[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (577)

Search Parameters:
Keywords = flash flood

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 3966 KiB  
Article
The Hydrologic Mitigation Effectiveness of Bioretention Basins in an Urban Area Prone to Flash Flooding
by Brian G. Laub, Eugene Von Bon, Lani May and Mel Garcia
Water 2024, 16(18), 2597; https://doi.org/10.3390/w16182597 - 13 Sep 2024
Viewed by 266
Abstract
The hydrologic performance and cost-effectiveness of green stormwater infrastructure (GSI) in climates with highly variable precipitation is an important subject in urban stormwater management. We measured the hydrologic effects of two bioretention basins in San Antonio, Texas, a growing city in a region [...] Read more.
The hydrologic performance and cost-effectiveness of green stormwater infrastructure (GSI) in climates with highly variable precipitation is an important subject in urban stormwater management. We measured the hydrologic effects of two bioretention basins in San Antonio, Texas, a growing city in a region prone to flash flooding. Pre-construction, inflow, and outflow hydrographs of the basins were compared to test whether the basins reduced peak flow magnitude and altered the metrics of flashiness, including rate of flow rise and fall. We determined the construction and annual maintenance cost of one basin and whether precipitation magnitude and antecedent moisture conditions altered hydrologic mitigation effectiveness. The basins reduced flashiness when comparing inflow to outflow and pre-construction to outflow hydrographs, including reducing peak flow magnitudes by >80% on average. Basin performance was not strongly affected by precipitation magnitude or antecedent conditions, though the range of precipitation magnitudes sampled was limited. Construction costs were higher than previously reported projects, but annual maintenance costs were similar and no higher than costs to maintain an equivalent landscaped area. Results indicate that bioretention basins effectively mitigate peak flow and flashiness, even in flash-flood-prone environments, which should benefit downstream ecosystems. The results provide a unique assessment of bioretention basin performance in flash-flood-prone environments and can inform the optimization of cost-effectiveness when implementing GSI at watershed scales in regions with current or future similar precipitation regimes. Full article
(This article belongs to the Section Urban Water Management)
Show Figures

Figure 1

Figure 1
<p>Map showing (<b>a</b>) the location of the Leon Creek watershed, San Antonio, and the Edwards Aquifer zones within the state of Texas, (<b>b</b>) the location of the central and west campus bioretention basins on the UTSA campus, (<b>c</b>) the UTSA campus and Leon Creek watershed along with the Edwards Aquifer zones, and (<b>d</b>) schematic diagram of the central campus basin showing the north and south basins divided by an earthen berm and connected by an overflow pipe from the north basin to the south basin. The red areas in (<b>c</b>) show urban developed land in and around the city of San Antonio. Also shown in (<b>d</b>) are major inflow points and the sump housing where water is pumped out of the basin as outflow. The contour lines in (<b>d</b>) are 0.3 m.</p>
Full article ">Figure 2
<p>Example showing how with- and without-basin hydrographs were constructed from changes in water depth over time in the bioretention basins. Panel (<b>a</b>) shows the recorded changes in depth (black line) in the south basin during a runoff event on 3 November 2021. An increase in depth represents an inflow to the basin (highlighted by orange arrows), which would have passed downstream as flow without the basin in place. The decrease in depth represents the draining of the basin (highlighted by blue arrow), which was pumped downstream out of the basin. Panel (<b>b</b>) shows the resulting flow rate that would have occurred downstream of the basin without the basin (orange line) and the actual flow rate with the basin in place (blue line).</p>
Full article ">Figure 3
<p>Box plots comparing flow metrics between pre-construction (Pre), with-basin, and without-basin hydrographs for the central campus basin. Boxes show 25th and 75th percentile (interquartile range) with the dark line indicating the median. Whiskers extend ±1.5 times the interquartile range, with values outside whiskers indicated as individual points.</p>
Full article ">Figure 4
<p>Box plots comparing with-basin and without-basin flow metrics for the west campus basin. Boxes show 25th and 75th percentile (interquartile range) with the dark line indicating the median. Whiskers extend ±1.5 times the interquartile range, with values outside whiskers indicated as individual points.</p>
Full article ">
23 pages, 16575 KiB  
Article
Remote Sensing of Floodwater-Induced Subsurface Halite Dissolution in a Salt Karst System, with Implications for Landscape Evolution: The Western Shores of the Dead Sea
by Gidon Baer, Ittai Gavrieli, Iyad Swaed and Ran N. Nof
Remote Sens. 2024, 16(17), 3294; https://doi.org/10.3390/rs16173294 - 4 Sep 2024
Viewed by 631
Abstract
We study the interrelations between salt karst and landscape evolution at the Ze’elim and Hever alluvial fans, Dead Sea (DS), Israel, in an attempt to characterize the ongoing surface and subsurface processes and identify future trends. Using light detection and ranging, interferometric synthetic [...] Read more.
We study the interrelations between salt karst and landscape evolution at the Ze’elim and Hever alluvial fans, Dead Sea (DS), Israel, in an attempt to characterize the ongoing surface and subsurface processes and identify future trends. Using light detection and ranging, interferometric synthetic aperture radar, drone photography, time-lapse cameras, and direct measurements of floodwater levels, we document floodwater recharge through riverbed sinkholes, subsurface salt dissolution, groundwater flow, and brine discharge at shoreline sinkholes during the years 2011–2023. At the Ze’elim fan, most of the surface floodwater drains into streambed sinkholes and discharges at shoreline sinkholes, whereas at the Hever fan, only a small fraction of the floodwater drains into sinkholes, while the majority flows downstream to the DS. This difference is attributed to the low-gradient stream profiles in Ze’elim, which enable water accumulation and recharge in sinkholes and their surrounding depressions, in contrast with the higher-gradient Hever profiles, which yield high-energy floods capable of carrying coarse gravel that eventually fill the sinkholes. The rapid drainage of floodwater into sinkholes also involves slope failure due to pore-pressure drop and cohesion loss within hours after each drainage event. Surface subsidence lineaments detected by InSAR indicate the presence of subsurface dissolution channels between recharge and discharge sites in the two fans and in the nearby Lynch straits. Subsidence and streambed sinkholes occur in most other fans and streams that flow to the DS; however, with the exception of Ze’elim, all other streams show only minor or no recharge along their course. This is due to either the high-gradient profiles, the gravelly sediments, the limited floods, or the lack of conditions for sinkhole development in the other streambeds. Thus, understanding the factors that govern the flood-related karst formation is of great importance for predicting landscape evolution in the DS region and elsewhere and for sinkhole hazard assessment. Full article
(This article belongs to the Special Issue Remote Sensing of the Dead Sea Region)
Show Figures

