model has conserved values with r2 = 1 and an error margin of ±6% for all return periods. This will introduce economy in the design of hydraulic structures. Excess runoffs in Calabar were, therefore, related to frequent short-duration rainfall with low intensities. Full article
(This article belongs to the Proceedings of The 4th International Electronic Conference on Applied Sciences)
Show Figures

Figure 1

Figure 1
<p>Non-stationary rainfall IDF curves for Calabar on logarithmic scales.</p>
Full article ">
27 pages, 7879 KiB  
Article
Changes in Magnitude and Shifts in Timing of Australian Flood Peaks
by Mohammed Abdul Bari, Gnanathikkam Emmanuel Amirthanathan, Fitsum Markos Woldemeskel and Paul Martinus Feikema
Water 2023, 15(20), 3665; https://doi.org/10.3390/w15203665 - 19 Oct 2023
Cited by 2 | Viewed by 2111
Abstract
We analysed changes in magnitude and timing of the largest annual observed daily flow (Amax), in each water year, for 596 stations in high-value water resource catchments and flood risk locations across Australia. These stations are either included in the Bureau of Meteorology’s [...] Read more.
We analysed changes in magnitude and timing of the largest annual observed daily flow (Amax), in each water year, for 596 stations in high-value water resource catchments and flood risk locations across Australia. These stations are either included in the Bureau of Meteorology’s Hydrologic Reference Stations or used in its operational flood forecasting services. Monotonic trend (which is either consistently increasing or decreasing) analyses of the magnitude and timing of flood peaks (estimated using Amax) were performed using the Theil–Sen and Mann–Kendall approaches and circular statistics to identify the strength of seasonality and timing. We analysed regional significance across different drainage divisions using the Walker test. Monotonic decreasing trends in Amax flood magnitude were found in the Murray–Darling River Basin and in other drainage divisions in Victoria, southwest and midwest of Western Australia and South Australia. No significant obvious pattern in Amax magnitude was detected in northern Queensland, coastal NSW, central Australia and Tasmania. Monotonic increasing trends were only found in the Tanami–Timor Sea Coast drainage division in northern Australia. Monotonic trends in Amax magnitude were regionally significant at the drainage division scale. We found two distinct patterns in flood seasonality and timing. In the northern and southern parts of Australia, flood peaks generally occur from February to March and August to October, respectively. The strength of this seasonality varies across the country. Weaker seasonality was detected for locations in the Murray–Darling River Basin, and stronger seasonality was evident in northern Australia, the southwest of Western Australia, South Australia, Victoria and Tasmania. The trends of seasonality and timing reveal that in general, flood peaks have occurred later in the water year in recent years. In northern Australia, flood peaks have generally occurred earlier, at a rate of 12 days/decade. In Victoria, New South Wales and Tasmania, the trends in timing are generally mixed. However, in the southwest of Western Australia, the largest change in timing was evident, with Amax peaks commencing later at a rate of 15 days/decade. Decadal variability in flood timing was found at the drainage division scale as well. Most stations show a decreasing trend in Amax magnitude, but how that trend is associated with the change in timing is not clear. Full article
Show Figures

Figure 1

Figure 1
<p>Map showing climate zones [<a href="#B52-water-15-03665" class="html-bibr">52</a>], drainage divisions [<a href="#B53-water-15-03665" class="html-bibr">53</a>] and location of streamflow measurement stations.</p>
Full article ">Figure 2
<p>Record length for the 596 locations and its distribution across the 11 drainage divisions [<a href="#B53-water-15-03665" class="html-bibr">53</a>] (SWP has no locations; NWP not shown).</p>
Full article ">Figure 3
<p>Flow diagram of trend analyses: Amax and shifts in timing.</p>
Full article ">Figure 4
<p>Box plot of Theil–Sen slope (<b>a</b>) in mm/day/decade for Amax and (<b>b</b>) days/decade for timing for stations (<span class="html-italic">n</span> = 211 for Amax and <span class="html-italic">n</span> = 65 for timing) showing significant statistical trends (<span class="html-italic">p</span> &lt; 0.1).</p>
