[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Issue
Volume 6, March
Previous Issue
Volume 5, September
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

GeoHazards, Volume 5, Issue 4 (December 2024) – 17 articles

Cover Story (view full-size image): Earthquakes pose a significant risk globally, yet accurately predicting their occurrence remains a scientific challenge. This paper investigates multiple state-of-the-art temporal, spatial, and spatiotemporal AI models for earthquake nowcasting, integrating advanced deep learning architectures and foundation models. Additionally, the study introduces two innovative models, Multi Foundation Quake and GNNCoder, designed to capture intricate spatial and temporal patterns in seismic activity in Southern California. The models use seismic records to integrate pre-trained foundation architectures with bespoke patterns and graph-based learning techniques. The findings highlight our new models’ superior performance over existing methods. This research provides a transformative step in utilizing AI for seismic prediction models, showcasing a multidisciplinary understanding of earthquake dynamics. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Select all
Export citation of selected articles as:
19 pages, 26960 KiB  
Article
The Northern Giona Fault Zone, a Major Active Structure Through Central Greece
by Leonidas Gouliotis and Dimitrios Papanikolaou
GeoHazards 2024, 5(4), 1370-1388; https://doi.org/10.3390/geohazards5040065 - 18 Dec 2024
Viewed by 587
Abstract
The steep northern slopes of Giona Mt in central continental Greece are the result of an E-W normal fault dipping 35–45° to the north, extending from the Mornos River in the west to the village of Gravia in the east. This fault creates [...] Read more.
The steep northern slopes of Giona Mt in central continental Greece are the result of an E-W normal fault dipping 35–45° to the north, extending from the Mornos River in the west to the village of Gravia in the east. This fault creates a significant elevation difference of approximately 1500 m between the northern Giona footwall and the southern Iti hanging wall. The footwall comprises imbricated Mesozoic carbonates of the Parnassos unit, which exhibit large-scale drag folding near and parallel to the fault. The hanging wall comprises deformed sedimentary rocks of the Beotian unit and tectonic klippen of the Eastern Greece unit, forming a southward-tilted neotectonic block with subsidence near the Northern Giona Fault and uplift near the Ypati fault to the north. These two E-W faults represent younger structures disrupting the older NNW-trending tectonic framework. Fault scarps are observed all along the 14 km length of the Northern Giona fault accompanied by cataclastic zones, separating the carbonate formations of the Parnassos Unit from thick scree, slide blocks, boulders and olistholites. Inversion of fault-slip data has shown a mean slip vector of 45°, N004°E, which aligns with the current regional extensional deformation of the area, as confirmed by focal mechanism solutions. Based on the general asymmetry of the alpine units in the hanging wall, we interpret a listric fault geometry at depth using slip-line analysis and we forward modelled a disrupted fault-propagation fold using kinematic trishear algorithms, estimating a total displacement of 6500 m and a throw of approximately 2000 m. Seismic activity in the area of the Northern Giona Fault includes a magnitude 6.1 earthquake in 1852, which caused casualties, rockfalls and extensive damage, as well as a magnitude 5.1 event in 1983. The expected seismic magnitude is deterministically estimated between 6.2 and 6.7, depending on the potential westward continuation of the Northern Giona Fault beyond the Mornos River to the Northern Vardoussia saddle. The seismic hazard zone includes several villages located near the fault, particularly on the hanging wall, where intense landslide activity during seismic events could result in severe damage to regional infrastructure. The neotectonic development of the Northern Giona Fault highlights the importance of extending seismotectonic research into the mountainous regions of central Greece within the alpine formations, beyond the post-orogenic sedimentary basins. Full article