Figure 1

Figure 1
<p>LiDAR topography of the two study areas draped upon hill−shaded DSMs. (<b>a</b>) Ze’elim fan. Gully numbers (in white) are after [<a href="#B3-remotesensing-16-03294" class="html-bibr">3</a>]. (<b>b</b>) Hever fan. (<b>c</b>) Location maps of the study areas.</p>
Full article ">Figure 2
<p>Elevation profiles along Ze’elim and Hever riverbeds, May 2020. Note the low gradients (1−3%) and fine-grained composition of the Ze’elim riverbeds (blue-green profiles, for location, see <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>a), in contrast with the high gradients (3−4.5%) and coarse gravel sediments of the Hever riverbeds (red-brown profiles; for location, see <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>b), which decrease to 1−1.5% only at their easternmost parts.</p>
Full article ">Figure 3
<p>(<b>a</b>) Time-lapse camera and drone, overlooking gully 14 recharge sinkhole. The blue arrow marks the flow direction from west to east. (<b>b</b>) View south at the DSW canal as floodwater crosses the overpasses. Photo courtesy of DSW. (<b>c</b>) Locations of hydrometers (marked by white arrows) that are installed at an overpass. E and W mark eastern and western hydrometers.</p>
Full article ">Figure 4
<p>Photos of recharge sinkholes at Ze’elim fan streambeds. Blue arrows mark the flow direction. Ab—abandoned gullies, overhanging downstream of the recharge sinkholes. For location, see <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>a. (<b>a</b>) Gully 6. (<b>b</b>) Gully 7. (<b>c</b>) Gully 13. (<b>d</b>) Gully 14. Drone picture was taken by Liran Ben Moshe.</p>
Full article ">Figure 5
<p>(<b>a</b>) Floodwater recharge (red circles) and discharge (blue rectangles) sites at the Ze’elim fan. Red and yellow triangles mark locations and operation intervals of the TLCs. Yellow numbers denote gully numbers (after [<a href="#B3-remotesensing-16-03294" class="html-bibr">3</a>]). (<b>b</b>) Drone photograph, 2 January 2020, showing the discharge sinkholes (rectangles) and TLCs (triangles) within and around the shoreline sinkholes of gully 10. Note that not all TLCs operated simultaneously.</p>
Full article ">Figure 6
<p>Discharge sinkhole 10a (see location in <a href="#remotesensing-16-03294-f005" class="html-fig">Figure 5</a>b). (<b>a</b>) View east, February 2024. (<b>b</b>) TLC picture showing water discharge following the 25 March 2019 flood. (<b>c</b>) The nested sinkhole at the northern wall of the major sinkhole, February 2024, exposing the “Sinkhole Salt” layer (white layers with small cavities), the dissolution channel openings, and groundwater flow.</p>
Full article ">Figure 7
<p>Sinkholes and subsidence along the course of gully 3. (<b>a</b>) LiDAR DSM, July 2023 (see location in <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>). The location of profile A−A’ (panel <b>d</b>) is shown in a dashed white line. The area around the recharge sinkhole is marked by a circle. (<b>b</b>) Drone photograph of the recharge area (sinkhole marked by white circle), January 2024, taken by Liran Ben Moshe. (<b>c</b>) Streambed sinkhole recharging floodwater after the 15.2.2024 flood. (<b>d</b>) Elevation profile A−A’ along the gully, September 2023 (location shown in panel <b>a</b>). Note that this recharge sinkhole does not appear in 2020 (see Ze’elim 3 profile in <a href="#remotesensing-16-03294-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 8
<p>Water levels at five overpasses during the 21–22 November 2021 flood in Ze’elim. See inset for location. The overpasses are marked by white numbers, and streams are marked by yellow numbers. E and W stand for eastern and western hydrographs, respectively (<a href="#remotesensing-16-03294-f003" class="html-fig">Figure 3</a>c).</p>
Full article ">Figure 9
<p>An interferogram of the Ze’elim fan spanning 44 days in early 2024, showing subsidence lineaments that are interpreted as surface manifestations of subsurface dissolution channels. The two acquisition times are 13 January 2024 and 26 February 2024. WZSL, EZSL, and NZSL stand for western, eastern, and northern Ze’elim subsidence lineaments, shown by white, orange, and yellow arrows, respectively. Gully numbers are marked in white (after [<a href="#B3-remotesensing-16-03294" class="html-bibr">3</a>]).</p>
Full article ">Figure 10
<p>Surface (dashed white lines) and proposed subsurface water pathways (dashed yellow lines) in Ze’elim: a southern pathway from gully 14 to the western side of sinkhole 10 (10a in <a href="#remotesensing-16-03294-f005" class="html-fig">Figure 5</a>b), and central pathways from gullies 5, 6, and 7 to the eastern side of sinkhole 10 (10f in <a href="#remotesensing-16-03294-f005" class="html-fig">Figure 5</a>b). Gully numbers are after [<a href="#B3-remotesensing-16-03294" class="html-bibr">3</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>,<b>b</b>) Annual surface elevation changes in the Hever fan, draped upon LiDAR DSMs. River incision, riverbank collapse, subsidence, and sinkholes are displayed by negative (blue) values. Aggradation of alluvial material along the streambeds and within sinkholes is displayed by positive (red) values and by white arrows. White ellipses mark subsidence around sinkhole clusters. The black arrow in (<b>a</b>) points at a meandering subsidence lineament, interpreted as the surface manifestation of a subsurface dissolution channel. (<b>c</b>) Interferogram showing sinkhole-related subsidence (semi-circular fringe colors) and a meandering subsidence lineament (marked by white arrows) that is interpreted as the surface manifestation of a subsurface dissolution channel between the western cluster of recharging sinkholes and the eastern subsidence zone (similar to the lineament in panel <b>a</b>). The acquisition times of the two images are 7 June 2018 and 18 June 2018.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) TLC photos showing recharge of floodwater at sinkholes in the northern branch of Hever fan during the 20 February 2015 flood. Blue arrows mark the braided streambed flow direction. Note the water overflow and the filling of the sinkhole with gravel at the final hours of the flood (panels (<b>b</b>) and (<b>c</b>), respectively). (<b>d</b>) Drone picture of 7 February 2019 floodwater drained into recharge sinkholes along the northern Hever branch with overflowing water continuing downstream.</p>
Full article ">Figure 13
<p>(<b>a</b>) Discharge sinkholes at the lower part of the southern Hever branch. (<b>b</b>) Offshore discharge sites at the Hever shoreline (white arrows). (<b>c</b>) A small salt chimney 2 m offshore in Hever. (<b>d</b>) <span class="html-italic">Anabasis setifera</span> vegetation at the lower Hever southern streambed. For location, see the black arrow in <a href="#remotesensing-16-03294-f011" class="html-fig">Figure 11</a>a.</p>
Full article ">Figure 14
<p>Linear and meandering subsidence patterns at the Lynch straits (for location, see <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>c). (<b>a</b>) Subtraction map of LiDAR DSMs between 2023 and 2014. (<b>b</b>) Interferogram between 2 and 13 April 2018. The white arrows point to subsidence lineaments that are interpreted to form above subsurface salt dissolution channels.</p>
Full article ">Figure 15
<p>Stream gradients along the western shoreline of the DS. Color legend distinguishes between low-gradient, mud-dominated streams (green); high-gradient, gravel-dominated streams (red); and intermediate-gradient mixed mud-gravel streams (yellow).</p>
Full article ">
21 pages, 6160 KiB  
Article
Challenges of Using a Geographic Information System (GIS) in Managing Flash Floods in Shah Alam, Malaysia
by Adam Narashman Leeonis, Minhaz Farid Ahmed, Mazlin Bin Mokhtar, Chen Kim Lim and Bijay Halder
Sustainability 2024, 16(17), 7528; https://doi.org/10.3390/su16177528 - 30 Aug 2024
Viewed by 955
Abstract
A geographic information system (GIS) is a tool and technology capable of addressing the effects and challenges of natural disasters, particularly flash floods. GIS applications are used to generate flood risk maps to tackle flood issues. However, various challenges and problems arise when [...] Read more.
A geographic information system (GIS) is a tool and technology capable of addressing the effects and challenges of natural disasters, particularly flash floods. GIS applications are used to generate flood risk maps to tackle flood issues. However, various challenges and problems arise when employing GIS to manage flash flood disasters in Shah Alam, Malaysia. Hence, this study aims to identify these challenges and gaps in GIS utilisation by Malaysian agencies for flash flood management in Shah Alam. Using the quadruple helix model technique, informal interviews were conducted as part of the study’s qualitative methodology. Five respondents were chosen from each of the four main sectors for primary data collection: government, academia, business, and community/NGO. The data were analysed using Taguette qualitative theme analysis. The findings reveal that the primary challenges lie in government management, particularly in providing equipment and access to GIS for all stakeholders, including the public. This challenge is attributed to the high costs and complexity associated with GIS data usage, limiting accessibility. Furthermore, there is a lack of expertise and research on GIS in Malaysian universities concerning flash flood management. The government should take proactive steps to improve flash flood management in Shah Alam, Malaysia, in order to solve these issues. Specifically, GIS training should be given to stakeholders, particularly those in the government and academic sectors, in order to develop GIS specialists who will be necessary for efficient flood management in Malaysia. Full article
(This article belongs to the Special Issue Sustainable Resilience Planning for Natural Hazard Events)
Show Figures

Figure 1

Figure 1
<p>Co-occurrence authors keywords for flash flooding in Malaysia based on the Scopus database and VOSviewer software version 1.6.19.</p>
Full article ">Figure 2
<p>Map of Shah Alam, Selangor, Malaysia (Source: Laboratory work, 2022).</p>
Full article ">Figure 3
<p>Location map of Shah Alam, Selangor, Malaysia (Source: Shah Alam City Council, 2022).</p>
Full article ">Figure 4
<p>Quadruple helix model based on four sectors.</p>
Full article ">Figure 5
<p>Proposed framework of challenges in the use of GIS in flash flood management in Shah Alam, Malaysia.</p>
Full article ">
28 pages, 37291 KiB  
Article
Probabilistic Cascade Modeling for Enhanced Flood and Landslide Hazard Assessment: Integrating Multi-Model Approaches in the La Liboriana River Basin
by Johnny Vega, Laura Ortiz-Giraldo, Blanca A. Botero, César Hidalgo and Juan Camilo Parra
Water 2024, 16(17), 2404; https://doi.org/10.3390/w16172404 - 27 Aug 2024
Viewed by 530
Abstract
Extreme rainfall events in Andean basins frequently trigger landslides, obstructing river channels and causing flash flows, loss of lives, and economic damage. This study focused on improving the modeling of these events to enhance risk management, specifically in the La Liboriana basin in [...] Read more.
Extreme rainfall events in Andean basins frequently trigger landslides, obstructing river channels and causing flash flows, loss of lives, and economic damage. This study focused on improving the modeling of these events to enhance risk management, specifically in the La Liboriana basin in Salgar (Colombia). A cascading modeling methodology was developed, integrating the spatially distributed rainfall intensities, hazard zoning with the SLIDE model, propagation modeling with RAMMS using calibrated soil rheological parameters, the distributed hydrological model TETIS, and flood mapping with IBER. Return periods of 2.33, 5, 10, 25, 50, and 100 years were defined and applied throughout the methodology. A specific extreme event (18 May 2015) was modeled for calibration and comparison. The spatial rainfall intensities indicated maximum concentrations in the northwestern upper basin and southeastern lower basin. Six landslide hazard maps were generated, predicting landslide-prone areas with a slightly above random prediction rate for the 2015 event. The RAMMS debris flow modeling involved 30 simulations, indicating significant deposition within the river channel and modifying the terrain. Hydraulic modeling with the IBER model revealed water heights ranging from 0.23 to 7 m and velocities from 0.34 m/s to 6.98 m/s, with urban areas showing higher values, indicating increased erosion and infrastructure damage potential. Full article
(This article belongs to the Section Hydrogeology)
Show Figures