Full article ">Figure 5
<p>Examples of typical trends in the magnitude of Amax (<b>a</b>) decreasing in TAS and (<b>b</b>) increasing in TTS drainage divisions.</p>
Full article ">Figure 6
<p>Maps showing trends in Amax streamflow using (<b>a</b>) MK1, (<b>b</b>) (MK3) and (<b>c</b>) MK3bs tests at <span class="html-italic">p</span> &lt; 0.10. Upward (green) and downward (red) pointing triangles indicate significant increasing and decreasing trends, respectively. Blue dots indicate stations with no trends. Divisions with positive and negative trends with regional significance at <span class="html-italic">p</span> &lt; 0.10 are coloured blue and yellow, respectively.</p>
Full article ">Figure 7
<p>Streamflow timing: (<b>a</b>) Start of site water year, (<b>b</b>) Test of circular uniformity using the Rayleigh test with null hypothesis; that the distribution is uniform shows that the 592 sites are distributed non-uniformly.</p>
Full article ">Figure 8
<p>Timing of Amax peaks (percentage of stations). Amax peaks are concentrated in February–March and August–September.</p>
Full article ">Figure 9
<p>Observed average timing and seasonality of Amax flood peaks across Australia. Each arrow represents one monitoring station (<span class="html-italic">n</span> = 592). Arow colour, direction and length indicate the average timing and the concentration of Amax (R) within the water year, respectively (0: evenly distributed throughout the year; 1: all occur on the same date).</p>
Full article ">Figure 10
<p>Linear trend in (<b>a</b>) magnitude and (<b>b</b>) timing using the Theil–Sen estimator for flood (1950–2022). Each dot represents the median trend of the station (<span class="html-italic">n</span> = 596). The trend is expressed in (<b>a</b>) mm per decade and (<b>b</b>) days per decade, with red colour representing (<b>a</b>) a decreasing trend in magnitude and (<b>b</b>) a shift to earlier in the water year and blue colour representing (<b>a</b>) an increasing trend in magnitude and (<b>b</b>) a shift to later in the water year. The sites in (<b>a</b>) (<span class="html-italic">n</span> = 212) and (<b>b</b>) (<span class="html-italic">n</span> = 65) with dark outer circles have significant trends (<span class="html-italic">p</span> &lt; 10%).</p>
Full article ">Figure 11
<p>Long-term temporal changes in timing of floods in four selected drainage divisions: (<b>a</b>) NEC, (<b>b</b>) MDB, (<b>c</b>) SWC and (<b>d</b>) TTS. Solid lines show median timing over the entire drainage division; shaded bands indicate variability of timing within the year (±standard deviations). All data were subjected to a 10-year moving average filter. Other divisions are shown in <a href="#water-15-03665-f0A1" class="html-fig">Figure A1</a>, <a href="#app1-water-15-03665" class="html-app">Appendix A</a>.</p>
Full article ">Figure 12
<p>Relationship of trends in the magnitude and changes in timing of Amax. Negative and positive values in timing indicate earlier and later changes, respectively (<span class="html-italic">n</span> = 30).</p>
Full article ">Figure A1
<p>Long – term temporal changes in timing of floods in all drainage divisions: (<b>a</b>) NEC, (<b>b</b>) SEN, (<b>c</b>) SEV, (<b>d</b>) TAS, (<b>e</b>) MDB, (<b>f</b>) SAG, (<b>g</b>) SWC, (<b>h</b>) PG, (<b>i</b>) TTS, (<b>j</b>) CC and (<b>k</b>) LEB. Solid lines show median timing over the entire drainage division; shaded bands indicate variability of timing within the year (±standard deviations). All data were subjected to a 10 – year moving average filter.</p>
Full article ">
23 pages, 13587 KiB  
Article
Changes of Hydrological Extremes in the Center of Eastern Europe and Their Plausible Causes
by Irina S. Danilovich, Vladimir F. Loginov and Pavel Y. Groisman
Water 2023, 15(16), 2992; https://doi.org/10.3390/w15162992 - 19 Aug 2023
Cited by 5 | Viewed by 2106
Abstract
Regional studies of precipitation changes over Europe show that its eastern part is characterized by small changes in annual precipitation and insignificant aridity trends compared to central and southern Europe. However, a frequency analysis over the past 30 years showed statistically significant increasing [...] Read more.