(This article belongs to the Special Issue Active Faulting and Seismicity—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Morphological map of the mountainous region of central Greece between the Corinth Gulf and the Sperchios Valley/Maliakos Gulf. The northern boundary is defined by the northern slopes of Iti Mt, where the Ypati Fault (YF) creates a significant topographic difference of 2000 m, separating the mountainous area from the Sperchios Valley. To the south, the northern slopes of Giona Mt. align with the Northern Giona Fault (NGF), marking a topographic difference of 1500 m between Giona Mt and Iti Mt.</p>
Full article ">Figure 2
<p>Three-dimensional perspective of the studied area with view from the east-northeast (<b>top</b>) and view from the west (<b>bottom</b>). In both views, the NGF and YF are indicated along the abrupt northern slopes of Giona and Iti Mts, respectively. These two subparallel faults have shaped the landscape, creating high mountain peaks at their northern edges on the uplifted footwalls and subsidence to the southern edges.</p>
Full article ">Figure 3
<p>Map illustrating the distribution of seismic epicenters for instrumental (solid circles) and historical earthquakes (purple triangles), along with focal mechanisms for M &gt; 4 (NOA—<a href="http://emsc-csem.org" target="_blank">http://emsc-csem.org</a>), GPS velocity vectors [<a href="#B7-geohazards-05-00065" class="html-bibr">7</a>] and neotectonic faults (black lines). Central Greece’s active deformation results primarily from displacements on E-W-trending faults and some NE-SW strike-slip events.</p>
Full article ">Figure 4
<p>(<b>Top</b>) Geotectonic map of the mountainous region of central Greece between the Sperchios valley to the north and the Corinth gulf to the south. Yellow dashed rectangle shows the extent of the detailed geological map along the NGF of Figure 6. (<b>Middle</b>) NNW—SSE cross section from the Orthrys Mt to the Corinth Gulf showing the geometry of the alpine units and the major neotectonic boundaries, including the YF and the NGF. Dashed blue line indicates the top carbonate of the Parnassos unit. (<b>Bottom</b>) 2D forward model across the NGF showing the deformation of a 10 km layer-cake model with the top horizon corresponding to the pre-Pliocene tectonic framework as built in Figure 11. The Mw 5.1 19 September 1983 earthquake is plotted on the profile alongside the NGF. 1: Pindos Unit, 2: Penteoria unit, 3: Vardoussia unit, 4: Parnassos unit, 5: Beotian unit, 6: Eastern Greece unit, 7: Late Oligocene–Miocene molassic sediments, 8: Late Miocene-Quaternary sediments, 9: Neotectonic and active fault, 10: Miocene Extensional Detachment, including the Itea-Amfissa detachment (IAD) 11: major thrust fault.</p>
Full article ">Figure 5
<p>Panoramic view looking SSW of the NGF (red dashed line) along the northern slopes of Giona Mt. The high-elevated area of northern Giona belonging to the Parnassos unit occur in the footwall whereas the ophiolites and related sediments of the uppermost Eastern Greece unit occur in the hanging wall.</p>
Full article ">Figure 6
<p>Geological map and cross section of the northern Giona region. 1: alluvial deposits, 2: scree deposits, 3: Neogene deposits, 4: Molassic sediments of Oligocene—Middle Miocene, 5: Triassic—Jurassic carbonates of the SubPelagonian unit, 6: Jurassic ophiolites, 7: Beotian Unit with Jurassic—Cretaceous pelagic limestones, 8: Eocene flysch of the Parnassos unit, 9: Mesozoic Carbonate platform of the Parnassos unit, with an older b2 (black) and a younger b3 (red) bauxite horizons in the cross section, 10: Eocene flysch of the Vardousssia unit, 11: Olistholites mainly of Carbonate rocks, 12: Neotectonic and active normal fault, 13: IAD, Itea-Amfissa Detachment, Miocene extensional Detachment, 14: Overthrust, 15: thrust, 16. Base of gravity slide. FW1, HW1: Imbricated tectonic units of the Parnassos nappe. FW, HW: NGF’s footwall and hanging wall.</p>