Figure 1

Figure 1
<p>Location map of the study area: (<b>a</b>) continental scale; (<b>b</b>) country scale; and (<b>c</b>) basin scale.</p>
Full article ">Figure 2
<p>Characterization of the study area: (<b>a</b>) elevation; (<b>b</b>) slope; (<b>c</b>) landforms; (<b>d</b>) soil depth (thickness); (<b>e</b>) geology; and (<b>f</b>) landcover.</p>
Full article ">Figure 3
<p>Schematic workflow for the multi-hazard assessment.</p>
Full article ">Figure 4
<p>Rainfall intensities (m/h) for different return periods.</p>
Full article ">Figure 5
<p>Landslide hazard assessment (in terms of the FoS values) considering different return periods.</p>
Full article ">Figure 6
<p>Maximum deposition height (m) according for the considered return period.</p>
Full article ">Figure 7
<p>Analysis of the deposition heights (m) in the river section k18+910 according to the considered return periods.</p>
Full article ">Figure 8
<p>Location of the simulation points (flow hydrographs) for the hydraulic model and rainfall depths (mm) for the considered return periods.</p>
Full article ">Figure 9
<p>Torrential flood event hydrograph simulated with the TETIS model.</p>
Full article ">Figure 10
<p>Landcover of the flood zone along the La Liboriana river, modified based on the analysis of aerial photographs.</p>
Full article ">Figure 11
<p>Flow hydrographs at the reach outlet. Results comparison of the IBER and TETIS models.</p>
Full article ">Figure 12
<p>Hydraulic modeling results for the torrential flood event of 18 May 2015. (<b>a</b>,<b>b</b>) Maximum water height (m); (<b>c</b>) sites with a water level reported in Velasquez et al. [<a href="#B38-water-16-02404" class="html-bibr">38</a>] used for calibrating the hydraulic model; and (<b>d</b>,<b>e</b>) maximum velocity (m/s).</p>
Full article ">Figure 13
<p>Hydraulic modeling results for a 100-year return period torrential flood for La Margarita rural zone and the urban area of Salgar: (<b>top</b>) maximum water height (m); and (<b>bottom</b>) maximum velocity (m/s).</p>
Full article ">
20 pages, 5118 KiB  
Article
Co-Occurrence of Cyanotoxins and Phycotoxins in One of the Largest Southeast Asian Brackish Waterbodies: A Preliminary Study at the Tam Giang—Cau Hai Lagoon (Vietnam)
by Devleena Sahoo, Ngoc Khanh Ni Tran, Thi Gia-Hang Nguyen, Thi Thu Hoai Ho, Thi Thuy Hang Phan, Duong Thu Huong Hoang, Ngo Huu Binh, Thi Thu Lien Nguyen, Luong Quang Doc, Noureddine Bouaïcha and Tri Nguyen-Quang
Limnol. Rev. 2024, 24(3), 335-353; https://doi.org/10.3390/limnolrev24030020 - 25 Aug 2024
Viewed by 482
Abstract
The Tam Giang-Cau Hai lagoon (TGCH) in Thua Thien Hue province (Vietnam) is a marsh/lagoon system and ranks among the largest waterbodies in Southeast Asia. It plays a significant role in terms of both socio-economic and environmental resources. However, anthropogenic stress, as well [...] Read more.
The Tam Giang-Cau Hai lagoon (TGCH) in Thua Thien Hue province (Vietnam) is a marsh/lagoon system and ranks among the largest waterbodies in Southeast Asia. It plays a significant role in terms of both socio-economic and environmental resources. However, anthropogenic stress, as well as the discharge of untreated domestic and industrial sewage with agricultural runoff from its three major tributaries, dramatically damages the water quality of the lagoon. Especially after heavy rain and flash floods, the continuous degradation of its water quality, followed by harmful algal and cyanobacterial bloom patterns (HABs), is more perceptible. In this study, several physicochemical factors, cyanotoxins (anatoxins (ATXs), saxitoxins (STXs), microcystins (MCs)), phycotoxins (STXs, okadaic acid (OA), and dinophysistoxins (DTXs)) were analyzed in water and shellfish samples from 13 stations in June 2023 from 13 stations, using enzyme-linked immunosorbent assay (ELISA) kits for the ATXs and STXs, and the serine/threonine phosphatase type 2A (PP2A) inhibition assay kit for the MCs, OA, and DTXs. The results showed for the first time the co-occurrence of freshwater cyanotoxins and marine phycotoxins in water and shellfish samples in this lagoon. Traces of ATXs and STXs were detected in the shellfish and the orders of magnitude were below the seafood safety action levels. However, toxins inhibiting the PP2A enzyme, such as MCs and nodularin (NODs), as well as OA and DTXs, were detected at higher concentrations (maximum: 130.4 μg equiv. MC-LR/kg shellfish meat wet weight), approaching the actionable level proposed for this class of toxin in shellfish (160 μg of OA equivalent per kg of edible bivalve mollusk meat). It is very important to note that due to the possible false positives produced by the ELISA test in complex matrices such as a crude shellfish extract, this preliminary and pilot research will be repeated with a more sophisticated method, such as liquid chromatography coupled with mass spectroscopy (LC-MS), in the upcoming research plan. Full article
(This article belongs to the Special Issue Hot Spots and Topics in Limnology)
Show Figures

Figure 1

Figure 1
<p>Sampling locations with a zoomed view of Tam Giang lagoon, Vietnam.</p>
Full article ">Figure 2
<p>A local market in the fishery village at Tam Giang lagoon. Below: Seven different batches of shellfish samples collected from the lagoon at TC 13, including: (1) <span class="html-italic">Cyrenobatissa subsulcata</span>; (2) <span class="html-italic">Corbicula subsulcata</span>; (3) <span class="html-italic">Cristaria plicata</span>; (4) <span class="html-italic">Pila polita</span>; (5, 6) <span class="html-italic">Crasscostrea rivularis</span>; (7) <span class="html-italic">Perna viridis</span>.</p>
Full article ">Figure 3
<p>(<b>A</b>) Secchi depth from Stations TC1 to TC13; (<b>B</b>) trend of dissolved oxygen (DO), pH, and Salinity in TG-CH Lagoon.</p>
Full article ">Figure 4
<p>Relative abundances of phytoplankton from all phyla identified at 13 sites in the Tam Giang lagoon (Vietnam) during June 2023.</p>
Full article ">Figure 5
<p>Concentrations (µg/L) of paralytic shellfish toxins (PSTs) expressed as saxitoxin (STX) equivalent in water samples collected from the different sites (TC1 to TC13) in the Tam Giang lagoon, Vietnam. The regulatory guidance level for saxitoxin is 3 µg/L in drinking water and 30 µg/L in recreational water.</p>
Full article ">Figure 6
<p>Concentrations (µg/L) of anatoxins (ATXs) expressed as anatoxin-a (ATX-a) equivalent in water samples collected from the different sites (TC1 to TC13) in the Tam Giang lagoon, Vietnam. The regulatory guidance level for ATX-a is 30 µg/L in drinking water and 60 µg/L in recreational water.</p>
Full article ">Figure 7
<p>Concentrations (µg/L) of toxins inhibiting the PP2A, such as microcystins (MCs), nodularin (NODs), okadaic acid (OA), and dinophysistoxins (DTXs), expressed as microcystin-LR (MC-LR) in water samples collected from the different sites (TC1 to TC13) in the Tam Giang lagoon, Vietnam. The regulatory guidance level for MC-LR is 1 µg/L in drinking water (dotted red line) and 24 µg/L in recreational water.</p>
Full article ">Figure 8
<p>Levels (µg/kg) of paralytic shellfish toxins (PSTs) expressed as saxitoxin (STX) equivalent in shellfish samples collected from the Tam Giang lagoon, Vietnam. The regulatory guidance level is 800 µg STX equivalent/kg.</p>
Full article ">Figure 9
<p>Levels (µg/kg) of anatoxins (ATXs) expressed as anatoxin-a (ATX-a) equivalent in shellfish samples collected from the Tam Giang lagoon, Vietnam. No regulatory guidelines are mentioned for ATX in shellfish samples.</p>
Full article ">Figure 10
<p>Levels (µg/kg) of toxins inhibiting the PP2A, such as microcystins (MCs), nodularin (NODs), okadaic acid (OA), and dinophysistoxins (DTXs), expressed as microcystin-LR (MC-LR) equivalent in shellfish samples collected from the Tam Giang lagoon, Vietnam. The regulatory guidance level (dotted red line) for diarrheic toxins in shellfish is 160 μg of OA equivalent per kg of edible bivalve mollusk meat, by total amounts of OA, DTXs, and pectenotoxins.</p>
Full article ">
28 pages, 9121 KiB  
Article
Flood Hazard and Risk Assessment of Flash Floods for Petra Catchment Area Using Hydrological and Analytical Hierarchy (AHP) Modeling
by Mustafa Al Kuisi, Naheel Al Azzam, Tasneem Hyarat and Ibrahim Farhan
Water 2024, 16(16), 2283; https://doi.org/10.3390/w16162283 - 13 Aug 2024
Viewed by 765
Abstract
Floods are a widespread natural disaster that occur in most areas of the world, except for the polar regions. To minimize the damage caused by floods, effective management strategies and policies must be implemented. Petra and Wadi Musa areas are prone to floods, [...] Read more.
Floods are a widespread natural disaster that occur in most areas of the world, except for the polar regions. To minimize the damage caused by floods, effective management strategies and policies must be implemented. Petra and Wadi Musa areas are prone to floods, which happen every 2–3 years and result in significant harm to both lives and properties. To address this issue, a composite hazard and vulnerability index is commonly utilized to evaluate flood risk and guide policy formation for flood risk reduction. These tools are efficient and cost-effective in generating accurate results. Accordingly, the present study aims to determine the morphological and hydrometeorological parameters that affect flash floods in Petra catchment area and to identify high-risk zones using GIS, hydrological, and analytical hierarchy (AHP) modeling. Nine factors, including Elevation (E), Landuse/Landcover LULC, Slope (S), Drainage density (DD), Flood Control Points (FCP) and Rainfall intensity (RI), which make up the six risk indices, and Population Density (PD), Cropland (C), and Transportation (Tr), which make up the three vulnerability indices, were evaluated both individually and in combination using AHP in ArcGIS 10.8.2 software. These parameters were classified as hazard and vulnerability indicators, and a final flood map was generated. The map indicated that approximately 37% of the total area in Petra catchment is at high or very high risk of flooding, necessitating significant attention from governmental agencies and decision-makers for flood risk mitigation. The AHP method proposed in this study is an accurate tool for flood mapping that can be easily applied to other regions in Jordan to manage and prevent flood hazards. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the study area.</p>
Full article ">Figure 2
<p>(<b>a</b>). Geological map, (<b>b</b>). soil texture, (<b>c</b>). soil hydrological group, (<b>d</b>). DEM (<b>e</b>). slope and (<b>f</b>). land use.</p>
Full article ">Figure 3
<p>Flowchart showing the methodology for this study.</p>
Full article ">Figure 4
<p>Sub-catchments of the study area.</p>
Full article ">Figure 5
<p>Annual rainfall (mm) with the gauge stations.</p>
Full article ">Figure 6
<p>Long-term annual rainfall of (<b>a</b>) Wadi Musa and (<b>b</b>) Petra rainfall gauging stations with a nine-year moving average.</p>
Full article ">Figure 7
<p>IDF curves: (<b>a</b>) Wadi Musa and (<b>b</b>) Petra rain gauging stations.</p>
Full article ">Figure 8
<p>CN distribution value for the Petra catchment.</p>
Full article ">Figure 9
<p>The flood inflow hydrographs of the Petra catchment outlet created by HEC-1 for the different return periods (5, 10, 25, 50, 100 and 1000 years) with rainfall intensities of 10, 30, 60 and 180 min and 24 h.</p>
Full article ">Figure 10
<p>Perpendicular cross sections and water depth along the Wadi course.</p>
Full article ">Figure 11
<p>Flood inundation map with 1 and 24 h’ rainfall intensity for the return periods of 10, 25, 50, 100, and 1000 years.</p>
Full article ">Figure 12
<p>The thematic standardized maps for the hazard and vulnerability indicators, (<b>a</b>). Rainfall Intensities, (<b>b</b>). Elevation, (<b>c</b>). Slope, (<b>d</b>). Flood Control Points, (<b>e</b>). Drainage Density, (<b>f</b>). Land Use/Land Cover, (<b>g</b>). Cropland, (<b>h</b>). Transportation, and (<b>i</b>). Population Density.</p>
Full article ">Figure 13
<p>Flood hazard, vulnerability and risk maps.</p>
Full article ">
23 pages, 6275 KiB  
Article
Understanding Multi-Hazard Interactions and Impacts on Small-Island Communities: Insights from the Active Volcano Island of Ternate, Indonesia
by Mohammad Ridwan Lessy, Jonatan Lassa and Kerstin K. Zander
Sustainability 2024, 16(16), 6894; https://doi.org/10.3390/su16166894 - 11 Aug 2024
Viewed by 1185
Abstract
Drawing on a case study from Ternate Island, a densely populated volcanic island in Eastern Indonesia, this research illustrates how multi-hazards and extreme weather events are likely to compound and cascade, with serious consequences for sustainable development in small island context. At the [...] Read more.
Drawing on a case study from Ternate Island, a densely populated volcanic island in Eastern Indonesia, this research illustrates how multi-hazards and extreme weather events are likely to compound and cascade, with serious consequences for sustainable development in small island context. At the heart of Ternate Island sits the active Gamalama volcano, posing a constant eruption threat. Its location within the Ring of Fire further exposes the island to the risks of tsunamis and earthquakes. Additionally, the island’s physical features make it highly susceptible to flooding, landslides, and windstorms. Rapid urbanization has led to significant coastal alterations, increasing exposure to hazards. Ternate’s small-island characteristics include limited resources, few evacuation options, vulnerable infrastructure, and inadequate resilience planning. Combining GIS multi-hazard mapping with a structured survey in 60 villages in Ternate, this case study investigates the multi-hazard exposure faced by the local population and land coverage. The findings suggest significant gaps between village chiefs’ perceptions of the types of hazards and the multi-hazard assessment in each village. Out of 60 village chiefs surveyed, 42 (70%) are aware of earthquake risks, 17 (28%) recognize tsunami threats, and 39 see volcanoes as a danger. GIS assessments show that earthquakes could impact all villages, tsunamis could affect 46 villages (77%), and volcanoes could threaten 39 villages. The hazard map indicates that 32 villages are at risk of flash floods and 37 are at risk of landslides, and extreme weather could affect all villages. Additionally, 42 coastal villages on Ternate Island face potential extreme wave and abrasion disasters, but only 18 chiefs acknowledge extreme weather as a threat. The paper argues that addressing the cognitive biases reflected in the perceptions of community leaders requires transdisciplinary dialogue and engagement. Full article
Show Figures