Regional studies of precipitation changes over Europe show that its eastern part is characterized by small changes in annual precipitation and insignificant aridity trends compared to central and southern Europe. However, a frequency analysis over the past 30 years showed statistically significant increasing dryness trends in eastern Europe and an increase in the occurrence of extremely high rainfall as well as prolonged no-rain intervals during the warm season. The largest increase in aridity was observed in the western and central parts of Belarus. During 1990–2020, the frequency of dry periods doubled in all river basins along the Black, Caspian, and Baltic Sea water divide areas of eastern Europe. From 1970 to 1990, there were high streamflow rates during the winter low-flow season. Consequently, over the past 50 years, in spring, we observed here a continued decrease in maximal discharges across all river basins. In summer, we detected a statistically significant increase in the number of days with anticyclonic weather over eastern Europe, a decrease in rainfall duration by 15–20%, an increase in daily precipitation maxima by 20–30%, and an increase in the number of days with a low relative humidity by 1–4 days per decade. Full article
(This article belongs to the Section Hydrology)
Show Figures

Figure 1

Figure 1
<p>The territory of Belarus in Europe (<b>a</b>) and location of its meteorological and hydrological stations (<b>b</b>) In the Belarussian map, green station numbers correspond to the meteorological stations and blue numbers to hydrological stations, listed in <a href="#app1-water-15-02992" class="html-app">Appendix A</a>.</p>
Full article ">Figure 2
<p>Spatial distribution of 30 year mean values of annual air temperature (<b>a</b>), precipitation totals (<b>b</b>), snow water equivalent (<b>c</b>), and difference between precipitation and river runoff (<b>d</b>) for the 1991–2020 period.</p>
Full article ">Figure 3
<p>Seasonal changes in nationally averaged surface air temperature (<b>a</b>) and precipitation (<b>b</b>) over Belarus during the past 75 years (1945–2020). Mean rates of change (linear trend estimates in °C y<sup>−1</sup> and mm y<sup>−1</sup>, respectively) are shown by red lines. Estimates of statistical significance of linear trends using the <span class="html-italic">t</span>-test are also shown.</p>
Full article ">Figure 4
<p>Spatial distribution of the winter runoff fraction (%), 1945–2020.</p>
Full article ">Figure 5
<p>Spatial distribution of winter flood peaks (L s<sup>−1</sup> km<sup>−2</sup>), 1945–2020.</p>
Full article ">Figure 6
<p>Spatial distribution of spring flood peaks (L s<sup>−1</sup> km<sup>−2</sup>), 1945–2020.</p>
Full article ">Figure 7
<p>Spatial distribution of minimal discharge trends (L s<sup>−1</sup> km<sup>−2</sup> per 75 years), 1945–2020.</p>
Full article ">Figure 8
<p>The probability of discharge exceedance for two periods (1945–1990 and 1990–2020) at three rivers, Western Dvina near Vitebsk, Neman near Grodno, and Dnepr near Rechitsa.</p>
Full article ">Figure 9
<p>Number of days with anticyclone weather over eastern Europe (within a radius of 1500 km from the city of Minsk for at least one point of an anticyclone track).</p>
Full article ">
17 pages, 1812 KiB  
Article
Intensity–Duration–Frequency Curves for Dependent Datasets
by Wafaa El Hannoun, Anas Boukili Makhoukhi, Abdelhak Zoglat and Salah-Eddine El Adlouni
Water 2023, 15(14), 2641; https://doi.org/10.3390/w15142641 - 20 Jul 2023
Cited by 2 | Viewed by 1948
Abstract
Intensity–duration–frequency (IDF) curves of precipitation are a reference decision support tool used in hydrology. They allow the estimation of extreme precipitation and its return periods. Typically, IDF curves are estimated using univariate frequency analysis of the maximum annual intensities of precipitation for different [...] Read more.