Full article ">Figure 7
<p>View to the west of the footwall at the central part of the NGF, directly east of the Vrayla peak, at 1800–2000 m altitude. The Parnassos carbonate sequence is characterized by two members in this site: a lower one of Late Jurassic age (Js) and an upper one of Late Jurassic-Early Cretaceous age (J13-K6) separated by a bauxite horizon (b2—[<a href="#B24-geohazards-05-00065" class="html-bibr">24</a>]). In the sketch, solid lines indicate W-dipping strata, while dashed lines indicate N-dipping strata. This change in dip azimuth is characteristic of a kilometric scale normal drag developed near the NGF.</p>
Full article ">Figure 8
<p>View from the east of the NGF. Two notable sites (<b>A</b>,<b>B</b>) where the grooved and striated fault surface is exposed and measurable, showing top-to-N movement.</p>
Full article ">Figure 9
<p>Outcrops of the fault surfaces along the northern Giona slopes. (<b>A</b>) Fault surface (red arrows) at the central part of the NGF. (<b>B</b>) Close view of a north-dipping fault plane along the high slopes at the western side of the NGF. (<b>C</b>) Characteristic cross-section of the fault zone in the eastern part of the NGF, showing a well-developed damage zone that grade to thick fault core (red arrow). (<b>D</b>) Curved fault surface at the eastern termination of the NGF close to the Gravia village.</p>
Full article ">Figure 10
<p>Fault-slip data and calculation of stress tensor with the method of direct inversion for the NGF. Lower hemisphere, equal-area stereographic projections. <b>Left pane</b> shows fault-slip data and the calculated stress axes (σ1, σ2, σ3). <b>Middle pane</b> is a fluctuation histogram of the deviation angle (angle between measured and calculated slip vectors) and stress ratio R(σ2 − σ3)/(σ1 − σ3). <b>Right pane</b> shows the P–T axes.</p>
Full article ">Figure 11
<p>Construction of a 10 km-layer cake model illustrating a five-stage progressive deformation of the NGF footwall and hanging wall through the application of a trishear kinematic model with increasing displacements of 1300 m, 2600 m, 3900 m, 5200 &amp; 6500 m. Details of the trishear model are provided within text. The top layer represents the regional pre-tectonic level, corresponding to the pre-Pliocene deformed state, which includes early orogenic Late Eocene thrust faults (solid lines with triangles) overlying the Parnassos flysch. The Parnassos flysch comprises two members: the lower red pelites and the upper pelitic-sandstone, separated by a dashed line. Unconformably overlying these units are late-orogenic Miocene molasse deposits (wavy brown line) within the Iti and northern Giona fault blocks. The northward-dipping listric geometry of the NGF at depth is based on slip-line analysis [<a href="#B45-geohazards-05-00065" class="html-bibr">45</a>].</p>
Full article ">Figure 12
<p>Map showing the spatial distribution of seismic intensity recorded by various catastrophic phenomena associated with the two significant earthquakes in the region. The solid ellipse represents the macroseismic intensities from the 14 July 1852 earthquake, which had a magnitude of 6.1 (yellow star). The dashed ellipse outlines the area affected by the microseismicity (magnitudes between 2.0 and 4.2) that followed the 19 September 1983, earthquake, which had a magnitude of 5.1 and a fault plane solution of an ENE-WSW normal fault.</p>
Full article ">
24 pages, 59583 KiB  
Article
Proposed Solution for Stony Debris-Flow Control Works in Two Headwater Basins with Morphological Changes
by Mauro Boreggio, Matteo Barbini, Martino Bernard, Massimo Degetto and Carlo Gregoretti
GeoHazards 2024, 5(4), 1346-1369; https://doi.org/10.3390/geohazards5040064 - 18 Dec 2024
Viewed by 489
Abstract
Stony debris flows originating from the two basins of Jaron di Sacomedan and Jaron dei Ross pose a significant threat to the inhabited area of Chiapuzza (Dolomites, Northeastern Italian Alps) and the national road SS 51. In the upper part of the Jaron [...] Read more.