Figure 1

Figure 1
<p>North Maluku and Ternate Island.</p>
Full article ">Figure 2
<p>Multi-hazard risk assessment flowchart.</p>
Full article ">Figure 3
<p>Hydro-meteorological hazard maps of Ternate Island: (<b>a</b>) flash flood; (<b>b</b>) landslide; (<b>c</b>) extreme weather; and (<b>d</b>) extreme wave and abrasion.</p>
Full article ">Figure 4
<p>Geological hazard maps of Ternate Island: (<b>a</b>) earthquake; (<b>b</b>) tsunami; (<b>c</b>) volcano eruption.</p>
Full article ">Figure 5
<p>Integrated hazard map of Ternate Island.</p>
Full article ">Figure 6
<p>Population density (<b>a</b>) and land use (<b>b</b>) maps of Ternate Island.</p>
Full article ">Figure 7
<p>The vulnerability map (<b>a</b>) and multi-hazard risk map (<b>b</b>) of Ternate Island.</p>
Full article ">Figure 8
<p>Identification of hazard interactions.</p>
Full article ">Figure 9
<p>Knowledge of village officials about disaster threats (N = 60).</p>
Full article ">
20 pages, 12795 KiB  
Article
Building Reservoirs as Protection against Flash Floods and Flood Basins Management—The Case Study of the Stubo–Rovni Regional Water-Management System
by Ljubiša Bezbradica, Boško Josimović, Boris Radić, Siniša Polovina and Tijana Crnčević
Water 2024, 16(16), 2242; https://doi.org/10.3390/w16162242 - 8 Aug 2024
Viewed by 644
Abstract
Global warming and climate change cause large temperature oscillations and uneven annual rainfall patterns. The rainy cycles characterized by frequent high-intensity rainfall in the area of the Stubo–Rovni water reservoir, which in 2014 peaked at 129 mm of water in 24 h (the [...] Read more.
Global warming and climate change cause large temperature oscillations and uneven annual rainfall patterns. The rainy cycles characterized by frequent high-intensity rainfall in the area of the Stubo–Rovni water reservoir, which in 2014 peaked at 129 mm of water in 24 h (the City of Valjevo, the Republic of Serbia), caused major floods in the wider area. Such extremes negatively affect erosion processes, sediment production, and the occurrence of flash floods. The erosion coefficient before the construction of the water reservoir was Zm = 0.40, while the specific sediment production was about 916.49 m3∙km−2∙year−1. A hydrological study at the profile near the confluence of the Jadar and Obnica rivers, i.e., the beginning of the Kolubara river, the right tributary of the Sava (in the Danube river basin), indicates that the natural riverbed can accommodate flows with a 20% to 50% probability of occurrence (about 94 m3/s), while centennial flows of about 218 m3/s exceed the capacities of the natural riverbed of the Jadar river, causing flooding of the terrain and increasing risks to the safety of the population and property. The paper presents the impacts of the man-made Stubo–Rovni water reservoir on the catchment area and land use as the primary condition for preventing erosion processes (specific sediment production has decreased by about 20%, the forest cover increased by about 25%, and barren land decreased by 90%). Moreover, planned and controlled management of the Stubo–Rovni reservoir has significantly influenced the downstream flow, reducing the risks of flash floods. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the Stubo–Rovni water-management system [<a href="#B30-water-16-02242" class="html-bibr">30</a>,<a href="#B31-water-16-02242" class="html-bibr">31</a>].</p>
Full article ">Figure 2
<p>Elevation of land in the Stubo–Rovni reservoir [<a href="#B30-water-16-02242" class="html-bibr">30</a>,<a href="#B31-water-16-02242" class="html-bibr">31</a>].</p>
Full article ">Figure 3
<p>Land use in the Stubo–Rovni reservoir [<a href="#B30-water-16-02242" class="html-bibr">30</a>,<a href="#B31-water-16-02242" class="html-bibr">31</a>].</p>
Full article ">Figure 4
<p>Slopes of the Stubo–Rovni reservoir basin [<a href="#B30-water-16-02242" class="html-bibr">30</a>,<a href="#B31-water-16-02242" class="html-bibr">31</a>,<a href="#B32-water-16-02242" class="html-bibr">32</a>].</p>
Full article ">Figure 5
<p>Soil erosion in the Jablanica river basin up to the Stubo–Rovni profile [<a href="#B30-water-16-02242" class="html-bibr">30</a>,<a href="#B31-water-16-02242" class="html-bibr">31</a>,<a href="#B32-water-16-02242" class="html-bibr">32</a>].</p>
Full article ">Figure 6
<p>Specific erosion production in the Jablanica river basin from the Stubo–Rovni profile [<a href="#B30-water-16-02242" class="html-bibr">30</a>,<a href="#B31-water-16-02242" class="html-bibr">31</a>,<a href="#B32-water-16-02242" class="html-bibr">32</a>].</p>
Full article ">Figure 7
<p>Sketch of the natural riverbed, banks, and flow of the Jablanica river.</p>
Full article ">
17 pages, 9646 KiB  
Article
Declining Bank Erosion Rate Driven by Hydrological Alterations of a Small Sub-Alpine River
by Alexandra Pusztai-Eredics and Tímea Kiss
Hydrology 2024, 11(8), 114; https://doi.org/10.3390/hydrology11080114 - 31 Jul 2024
Viewed by 702
Abstract
In the 21st century, climate change and its consequences are getting more serious. The changes in temperature and precipitation alter the run-off conditions, subsequently influencing the channel processes of rivers. The study aims to analyse the hydrological changes in a small, sub-alpine river [...] Read more.
In the 21st century, climate change and its consequences are getting more serious. The changes in temperature and precipitation alter the run-off conditions, subsequently influencing the channel processes of rivers. The study aims to analyse the hydrological changes in a small, sub-alpine river (Rába/Raab River, Central Europe) and the bank erosional processes (1951–2024). The bank erosion was determined based on topographical maps, aerial photographs, and field (RTK–GPS) surveys. Short (2–3 days) floods were common between 1950 and 1980, and low stages occurred in 65–81% of a year. However, extreme regimes developed in the 21st century, as record-high, flash floods altered with long low stages (91–96% of a year). The bank erosion shows a cyclic temporal pattern, gradually increasing until it reaches a high value (4.1–4.9 m/y), followed by a limited erosional rate (2.2–2.8 m/y). However, the magnitude of the bank erosion is decreasing. This could be explained by (1) the lower transport capacity of the more common low stages and (2) the seasonal shift of the flood waves, which appear in the growing season when the riparian vegetation can more effectively protect the banks from erosion. Full article
Show Figures