Intensity–duration–frequency (IDF) curves of precipitation are a reference decision support tool used in hydrology. They allow the estimation of extreme precipitation and its return periods. Typically, IDF curves are estimated using univariate frequency analysis of the maximum annual intensities of precipitation for different durations. It is then assumed that the annual maxima of different durations are independent to simplify the parameter estimation. This strong hypothesis is not always verified for every climatic region. This study examines the effects of the independence hypothesis by proposing a multivariate model that considers the dependencies between precipitation intensities of different durations. The multivariate model uses D-vine copulas to explore the intraduration dependencies. The generalized extreme values distribution (GEV) is considered a marginal model that fits a wide range of tail behaviors. An illustration of the proposed approach is made for historical data from Moncton, in the province of New Brunswick (Eastern Canada), with climatic projections made through three scenarios of the Representative Concentration Pathway (RCP). Full article
(This article belongs to the Section Urban Water Management)
Show Figures

Figure 1

Figure 1
<p>Directed acyclic graph for a vine copula model with six durations showing the hierarchical representation by bivariate copulas. In tree 1, the first node (F1) represents the distribution of one of the four subhourly rainfall intensities (<math display="inline"><semantics><msub><mi>d</mi><mn>1</mn></msub></semantics></math>, <math display="inline"><semantics><msub><mi>d</mi><mn>2</mn></msub></semantics></math>, <math display="inline"><semantics><msub><mi>d</mi><mn>3</mn></msub></semantics></math> or <math display="inline"><semantics><msub><mi>d</mi><mn>4</mn></msub></semantics></math>), and the rest of the nodes <math display="inline"><semantics><mrow><mo>(</mo><mi>F</mi><mn>2</mn><mo>,</mo><mo>⋯</mo><mo>,</mo><mi>F</mi><mn>6</mn><mo>)</mo></mrow></semantics></math> are the distributions of the five hourly rainfall intensities <math display="inline"><semantics><mrow><mo>(</mo><msub><mi>d</mi><mn>5</mn></msub><mo>,</mo><mo>⋯</mo><mo>,</mo><msub><mi>d</mi><mn>9</mn></msub><mo>)</mo></mrow></semantics></math>. The edges of tree k, <math display="inline"><semantics><mrow><mi>k</mi><mo>=</mo><mn>1</mn><mo>,</mo><mo>⋯</mo><mo>,</mo><mn>4</mn></mrow></semantics></math> become nodes in trees <span class="html-italic">k</span> + 1, i.e., bivariate copulas for tree 2 and conditional bivariate copulas for the rest.</p>
Full article ">Figure 2
<p>Moncton weather station (Canada).</p>
Full article ">Figure 3
<p>Monthly average of total precipitation records for the Moncton location for the 1899–2010 period.</p>
Full article ">Figure 4
<p>Annual maximum rainfall intensity densities at Moncton station (1946–2016).</p>
Full article ">Figure 5
<p>Comparison between the quantiles obtained from the typical GEV model and quantile points from the copula method.</p>
Full article ">Figure 6
<p>Comparison between the IDF curves obtained from the typical IDF empirical formula and IDF points from the copula method.</p>
Full article ">Figure 7
<p>Spatial interpolation of the IDF curves for the RCP2.6 annual maximum series given by the proposed model for the Moncton location.</p>
Full article ">
19 pages, 4746 KiB  
Article
Bartlett–Lewis Model Calibrated with Satellite-Derived Precipitation Data to Estimate Daily Peak 15 Min Rainfall Intensity
by Md. Atiqul Islam, Bofu Yu and Nick Cartwright
Atmosphere 2023, 14(6), 985; https://doi.org/10.3390/atmos14060985 - 6 Jun 2023
Viewed by 2142
Abstract
Temporal variability of rainfall is extreme in the rangelands of northern Australia and occurs at annual, decadal, and even longer timescales. To maintain long-term productivity of the rangelands of northern Australia under highly variable rainfall conditions, suitable land management practices are assessed using [...] Read more.