Stony debris flows originating from the two basins of Jaron di Sacomedan and Jaron dei Ross pose a significant threat to the inhabited area of Chiapuzza (Dolomites, Northeastern Italian Alps) and the national road SS 51. In the upper part of the Jaron dei Ross basin, a large scree at the foot of a rocky amphitheater undergoes morphological changes due to frequent rockfalls. Previous mitigation efforts have proven inadequate, and after identifying the causes of their failure, new control measures are being planned. These works aim to direct debris flows towards a deposition area capable of intercepting flows from both the Jaron dei Ross and Jaron di Sacomedan basins. Essentially, the upper works in the Jaron dei Ross basin divert debris flows away from the populated area and channel them to a location where the sediment volume transported by debris flows from both basins can be stored. This solution is designed to protect both the Chiapuzza community and the SS51 national road. Full article
Show Figures

Figure 1

Figure 1
<p>Frontal (<b>a</b>) and plan (<b>b</b>) views of the Sacomedan and Jaron dei Ross basins with the hydrological Catchments (1–4) that contribute to debris-flow occurrences.</p>
Full article ">Figure 2
<p>Plan view of the lower part of the Jaron di Sacomedan and Jaron dei Ross basins showing the control works, including DF channels 1–5, the wall, the old deposition area (red dotted line), and the new extension of the deposition area (red continuous line).</p>
Full article ">Figure 3
<p>Downstream view of the wall and the upper part of the Jaron dei Ross debris-flow channel in 2013 (<b>a</b>) and after the VAIA storm in 2018 (<b>b</b>). Panels (<b>c</b>,<b>d</b>) show lateral views of the wall after the Vaia storm in 2018 and the rock failure in 2021, respectively.</p>
Full article ">Figure 4
<p>Sampling area on DF channel 1 just upstream of the wall (<b>a</b>) and the grain-size analysis of all the samples (<b>b</b>).</p>
Full article ">Figure 5
<p>Aerial map of the two basins, showing the lines on the scree of the Jaron dei Ross basin where the electrical tomography was carried out.</p>
Full article ">Figure 6
<p>The electrical tomography on the upper part of the scree below the cliffs in the Jaron dei Ross basin (<b>a</b>) and the resulting contour lines of the bedrock with the main drainage lines (<b>b</b>) (blue arrows).</p>
Full article ">Figure 7
<p>Frontal view of the upper part of the Jaron dei Ross scree in 2013 (<b>a</b>) and 2018 (<b>b</b>). Panels (<b>c</b>,<b>d</b>) show details from 2018 and 2021, respectively, while (<b>e</b>,<b>f</b>) display the corresponding shaded DEMs. The red line in panel (<b>e</b>) indicates the border of the area hit by the rock failure in 2021.</p>
Full article ">Figure 8
<p>Frontal views of the upper (<b>a</b>) and lower (<b>b</b>) parts of the Jaron di Sacomedan basin.</p>
Full article ">Figure 9
<p>Simulated runoff hydrographs of the Jaron di Sacomedan basin and Catchments 1–4 of Jaron dei Ross basin corresponding to the MPD (<b>a</b>) and MV (<b>b</b>).</p>
Full article ">Figure 10
<p>Computed solid-liquid hydrographs of the Jaron di Sacomedan basin and Catchment 2 of the Jaron dei Ross basin corresponding to the MPD (<b>a</b>) and MV (<b>b</b>).</p>
Full article ">Figure 11
<p>Plan view of control works in the upper part of the Jaron dei Ross scree with the red lines delimiting the new channel, (<b>a</b>) together with the cross and longitudinal sections of DF channel 1 (<b>b</b>) and the mechanically stabilized earth-retaining wall (<b>c</b>).</p>
Full article ">Figure 12
<p>The new deposition area that intercepts the Sacomedan and the DF channels 1–4.</p>
Full article ">Figure 13
<p>The results of hydraulic modeling in terms of maximum flow depth. The case of a solid-liquid input into the Sacomedan channel, solid-liquid and liquid inputs into DF channel 1, and liquid input into DF channel 2 for the MPD (<b>a</b>) and MV (<b>b</b>) scenarios, respectively; the case of solid-liquid input into the Sacomedan channel, liquid input into DF channel 1, and solid-liquid and liquid inputs into DF channel 2 for the MV (<b>c</b>) and MV (<b>d</b>)) scenarios, respectively.</p>