Figure 1

Figure 1
<p>The catchment of the Rába/Raab River is located in Central Europe (<b>A</b>). Its middle reach was studied in detail (<b>B</b>). The studied reach was divided into 14 units (U1–14) based on the degree of human impacts, and along some meanders (meander ID: A–U), the bank erosion was measured using an RTK–GPS (<b>C</b>).</p>
Full article ">Figure 2
<p>The mean bank erosion was calculated based on the length of lines perpendicular to a former bankline (<b>A</b>). The maximum bank erosion was calculated within circles (<b>B</b>).</p>
Full article ">Figure 3
<p>Annual highest, mean, and minimum water stages (<b>A</b>) and annual highest, mean, and minimum discharges (<b>B</b>) measured at Szentgotthárd gauging station.</p>
Full article ">Figure 4
<p>Monthly distribution (%) of overbank (H ≥ 250 cm) flood waves (<b>A</b>) and low stages (H ≤ 0 cm) measured at Szentgotthárd gauging station (<b>B</b>).</p>
Full article ">Figure 5
<p>Mean duration of low stages (≤0 cm) and overbank stages (≥250 cm) at Szentgotthárd.</p>
Full article ">Figure 6
<p>Daily water level (cm) and discharge (m<sup>3</sup>/s) changes in the Rába River measured at Szentgotthárd gauging station during the field survey.</p>
Full article ">Figure 7
<p>Bank erosion rates of single bends of the studied reach of the Rába River (<b>A</b>) and the characteristic annual stages during the same time (<b>B</b>).</p>
Full article ">Figure 8
<p>Mean bank erosion (<b>A</b>) and maximum bank erosion (<b>B</b>) of some meanders between the RTK–GPS surveys performed in every four months between April 2022 and April 2024.</p>
Full article ">Figure 9
<p>Short- (2022–2024) and long-term (1951–2022) changes in the position of the bankline at selected meanders. Meander B represents a bend where the location of an island influences the bank erosion. Meander C shows a classical meander expansion. North of Meander E is a terrace rim; thus, the meander development is confined. On Meander H, secondary bends develop.</p>
Full article ">Figure 10
<p>Conceptual model of declining bank erosion driven by hydrological changes.</p>
Full article ">
28 pages, 3401 KiB  
Review
A Critical Review of Emerging Technologies for Flash Flood Prediction: Examining Artificial Intelligence, Machine Learning, Internet of Things, Cloud Computing, and Robotics Techniques
by Ghazi Al-Rawas, Mohammad Reza Nikoo, Malik Al-Wardy and Talal Etri
Water 2024, 16(14), 2069; https://doi.org/10.3390/w16142069 - 22 Jul 2024
Viewed by 1070
Abstract
There has been growing interest in the application of smart technologies for hazard management. However, very limited studies have reviewed the trends of such technologies in the context of flash floods. This study reviews innovative technologies such as artificial intelligence (AI)/machine learning (ML), [...] Read more.
There has been growing interest in the application of smart technologies for hazard management. However, very limited studies have reviewed the trends of such technologies in the context of flash floods. This study reviews innovative technologies such as artificial intelligence (AI)/machine learning (ML), the Internet of Things (IoT), cloud computing, and robotics used for flash flood early warnings and susceptibility predictions. Articles published between 2010 and 2023 were manually collected from scientific databases such as Google Scholar, Scopus, and Web of Science. Based on the review, AI/ML has been applied to flash flood susceptibility and early warning prediction in 64% of the published papers, followed by the IoT (19%), cloud computing (6%), and robotics (2%). Among the most common AI/ML methods used in susceptibility and early warning predictions are random forests and support vector machines. However, further optimization and emerging technologies, such as computer vision, are required to improve these technologies. AI/ML algorithms have demonstrated very accurate prediction performance, with receiver operating characteristics (ROC) and areas under the curve (AUC) greater than 0.90. However, there is a need to improve on these current models with large test datasets. Through AI/ML, IoT, and cloud computing technologies, early warnings can be disseminated to targeted communities in real time via electronic media, such as SMS and social media platforms. In spite of this, these systems have issues with internet connectivity, as well as data loss. Additionally, Al/ML used a number of topographical variables (such as slope), geological variables (such as lithology), and hydrological variables (such as stream density) to predict susceptibility, but the selection of these variables lacks a clear theoretical basis and has inconsistencies. To generate more reliable flood risk assessment maps, future studies should also consider sociodemographic, health, and housing data. Considering future climate change impacts, susceptibility or early warning studies may be projected under different climate change scenarios to help design long-term adaptation strategies. Full article
(This article belongs to the Section Hydrology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of articles (50) published in 2010–2023 according to (<b>a</b>) the number of papers and by year of publication and (<b>b</b>) categories of technologies such as artificial intelligence/machine learning (AI/ML), the Internet of Things (IoT), cloud computing, robotics, and other technologies (e.g., storm cell identification, video-based surveillance, interactive voice response, digital image analysis).</p>
Full article ">Figure 2
<p>Geographical distribution (<b>a</b>) and frequencies (<b>b</b>) of published papers (2010–2023) on flash flood susceptibility and early warnings using AI/ML, the IoT, cloud computing, robotics, and other technologies (e.g., storm cell identification, video-based surveillance, interactive voice response, digital image analysis).</p>
Full article ">Figure 3
<p>The number of different AI/ML algorithms (N = 51) used for flash flood susceptibility and early warning predictions (2010–2023).</p>
Full article ">Figure 4
<p>The number of AI/ML algorithms per year (2010–2023) applied in flash flood susceptibility and warning predictions.</p>
Full article ">Figure 5
<p>Study design, main findings, limitations, and future perspectives.</p>
Full article ">
23 pages, 11525 KiB  
Article
Agent-Based Modeling for Household Decision-Making in Adoption of Private Flood Mitigation Measures: The Upper Kan Catchment Case Study
by Shima Nabinejad and Holger Schüttrumpf
Water 2024, 16(14), 2027; https://doi.org/10.3390/w16142027 - 17 Jul 2024
Viewed by 648
Abstract
Residential areas in developing arid and semi-arid countries are highly vulnerable to flooding, and water shortages have forced their inhabitants to live close to rivers. While climate change is expected to cause more extreme weather conditions in the future, adopting private loss-reduction measures [...] Read more.
Residential areas in developing arid and semi-arid countries are highly vulnerable to flooding, and water shortages have forced their inhabitants to live close to rivers. While climate change is expected to cause more extreme weather conditions in the future, adopting private loss-reduction measures can diminish flood risk. Although the number of flood models has grown significantly for developing arid and semi-arid lands, these models suffer from being incapable of performing micro-scale flood risk analysis and including household behaviors. This research work presents a novel socio-economic simulation model in the framework of flood risk management (FRM) to couple household adaptive responses with flood risk analysis. Agent-based modeling (ABM) embeds human behaviors in a flood-simulating environment. The focus of this research is the upper Kan catchment in Iran with a long history of severe flash flooding. Our results show the ability of the developed framework to address household participation in FRM activities through private precautionary measures. Moreover, the results indicate the importance of presenting such micro-level behaviors in flood modeling for a more realistic flood risk assessment. It is also demonstrated that household adaptation in a continuous way can lead to less flood risks in the region, even under climate change and the future economy. Finally, the results reveal that the remaining and diminished regional flood risks are influenced by the behavioral framework through which the individuals make decisions in adopting flood-loss-reduction measures. A similar pattern is, however, observed in household contributions over time among the discussed behavioral approaches. Full article
Show Figures