Temporal variability of rainfall is extreme in the rangelands of northern Australia and occurs at annual, decadal, and even longer timescales. To maintain long-term productivity of the rangelands of northern Australia under highly variable rainfall conditions, suitable land management practices are assessed using rangeland biophysical models, e.g., GRASP (GRASs Production). The daily maxima of the 15 min rainfall intensity (I15) are used to predict runoff and moisture retention in the model. The performance of rangeland biophysical models heavily relies on the I15 estimates. As the number of pluviograph stations is very limited in northern Australian rangelands, an empirical I15 model (Fraser) was developed using readily available daily climate variables, i.e., daily rainfall total, daily diurnal temperature range, and daily minimum temperature. The aim of this study is to estimate I15 from daily rainfall totals using a well-established disaggregation scheme coupled with the Bartlett–Lewis rectangular pulse (BLRP) model. In the absence of pluviograph data, the BLRP models (RBL-E and RBL-G) were calibrated with the precipitation statistics estimated using the Integrated Multi-satellitE Retrievals for GPM (global precipitation measurement) (IMERG; 30 min, 0.1° resolution) precipitation product. The Fraser, RBL-E, and RBL-G models were assessed using 1 min pluviograph data at a single test site in Darwin. The results indicated that all three models tended to underestimate the observed I15, while a serious underestimation was observed for RBL-E and RBL-G. The underestimation by the Fraser, RBL-E, and RBL-G models consisted of 23%, 38%, and 50% on average, respectively. Furthermore, the Fraser model represented 29% of the variation in observed I15, whereas RBL-E and RBL-G represented only 7% and 11% of the variation, respectively. A comparison of RBL-E and RBL-G suggested that the difference in the spatial scales of IMERG and pluviograph data needs to be addressed to improve the performance of RBL-E and RBL-G. Overall, the findings of this study demonstrate that the BLRP model calibrated with IMERG statistics has the potential for estimating I15 for the GRASP biophysical model once the scale difference between IMERG and point rainfall data is addressed. Full article
(This article belongs to the Special Issue Precipitation Observations and Prediction)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the selected site in Australia and (<b>b</b>) the IMERG grids (0.1° × 0.1°) around the selected site with the nearest pixel center.</p>
Full article ">Figure 2
<p>Schematic of the Bartlett–Lewis rectangular pulse model. Note that the model allows overlapping of storms and cells.</p>
Full article ">Figure 3
<p>Flowchart illustrating the adjustment of 30 min IMERG data.</p>
Full article ">Figure 4
<p>Scatterplots of (<b>a</b>) wet-period mean (mm), (<b>b</b>) wet-period variance (mm<sup>2</sup>), (<b>c</b>,<b>d</b>) wet-period lag-1 auto-covariance (mm<sup>2</sup>), (<b>e</b>) proportion dry, and (<b>f</b>) wet-period skewness of the original-IMERG and observed datasets at the six aggregation intervals from 0.5 to 24 h. Panels (<b>c</b>,<b>d</b>) show negative and positive wet-period lag-1 auto-covariances, respectively. The circles and triangles represent the statistics over the summer half of the year (November–April) and winter half, respectively. The face colour of the circles and triangles indicates different aggregation intervals. Each colour has six circles and six triangles representing six months of each season. Note the log–log scale in panels (<b>a</b>,<b>b</b>,<b>d</b>).</p>
Full article ">Figure 5
<p>As in <a href="#atmosphere-14-00985-f004" class="html-fig">Figure 4</a>, but for corrected-IMERG dataset.</p>
Full article ">Figure 6
<p>Scatterplots of (<b>a</b>) wet-period mean intensity (mm/h), (<b>b</b>) wet-period standard deviation (mm/h), (<b>c</b>) wet-period lag-1 auto-correlation, (<b>d</b>) wet-period fraction, and (<b>e</b>) wet-period skewness of the disaggregated (RBL-E) and observed datasets at the seven aggregation intervals from 0.25 to 24 h. The circles and triangles represent the statistics over the summer half of the year (November–April) and winter half, respectively. The face colour of the circles and triangles indicates different aggregation intervals. Each colour has six circles and six triangles representing six months of each season. Note the log–log scale in panels (<b>a</b>,<b>b</b>,<b>d</b>).</p>
Full article ">Figure 7
<p>As in <a href="#atmosphere-14-00985-f006" class="html-fig">Figure 6</a>, but for disaggregated (RBL-G) dataset.</p>
Full article ">Figure 8
<p>Comparison of the observed and predicted I15 estimates from (<b>a</b>) Fraser, (<b>b</b>) RBL-E, and (<b>c</b>) RBL-G models for the days with rainfall equal to or greater than 15 mm. Performance metrics are shown in the upper left corner of each panel. The solid line shows the 1:1 line, while the dashed line shows the linear regression line between the observed and predicted estimates.</p>
Full article ">Figure 9
<p>Comparison of the ECDFs of I15 estimates obtained using the Fraser, RBL-E, and RBL-G models with the ECDF of observed I15. The green- and magenta-shaded areas show ECDFs of the 95% confidence intervals (CI) of the I15 values estimated using the RBL-E and RBL-G models, respectively. The panels are drawn considering days with rainfall equal to or greater than 15 mm (<b>a</b>) for 0–240 mm/h and (<b>b</b>) zoomed into the intensity range 40–240 mm/h. On the x-axis, the rainfall intensity is overlapped with different scales on the y-axis.</p>
Full article ">
Back to TopTop