Full article ">Figure 14
<p>The results of hydraulic modeling in terms of deposition and erosion depths. The case of solid-liquid input into the Sacomedan channel, solid-liquid and liquid inputs into DF channel 1, and liquid input into DF channel 2 for the MPD (<b>a</b>) and MV (<b>b</b>) scenarios, respectively; the case of solid-liquid input into the Sacomedan channel, liquid input into DF channel 1, and solid-liquid and liquid inputs into DF channel 2 for the MV (<b>c</b>) and MV (<b>d</b>) scenarios, respectively.</p>
Full article ">Figure A1
<p>A general view (<b>a</b>) of the wall subjected to piping with the particular of foundation (<b>b</b>). The images were taken on 2012 (courtesy of R. Mezzomo).</p>
Full article ">Figure A2
<p>Areal view of the mouth of the Sacomedan channel (<b>left</b>) with the in-time variations of the cross-sections.</p>
Full article ">Figure A3
<p>Deposition volume: schematic plan and profile views (<b>a</b>); detailed 3D (<b>b</b>,<b>d</b>); plan (<b>c</b>) views and the vertical section along the longitudinal axis (<b>e</b>) (modified and redrawn from [<a href="#B23-geohazards-05-00064" class="html-bibr">23</a>]).</p>
Full article ">
20 pages, 98934 KiB  
Article
Automated Snow Avalanche Monitoring and Alert System Using Distributed Acoustic Sensing in Norway
by Antoine Turquet, Andreas Wuestefeld, Guro K. Svendsen, Finn Kåre Nyhammer, Espen Lauvlund Nilsen, Andreas Per-Ola Persson and Vetle Refsum
GeoHazards 2024, 5(4), 1326-1345; https://doi.org/10.3390/geohazards5040063 - 17 Dec 2024
Viewed by 743
Abstract
Avalanches present substantial hazard risk in mountainous regions, particularly when avalanches obstruct roads, either hitting vehicles directly or leaving traffic exposed to subsequent avalanches during cycles. Traditional detection methods often are designed to cover only a limited section of a road stretch, hampering [...] Read more.
Avalanches present substantial hazard risk in mountainous regions, particularly when avalanches obstruct roads, either hitting vehicles directly or leaving traffic exposed to subsequent avalanches during cycles. Traditional detection methods often are designed to cover only a limited section of a road stretch, hampering effective risk management. This research introduces a novel approach using Distributed Acoustic Sensing (DAS) for avalanche detection. The monitoring site in Northern Norway is known to be frequently impacted by avalanches. Between 2022–2024, we continuously monitored the road for avalanches blocking the traffic. The automated alert system identifies avalanches affecting the road and estimates accumulated snow. The system provides continuous, real-time monitoring with competitive sensitivity and accuracy over large areas (up to 170 km) and for multiple sites on parallel. DAS powered alert system can work unaffected by visual barriers or adverse weather conditions. The system successfully identified 10 road-impacting avalanches (100% detection rate). Our results via DAS align with the previous works and indicate that low frequency part of the signal (<20 Hz) is crucial for detection and size estimation of avalanche events. Alternative fiber installation methods are evaluated for optimal sensitivity to avalanches. Consequently, this study demonstrates its durability and lower maintenance requirements, especially when compared to the high setup costs and coverage limitations of radar systems, or the weather and lighting vulnerabilities of cameras. Furthermore the system can detect vehicles on the road as important supplemental information for search and rescue operations, and thus the authorities can be alerted, thereby playing a vital role in urgent rescue efforts. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Map showing Existing Cable (blue) and New Cable extension (orange). Photos taken during the installation are attached from (i) the cabinet at the northern end of the monitoring system, (ii) the vehicle warning system end of the north section, (iii) the vehicle warning system beginning of the south section (iv) an example photo of microtrenching. (<b>b</b>) Map of Norway and the region surrounding the avalanche monitoring zone. Important places are marked, including Holmbuktura, the location of the installation. (<b>c</b>) A cross section sketch showing the details of microtrench cable installation (iv). Direct buried new cable is installed at 15 cm depth and plastic tube covered installation is done at 20 cm depth from the surface.</p>