Figure 1

Figure 1
<p>Conceptual framework of ABMhofo for flood adaptation behaviors of local households.</p>
Full article ">Figure 2
<p>Components and input data of the flood sub-module in ABMhofo.</p>
Full article ">Figure 3
<p>An example probability–damage curve generated for three flood events with return periods <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">T</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>,<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">T</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">T</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Activity diagram of the household behaviors and decision-making in the developed agent-based model.</p>
Full article ">Figure 5
<p>Map of Iran (<b>a</b>) and the study area (<b>b</b>).</p>
Full article ">Figure 6
<p>Land use map of Kan upper catchment.</p>
Full article ">Figure 7
<p>Spatially distributed water level and flood depths in upper Kan catchment under current climate for the: (<b>a</b>) 50-year, (<b>b</b>) 100-year, and (<b>c</b>) 500-year flood events.</p>
Full article ">Figure 8
<p>Generation of the house shapefiles based on Google satellite images.</p>
Full article ">Figure 9
<p>Depth–effectiveness curves of the private flood-loss-reduction measures.</p>
Full article ">Figure 10
<p>Inflation rate in Germany from 2005–2022.</p>
Full article ">Figure 11
<p>Contribution of rational household agents in adopting flood risk mitigation measures over time under three frameworks: expected utility (<b>a</b>), cost–benefit analysis (<b>b</b>), and individual flood risk (<b>c</b>).</p>
Full article ">Figure 12
<p>The impact of flood-loss-reduction measures in the last simulation year on the house’ damage ratios under the three rational perspectives: bars = means; points = medians; lines = upper and lower quartile.</p>
Full article ">Figure 13
<p>Residual and reduced flood risk of the whole area over time under three rational behavioral frameworks: expected utility (<b>a</b>), cost–benefit analysis (<b>b</b>), and individual flood risk (<b>c</b>).</p>
Full article ">Figure 14
<p>Change in aggregated residual flood risk for three behavioral frameworks expected utility (<b>a</b>), cost–benefit analysis (<b>b</b>), and individual flood risk (<b>c</b>).</p>
Full article ">Figure 15
<p>Total reduced flood risk of the population as well as the number of measures in place for simulation year 5 and year 8.</p>
Full article ">Figure 16
<p>Stacked bar graph of computed flood risk at the household-level in the last simulation year (40th time step) considering climate change when the household agents take no loss-reduction measures (blue bar chart) and when they take loss-reduction measures (orange bar chart).</p>
Full article ">
15 pages, 2273 KiB  
Article
Experimental Investigation on the Effect of Sequences of Unsteady Flows on Bedload Sediment Transport
by Zahra Askari, Luca Mao, Saeed Reza Khodashenas and Kazem Esmaili
Geosciences 2024, 14(7), 193; https://doi.org/10.3390/geosciences14070193 - 17 Jul 2024
Viewed by 517
Abstract
Flash floods in ephemeral streams are rare, short and difficult to forecast and thus to monitor. During these events, bedload transport reaches very high rates and most sediment transport occurs within a limited number of hours during the course of a year. Because [...] Read more.
Flash floods in ephemeral streams are rare, short and difficult to forecast and thus to monitor. During these events, bedload transport reaches very high rates and most sediment transport occurs within a limited number of hours during the course of a year. Because monitoring of bedload in ephemeral rivers is challenging, here we present the results of a series of flume experiments designed to simulate short, flashy floods. Since most flume experiments usually involve single events, here we add to existing evidence by testing the effects of sequences of multiple floods in rapid succession. The flume is 10 m long, 0.3 m wide and 0.5 m deep. Two bed sediment mixtures (well sorted and poorly sorted) with similar median grain size but a different standard deviation were used. Bedload was monitored continuously during each hydrograph, but no sediment was fed. The flume experiments used six triangular hydrographs with peak flows ranging from 0.0147 to 0.02 m3s1 and durations ranging from 150 to 400 s. Results indicate that the sediment transport rate decreases progressively from the first to the third hydrograph, and that this pattern is consistent for all permutations of peak discharge and flood duration. In all of the runs, the sediment transport rate at a specified flow was higher during the rising limb than the falling limb of the hydrograph, indicating clockwise hysteresis. Furthermore, in the subsequent repetitions of the same hydrograph, the degree of hysteresis generally diminishes in magnitude from the first to the last repetition for all the experiments, irrespective of their magnitude and duration. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic figure of the flume showing the location of the cameras and the verticals where the water depth was measured.</p>
Full article ">Figure 2
<p>Symmetrical hydrographs with different duration and magnitude simulated on the experiment flume. The hydrographs A, B and C have the same duration but different peak discharge and the hydrographs D, E and F have the same peak discharge but different duration.</p>
Full article ">Figure 3
<p>Graphs showing the repetitions of the hydrographs and the bedload transport rated (in g s<sup>−1</sup>) measured during the runs conducted with well-sorted and poorly sorted sediments in the flume. The graphs refer to the runs conducted repeating three times the hydrographs, and the label of the graph refers to the type of hydrograph (as depicted on <a href="#geosciences-14-00193-f002" class="html-fig">Figure 2</a>: (<b>A</b>–<b>F</b>). The data referring to the first repetition of the hydrograph B could not be retrieved and are unfortunately missing.</p>
Full article ">Figure 4
<p>Plots of sediment transport rate vs. liquid discharge for the experiments run with poorly sorted sediments, showing the temporal hysteresis from the first to the third repetition of each type of hydrographs. Graphs on the right (poorly sorted) and graphs on the left (well sorted). Note: <math display="inline"><semantics> <mrow> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> <mo> </mo> <mo>,</mo> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </mfenced> <mo>,</mo> <mo> </mo> <mo>(</mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math>, (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> <mo>,</mo> <mo>(</mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> <mo>,</mo> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> relate to hydrographs A, B, C, D, E and F, for well-sorted and poorly sorted, respectively. The subscripts w and p stand for well- and poorly sorted sediment mixtures, respectively, and the letters R and F refer to rising and falling limb of hydrographs, respectively.</p>
Full article ">Figure 4 Cont.
<p>Plots of sediment transport rate vs. liquid discharge for the experiments run with poorly sorted sediments, showing the temporal hysteresis from the first to the third repetition of each type of hydrographs. Graphs on the right (poorly sorted) and graphs on the left (well sorted). Note: <math display="inline"><semantics> <mrow> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> <mo> </mo> <mo>,</mo> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </mfenced> <mo>,</mo> <mo> </mo> <mo>(</mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math>, (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> <mo>,</mo> <mo>(</mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> <mo>,</mo> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> relate to hydrographs A, B, C, D, E and F, for well-sorted and poorly sorted, respectively. The subscripts w and p stand for well- and poorly sorted sediment mixtures, respectively, and the letters R and F refer to rising and falling limb of hydrographs, respectively.</p>
Full article ">Figure 4 Cont.
<p>Plots of sediment transport rate vs. liquid discharge for the experiments run with poorly sorted sediments, showing the temporal hysteresis from the first to the third repetition of each type of hydrographs. Graphs on the right (poorly sorted) and graphs on the left (well sorted). Note: <math display="inline"><semantics> <mrow> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> <mo> </mo> <mo>,</mo> <msub> <mrow> <mi>A</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </mfenced> <mo>,</mo> <mo> </mo> <mo>(</mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math>, (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> <mo>,</mo> <mo>(</mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo>)</mo> <mo>,</mo> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>w</mi> <mo> </mo> </mrow> </msub> <mo>,</mo> <mo> </mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> relate to hydrographs A, B, C, D, E and F, for well-sorted and poorly sorted, respectively. The subscripts w and p stand for well- and poorly sorted sediment mixtures, respectively, and the letters R and F refer to rising and falling limb of hydrographs, respectively.</p>
Full article ">Figure 5
<p>Comparison of the surface sediment size taken on three sections (upstream, middle and downstream end of the flume) before and after one of the runs.</p>
Full article ">Figure 6
<p>The comparison of the gradation size sediment transported in rising and falling hydrograph and the initial bed surfaces. Note: R1, R2 and R3 show the first repetition, the second repetition and the third repetition, respectively.</p>
Full article ">Figure 6 Cont.
<p>The comparison of the gradation size sediment transported in rising and falling hydrograph and the initial bed surfaces. Note: R1, R2 and R3 show the first repetition, the second repetition and the third repetition, respectively.</p>
Full article ">
26 pages, 10009 KiB  
Article
A Nationwide Flood Forecasting System for Saudi Arabia: Insights from the Jeddah 2022 Event
by Giulia Sofia, Qing Yang, Xinyi Shen, Mahjabeen Fatema Mitu, Platon Patlakas, Ioannis Chaniotis, Andreas Kallos, Mohammed A. Alomary, Saad S. Alzahrani, Zaphiris Christidis and Emmanouil Anagnostou
Water 2024, 16(14), 1939; https://doi.org/10.3390/w16141939 - 9 Jul 2024
Viewed by 1004
Abstract
Saudi Arabia is threatened by recurrent flash floods caused by extreme precipitation events. To mitigate the risks associated with these natural disasters, we implemented an advanced nationwide flash flood forecast system, boosting disaster preparedness and response. A noteworthy feature of this system is [...] Read more.
Saudi Arabia is threatened by recurrent flash floods caused by extreme precipitation events. To mitigate the risks associated with these natural disasters, we implemented an advanced nationwide flash flood forecast system, boosting disaster preparedness and response. A noteworthy feature of this system is its national-scale operational approach, providing comprehensive coverage across the entire country. Using cutting-edge technology, the setup incorporates a state-of-the-art, three-component system that couples an atmospheric model with hydrological and hydrodynamic models to enable the prediction of precipitation patterns and their potential impacts on local communities. This paper showcases the system’s effectiveness during an extreme precipitation event that struck Jeddah on 24 November 2022. The event, recorded as the heaviest rainfall in the region’s history, led to widespread flash floods, highlighting the critical need for accurate and timely forecasting. The flash flood forecast system proved to be an effective tool, enabling authorities to issue warnings well before the flooding, allowing residents to take precautionary measures, and allowing emergency responders to mobilize resources effectively. Full article
Show Figures

Figure 1

Figure 1
<p>Jeddah flood, images from videos available online.</p>
Full article ">Figure 2
<p>Overview of the NCM Flash Flood Forecasting System. The system combines the WRF weather model, the CREST hydrological model, and the HEC-RAS 2D hydraulic model. WRF hourly weather forecasts force CREST to generate hourly discharge time series four times daily. Finally, the hydrographs derived from the hydrological model serve as input hydrographs in the 2D HEC-RAS hydrodynamic model for flood inundation modeling and mapping.</p>
Full article ">Figure 3
<p>WRF model domain: black boxes highlight the coarse nest, including parts of Africa, Asia, and Europe, with a spatial resolution of 4.8 km, and the fine nest, covering the Arabian Peninsula with a spatial resolution of 1.6 km.</p>
Full article ">Figure 4
<p>CREST-HEC2D settings for the whole system, with a zoom to the midwest coast (red square in the map at the top and highlighted in panel (<b>a</b>) below that). Panel (<b>a</b>) compares the CREST watersheds and their outlets to the HEC2D domain for the 24 November 2022 event. The midwest coast domain encompasses the cities of Jeddah, Makkah, and Taif. The bottom images display a detailed view of an area in Jeddah (panel (<b>b</b>)), with the corresponding terrain at 30 m (panel (<b>c</b>)) and the high-resolution terrain at 2.5 m (panel (<b>d</b>)). The quality and resolution of the terrain allow the depiction of the footprints of buildings and roads, as highlighted by the roughness of the map in panel (<b>c</b>). High resolution (panel (<b>d</b>)) provides detailed information for improved flood modeling.</p>
Full article ">Figure 5
<p>Locations of the stations used for the evaluation.</p>
Full article ">Figure 6
<p>Simulated flood hydrograph derived from the tuned parameters compared with the observed record for 1985 at the outlet of Wadi_415.</p>
Full article ">Figure 7
<p><b>The</b> 24-h accumulated precipitation for 24 November 2022, as retrieved by (<b>a</b>) WRF, (<b>b</b>) IMERG-Late, and (<b>c</b>) bias-adjusted radar.</p>
Full article ">Figure 8
<p>Cumulative rainfall curves of WRF, IMERG-Late, bias-adjusted radar and station observations.</p>
Full article ">Figure 9
<p>Object analysis, WRF objects (<b>a</b>), IMERG-Late objects (<b>b</b>), IMERG outlines, and WRF objects (colored). Comparison of the overlapping objects is shown in (<b>c</b>). Object 1 in blue highlights the area of interest of Jeddah, whereas object 2 in red highlights a secondary area of rainfall.</p>
Full article ">Figure 10
<p>Comparison between the WRF-based forecast and radar-based simulation. The figure reports the flooded locations (<b>a</b>) and time series of depths and CREST discharge for two selected locations (<b>b</b>).</p>
Full article ">Figure 11
<p>Inundation depth over Jeddah, as simulated from discharge derived using WRF and radar forcings. The figure shows (<b>a</b>) the overall variability from the boxplots, as well as the density distribution of points across the domain, and the correspondence between WRF and radar data of inundation depth (<b>b</b>) and the percentiles of inundation depth (<b>c</b>).</p>
Full article ">Figure 12
<p>WRF-based forecasted depth, compared to selected crowdsourced videos. Letters A to E reports the location of each crowdsourced image on the right-hand side of the figure.</p>
Full article ">Figure 13
<p>CREST outlets. Quantiles of discharge from WRF forecast (<b>a</b>) compared to those obtained from radar data (<b>b</b>). One-to-one comparison for selected outlets is also shown (<b>c</b>).</p>
Full article ">Figure 14
<p>Inundation quantiles at selected locations. Neighboring zones (<b>a</b>) and quantiles of flood depth (<b>b</b>) compared to the historical distribution of depths for the same areas (<b>c</b>). The red line in (<b>c</b>) indicates the most frequent depth within those areas for the November 2022 event.</p>
Full article ">Figure 15
<p>Warning system available to the authorities: overall warning (<b>a</b>), maximum warning for the event (<b>b</b>), and warning time series at different moments of the available simulation window (<b>c</b>–<b>e</b>). The available information for each hexagon is shown in (<b>f</b>).</p>
Full article ">Figure 16
<p>Detailed comparison of the WRF-based flood inundation using the 30 m DEM (<b>a</b>) and the 2.5 m high-resolution one (<b>b</b>). The left panel image shows a satellite view taken over Jeddah on November 25 at 8.30 UTC by PLÉIADES NEO SATELLITE. While it is not possible to detect the actual flood extent, the image shows sediment being delivered to the sea, highlighted by the red arrow and circle, with input from the locations where the flood forecast system showed the highest floodwater accumulation.</p>
Full article ">Figure 17
<p>WRF-based flood warnings from 30 m (<b>a</b>,<b>d</b>) and 2.5 m DEM (<b>b</b>,<b>e</b>), as compared to flooded locations crowdsourced on November 24–25 (<b>c</b>) and traffic on November 24 (<b>f</b>). Crowdsourced flooding for 2011 and 2017 is also reported. (<b>c</b>,<b>f</b>) are from Flickr; credits to Efren Rodriguez (2011) and Saidalavi Mohamed (2017).</p>
Full article ">
23 pages, 15916 KiB  
Article
Park Heritage of the Island of Krk between Urban Transformations and Climate Change
by Koraljka Vahtar-Jurković, Renata Sokol Jurković and Jadran Jurković
Land 2024, 13(7), 1024; https://doi.org/10.3390/land13071024 - 8 Jul 2024
Viewed by 678
Abstract
The island of Krk in Primorje-Gorski Kotar County, Croatia, is also called the Golden Island because of its favorable geographical location, the diversity of natural and urban landscapes, the beauty of the coast and the sea, the wealth of tangible and intangible heritage, [...] Read more.
The island of Krk in Primorje-Gorski Kotar County, Croatia, is also called the Golden Island because of its favorable geographical location, the diversity of natural and urban landscapes, the beauty of the coast and the sea, the wealth of tangible and intangible heritage, and especially because of the opportunities for living and working. During the last century and in this century, urban landscapes have been exposed to dramatic changes that transformed old castles or former smaller settlements of the local population into tourist centers and desirable places for permanent or temporary residence. A significant part of their complex structure is the cultural and historical heritage, within which the island’s park heritage has so far been insufficiently recognized and valued. Therefore, this paper examines forty selected public park spaces in the area of all local self-government units of the island of Krk in the context of urban transformations and climate change. It is concluded that the island’s park heritage has often been created as a result of urban transformations in which, despite being exposed to constant changes, it is mostly preserved, but that recently, new elements of this heritage are emerging. A new challenge is the threat of the consequences of climate change—increasingly frequent and long-lasting droughts, extreme precipitation and flash floods, stormy winds, rising sea level and salinization, which further emphasizes the need to preserve the park heritage of the island of Krk in the context of resistance to climate change. Full article
(This article belongs to the Special Issue Urban Landscape Transformation vs. Heritage)
Show Figures

Figure 1

Figure 1
<p>The island of Krk and its location in Primorje-Gorski Kotar County, Croatia (map created in ArcGIS 10.6).</p>
Full article ">Figure 2
<p>Location of the public town parks, forest parks, hotel/camping parks, and promenades in settlements of the island of Krk (maps created in ArcGIS 10.6).</p>
Full article ">Figure 3
<p>Public parks, forest parks, hotel/camping parks, and promenades on the island of Krk (map created in ArcGIS 10.6).</p>
Full article ">Figure 4
<p>Some public town parks, forest parks, parks of hotels, and promenades of the island of Krk. (photos by K. Vahtar-Jurković, except Roof park DUBoak, Copyright 2024, Copyright Maritime Heritage Interpretation Center DUBoak).</p>
Full article ">Figure 4 Cont.
<p>Some public town parks, forest parks, parks of hotels, and promenades of the island of Krk. (photos by K. Vahtar-Jurković, except Roof park DUBoak, Copyright 2024, Copyright Maritime Heritage Interpretation Center DUBoak).</p>
Full article ">Figure 5
<p>Threat map of the coastal area of the island of Krk due to coastal flooding (Adapted with permission from Ref. [<a href="#B68-land-13-01024" class="html-bibr">68</a>]. Copyright 2024, copyright Assoc.Prof. Igor Ružić, PhD).</p>
Full article ">Figure 6
<p>Threats on the parks/promenades of the island of Krk from the consequences of climate change.</p>
Full article ">
27 pages, 3990 KiB  
Review
Navigating the Uncertain Terrain: Venezuela’s Future Using the Shared Socioeconomic Pathways Framework—A Systematic Review
by Isaias Lescher Soto, Alicia Villamizar, Barlin O. Olivares, María Eugenia Gutiérrez and Gustavo J. Nagy
Climate 2024, 12(7), 98; https://doi.org/10.3390/cli12070098 - 6 Jul 2024
Viewed by 1975
Abstract
We investigate Venezuela’s potential “futures” under Shared Socioeconomic Pathways (SSPs) through a systematic literature review, including systematic mapping and thematic analysis of 50 scientific articles. We categorised the SSP scenarios into two generational categories and classified the outcomes into positive, negative, and neutral [...] Read more.
We investigate Venezuela’s potential “futures” under Shared Socioeconomic Pathways (SSPs) through a systematic literature review, including systematic mapping and thematic analysis of 50 scientific articles. We categorised the SSP scenarios into two generational categories and classified the outcomes into positive, negative, and neutral futures. Under first-generation SSP scenarios, increasing poverty could be reversed, and the country’s economic growth could be stimulated by adopting unambitious climate measures. However, second-generation SSP scenarios paint a more challenging picture. They suggest that Venezuela could face heat waves, droughts, an increase in diseases, loss of biodiversity, and an increase in invasive species and pests during the remainder of the 21st century as a direct consequence of climate change. Venezuela’s geographic and topographic diversity could exacerbate these impacts of climate change. For instance, coastal areas could be at risk of sea-level rise and increased storm surges, while mountainous regions could experience more frequent and intense rainfall, leading to landslides and flash floods. The urgency of conducting additional research on the factors that could influence the severity of climate change’s impact, considering Venezuela’s geographic and topographic diversity, cannot be overstated. We also identified the critical need to explore alternative paths to move away from the current extractive development model. The potential actions in this regard could be instrumental in aligning the country with global adaptation and mitigation commitments. Full article
Show Figures