Full article ">Figure 2
<p>Aerial overview of the Holmbuktura region detailing characteristic avalanche paths and the avalanche monitoring setup. The image on the left (<b>a</b>) shows a comprehensive view of the valley with shaded areas for avalanche zones in north and south. The paths (1–5) along the slope show 5 characteristic avalanche paths, delineating the primary areas of avalanche activity. The cyan line represents the trajectory of the sensor cable installation, placed to capture both the dynamics of avalanches and the road traffic activity. The plot on the right (<b>b</b>) shows the altitude evolution along 5 selected paths, giving the impression of the topography of the region. Image © 2024 Google Earth, Image Landsat/Copernicus, Image © 2024 Maxar Technologies, Image © 2024 CNES/Airbus.</p>
Full article ">Figure 3
<p>Simplified flowchart of the automated avalanche detection and monitoring system. Data is continuously collected and processed through edge computing in two separate modules: (1) vehicle detection and (2) avalanche detection, which operate independently to avoid interference. Detected avalanches and vehicles are then transferred to a central repository and messaging module. This module evaluates risk levels, checks for stranded or at-risk vehicles, and prepares necessary visualizations and alerts. If the risk level exceeds a predefined threshold, the system sends alerts, including plots and messages, via SCADA message system and email using 4G communication.</p>
Full article ">Figure 4
<p>Examples of signals recorded during monitoring with the DAS system in Holmbuktura are shown. The strain rate waterfall plot (Z) highlights features of different events: (<b>a</b>) avalanche activity in the north, (<b>b</b>) avalanche activity in the south, (<b>c</b>) a passenger car, and (<b>d</b>) a snowplow.</p>
Full article ">Figure 5
<p>The power spectral density (PSD) was computed from signals recorded during monitoring with the DAS system in Holmbuktura. The signals represent distinct events, specifically: (<b>a</b>) avalanche activity in the north, (<b>b</b>) avalanche activity in the south, (<b>c</b>) a passenger car, and (<b>d</b>) a snowplow.</p>
Full article ">Figure 6
<p>Most energetic traces from all avalanches are presented as raw signals. In (<b>a</b>) avalanche signals are presented and marked with “Zone N” and “Zone S” showing where the avalanches happened. Event 0 is an avalanche which stopped right before the road it is presented for comparison. Corresponding mean frequency of the 200 s trace is computed and marked on the end of trace. In (<b>b</b>) we present the spectrogram of all avalanches. (<b>c</b>,<b>d</b>) we compare the power spectra of north avalanches and south avalanches respectively. The associated log-averaged power spectra are also plotted.</p>
Full article ">Figure 7
<p>Co-located direct buried “D” (red) and piped loopback cable “P” (blue) traces from avalanches only hitting the southern section are presented (<b>a</b>). Corresponding mean frequency of the entire trace is computed and marked on the trace as well. On right we compare the power spectra from direct buried cable (<b>b</b>) and piped buried (<b>c</b>). The associated log-average power spectra are also plotted.</p>
Full article ">Figure 8