Figure 1

Figure 1
<p>PRISMA flow diagram according to Page et al. [<a href="#B58-climate-12-00098" class="html-bibr">58</a>]. ** If automation tools were used, indicate how many records were excluded by a human and how many were excluded by automation tools.</p>
Full article ">Figure 2
<p>Geographic distribution of the first author in the scientific publications of the review and quantity of papers according to location (<span class="html-italic">n</span> = 50).</p>
Full article ">Figure 3
<p>(<b>a</b>) Circle Packing of total journals per database and the number of papers per journal. Coloured circles represent each scientific journal included in the review. The circles are nested based on the consulted database. Pink circles reveal the number of articles in IOPSCIENCE, orange in JSTOR, yellow in Publimed, green in ScienceDirect, and purple in Scopus. The size of each circle indicates the number of articles per journal. (<b>b</b>) Beeswarm plot of total papers for SSP generations. Years: 2013–2023 (<span class="html-italic">n</span> = 50). Lines and colours represent the generations of SSP scenarios. Light blue corresponds to the first generation, purple to the second generation, and pink to both generations. The lines group the publications for each scenario generation. Circles indicate the journals where the consulted publications were produced and the publication year. The circled area illustrates the extent of SSP studies over the years.</p>
Full article ">Figure 4
<p>Circular dendrogram of the areas of knowledge scientific articles (<span class="html-italic">n</span> = 50) and their citations, Zhang et al. [<a href="#B4-climate-12-00098" class="html-bibr">4</a>], Ávila Díaz et al. [<a href="#B6-climate-12-00098" class="html-bibr">6</a>]; Fan et al. [<a href="#B63-climate-12-00098" class="html-bibr">63</a>]; Viloria et al. [<a href="#B64-climate-12-00098" class="html-bibr">64</a>]; Wang et al. [<a href="#B65-climate-12-00098" class="html-bibr">65</a>]; Egli and Stünzi [<a href="#B66-climate-12-00098" class="html-bibr">66</a>]; Crespo Cuaresma et al. [<a href="#B67-climate-12-00098" class="html-bibr">67</a>]; Campagnolo and Davide [<a href="#B68-climate-12-00098" class="html-bibr">68</a>]; Ostadzadeh et al. [<a href="#B69-climate-12-00098" class="html-bibr">69</a>]; Adhikari et al. [<a href="#B70-climate-12-00098" class="html-bibr">70</a>]; Andrews et al. [<a href="#B71-climate-12-00098" class="html-bibr">71</a>]; Benveniste et al. [<a href="#B72-climate-12-00098" class="html-bibr">72</a>]; Carlson et al. [<a href="#B73-climate-12-00098" class="html-bibr">73</a>]; Chatting et al. [<a href="#B74-climate-12-00098" class="html-bibr">74</a>]; Chen et al. [<a href="#B75-climate-12-00098" class="html-bibr">75</a>]; Colón-González et al. [<a href="#B76-climate-12-00098" class="html-bibr">76</a>]; Colón-González et al. [<a href="#B77-climate-12-00098" class="html-bibr">77</a>]; Cooper et al. [<a href="#B78-climate-12-00098" class="html-bibr">78</a>]; Dosio et al. [<a href="#B79-climate-12-00098" class="html-bibr">79</a>]; Dutta et al. [<a href="#B80-climate-12-00098" class="html-bibr">80</a>]; Fernández-Alvarez et al. [<a href="#B81-climate-12-00098" class="html-bibr">81</a>]; Ganglo [<a href="#B82-climate-12-00098" class="html-bibr">82</a>]; Guirado et al. [<a href="#B83-climate-12-00098" class="html-bibr">83</a>]; Hanasaki et al. [<a href="#B84-climate-12-00098" class="html-bibr">84</a>]; Hernández et al. [<a href="#B85-climate-12-00098" class="html-bibr">85</a>]; Herrera-Feijoo et al. [<a href="#B86-climate-12-00098" class="html-bibr">86</a>]; Jin et al. [<a href="#B87-climate-12-00098" class="html-bibr">87</a>]; Johnston &amp;and Radeloff [<a href="#B88-climate-12-00098" class="html-bibr">88</a>]; Kemp et al. [<a href="#B89-climate-12-00098" class="html-bibr">89</a>]; Kinoshita et al. [<a href="#B90-climate-12-00098" class="html-bibr">90</a>]; Laporta et al. [<a href="#B91-climate-12-00098" class="html-bibr">91</a>]; Lenton et al. [<a href="#B92-climate-12-00098" class="html-bibr">92</a>]; Lopes et al. [<a href="#B93-climate-12-00098" class="html-bibr">93</a>]; Moo-Llanes et al. [<a href="#B94-climate-12-00098" class="html-bibr">94</a>]; Nkiriki et al. [<a href="#B95-climate-12-00098" class="html-bibr">95</a>]; Pérez et al. [<a href="#B96-climate-12-00098" class="html-bibr">96</a>]; Petrova et al. [<a href="#B97-climate-12-00098" class="html-bibr">97</a>]; Pretis et al. [<a href="#B98-climate-12-00098" class="html-bibr">98</a>]; Purse et al. [<a href="#B99-climate-12-00098" class="html-bibr">99</a>]; Rodríguez De Luque et al. [<a href="#B100-climate-12-00098" class="html-bibr">100</a>]; Sampedro et al. [<a href="#B101-climate-12-00098" class="html-bibr">101</a>]; Setter et al. [<a href="#B102-climate-12-00098" class="html-bibr">102</a>]; Shen et al. [<a href="#B103-climate-12-00098" class="html-bibr">103</a>]; Shepherd et al. [<a href="#B104-climate-12-00098" class="html-bibr">104</a>]; Tong et al. [<a href="#B105-climate-12-00098" class="html-bibr">105</a>]; Wainwright et al. [<a href="#B106-climate-12-00098" class="html-bibr">106</a>]; Wang et al. [<a href="#B107-climate-12-00098" class="html-bibr">107</a>]; Welsby et al. [<a href="#B108-climate-12-00098" class="html-bibr">108</a>]; Wiebe et al. [<a href="#B109-climate-12-00098" class="html-bibr">109</a>]; Zampieri et al. [<a href="#B110-climate-12-00098" class="html-bibr">110</a>]. Coloured nodes and lines represent the relationships between the Knowledge Field, Narrow Field [<a href="#B57-climate-12-00098" class="html-bibr">57</a>], and the subject areas identified in the consulted articles. Field: Manufacturing and Construction); NSMS (Natural Sciences, Mathematics and Statistics) and SSJI (Social Sciences, Journalism, and Information). Narrow Field: (A) (Agriculture), BRS (Biological and Related Sciences), EET (Engineering and Engineering Trades) E Environment, PS (Physical Sciences) SBS (Social and Behavioural Sciences). Subject areas: CS (Carbon Sequestration), CCI (Climate Change Impacts), CF (Climate Finance), C and EG (Conflict, Economic Growth), (C) (Crops), DL (Demand for Land-based Transportation Services) Eco (Ecosystems), GET (Global Energy Trade), GH Global hunger, (M) Migration (ME) Methane Emissions, P (Poverty), SG (Solar Geoengineering) (SE) Solar Energy, (URFF) (Unextractable Reserves of Fossil Fuels), V (Vulnerability) and WS Water Scarcity Further individual article details are available in the <a href="#app1-climate-12-00098" class="html-app">Supplementary Materials</a>. Box indicates: Green: AFFV; Orange: EMC; Blue: NSMS and Red SSJI. See <a href="#app1-climate-12-00098" class="html-app">Supplementary Materials</a> for references [<a href="#B4-climate-12-00098" class="html-bibr">4</a>,<a href="#B6-climate-12-00098" class="html-bibr">6</a>,<a href="#B63-climate-12-00098" class="html-bibr">63</a>,<a href="#B64-climate-12-00098" class="html-bibr">64</a>,<a href="#B65-climate-12-00098" class="html-bibr">65</a>,<a href="#B66-climate-12-00098" class="html-bibr">66</a>,<a href="#B67-climate-12-00098" class="html-bibr">67</a>,<a href="#B68-climate-12-00098" class="html-bibr">68</a>,<a href="#B69-climate-12-00098" class="html-bibr">69</a>,<a href="#B70-climate-12-00098" class="html-bibr">70</a>,<a href="#B71-climate-12-00098" class="html-bibr">71</a>,<a href="#B72-climate-12-00098" class="html-bibr">72</a>,<a href="#B73-climate-12-00098" class="html-bibr">73</a>,<a href="#B74-climate-12-00098" class="html-bibr">74</a>,<a href="#B75-climate-12-00098" class="html-bibr">75</a>,<a href="#B76-climate-12-00098" class="html-bibr">76</a>,<a href="#B77-climate-12-00098" class="html-bibr">77</a>,<a href="#B78-climate-12-00098" class="html-bibr">78</a>,<a href="#B79-climate-12-00098" class="html-bibr">79</a>,<a href="#B80-climate-12-00098" class="html-bibr">80</a>,<a href="#B81-climate-12-00098" class="html-bibr">81</a>,<a href="#B82-climate-12-00098" class="html-bibr">82</a>,<a href="#B83-climate-12-00098" class="html-bibr">83</a>,<a href="#B84-climate-12-00098" class="html-bibr">84</a>,<a href="#B85-climate-12-00098" class="html-bibr">85</a>,<a href="#B86-climate-12-00098" class="html-bibr">86</a>,<a href="#B87-climate-12-00098" class="html-bibr">87</a>,<a href="#B88-climate-12-00098" class="html-bibr">88</a>,<a href="#B89-climate-12-00098" class="html-bibr">89</a>,<a href="#B90-climate-12-00098" class="html-bibr">90</a>,<a href="#B91-climate-12-00098" class="html-bibr">91</a>,<a href="#B92-climate-12-00098" class="html-bibr">92</a>,<a href="#B93-climate-12-00098" class="html-bibr">93</a>,<a href="#B94-climate-12-00098" class="html-bibr">94</a>,<a href="#B95-climate-12-00098" class="html-bibr">95</a>,<a href="#B96-climate-12-00098" class="html-bibr">96</a>,<a href="#B97-climate-12-00098" class="html-bibr">97</a>,<a href="#B98-climate-12-00098" class="html-bibr">98</a>,<a href="#B99-climate-12-00098" class="html-bibr">99</a>,<a href="#B100-climate-12-00098" class="html-bibr">100</a>,<a href="#B101-climate-12-00098" class="html-bibr">101</a>,<a href="#B102-climate-12-00098" class="html-bibr">102</a>,<a href="#B103-climate-12-00098" class="html-bibr">103</a>,<a href="#B104-climate-12-00098" class="html-bibr">104</a>,<a href="#B105-climate-12-00098" class="html-bibr">105</a>,<a href="#B106-climate-12-00098" class="html-bibr">106</a>,<a href="#B107-climate-12-00098" class="html-bibr">107</a>,<a href="#B108-climate-12-00098" class="html-bibr">108</a>,<a href="#B109-climate-12-00098" class="html-bibr">109</a>,<a href="#B110-climate-12-00098" class="html-bibr">110</a>].</p>
Full article ">Figure 5
<p>Alluvial diagram of the correlations between categorical dimensions Scenario Generation, Representative Concentration Pathways (RCP), Thematic Focus, and ISCED-F 2013 Detailed Field (areas of knowledge) (<span class="html-italic">n</span> = 50).</p>
Full article ">Figure 6
<p>Possible futures for Venezuela under SSP Generation scenarios.</p>
Full article ">
Back to TopTop