<p>Detailed analysis of the most energetic trace from Event 5. The avalanche signal is analyzed using 20 s sliding time window to investigate avalanche dynamics. In (<b>a</b>), the normalized signal is shown in the time domain; (<b>b</b>) presents the mean frequency of the 20 s time window sliding every 1 s; and (<b>c</b>) displays the power spectra of selected time windows. The colored boxes in (<b>a</b>) indicate time windows, which are highlighted with markers in the mean frequency plot (<b>b</b>) and in the power spectra plot (<b>c</b>) in corresponding colors.</p>
Full article ">Figure A1
<p>Comparison of avalanche dates with historical data of environmental variables. Temperature, snow depth, rain and wind speed from the region covering October 2022 to May 2024 is obtained from OpenMeteo [<a href="#B69-geohazards-05-00063" class="html-bibr">69</a>] is presented. We have plotted the 200 h moving averaged data to visualize long term trends.</p>
Full article ">
18 pages, 6117 KiB  
Article
Multi-Objective Distributionally Robust Optimization for Earthquake Shelter Planning Under Demand Uncertainties
by Kai Tang and Toshihiro Osaragi
GeoHazards 2024, 5(4), 1308-1325; https://doi.org/10.3390/geohazards5040062 - 16 Dec 2024
Viewed by 548
Abstract
Deciding the locations of shelters and how to assign evacuees to these locations is crucial for effective disaster management. However, the inherent uncertainty in evacuation demand makes it challenging to make optimal decisions. Traditional stochastic or robust optimization models tend to be either [...] Read more.
Deciding the locations of shelters and how to assign evacuees to these locations is crucial for effective disaster management. However, the inherent uncertainty in evacuation demand makes it challenging to make optimal decisions. Traditional stochastic or robust optimization models tend to be either too aggressive or overly conservative, failing to strike a balance between risk reduction and cost. In response to these challenges, this research proposes a multi-objective distributionally robust optimization (MODRO) model tailored for shelter location and evacuation allocation. First, an ambiguity set (moment-based or distance-based) is constructed to capture the uncertainty in evacuation demand, reflecting the possible range of outcomes based on demand data from a disaster simulation model. Then, the distributionally robust optimization model considers the “worst-case” distribution within this ambiguity set to minimize construction cost, travel distance, and unmet demand/unused capacity, balancing the trade-off between overly conservative and overly optimistic approaches. The model aims to ensure that shelters are optimally located and evacuees are efficiently allocated, even under the most challenging scenarios. Furthermore, Pareto optimal solutions are obtained using the augmented ε-constraint method. Finally, a case study of Ogu, a wooden density built-up area in Tokyo, Japan, compares the DRO model with stochastic and robust optimization models, demonstrating that the cost obtained by the DRO model is higher than a stochastic model while lower than the worst-case robust model, indicating a more balanced approach to managing uncertainty. This research provides a practical and effective framework for improving disaster preparedness and response, contributing to the resilience and safety of urban populations in earthquake-prone areas. Full article
Show Figures

Figure 1

Figure 1
<p>Two-stage model.</p>
Full article ">Figure 2
<p>An example of a saddle point.</p>
Full article ">Figure 3
<p>Flowchart of PDHG.</p>
Full article ">Figure 4
<p>Location of the study area.</p>
Full article ">Figure 5
<p>Demand points and potential shelters.</p>
Full article ">Figure 6
<p>Sensitivity analysis of the size of the Wasserstein distance-based ambiguity set for <span class="html-italic">f</span><sub>3</sub>.</p>
Full article ">Figure 7
<p>Sensitivity analysis of the size of the moment-based ambiguity set for <span class="html-italic">f</span><sub>3</sub>.</p>
Full article ">Figure 8
<p>(<b>a</b>) Selection of the best solution from the Pareto set for MODRO (<span class="html-italic">r</span><sub>1</sub> = 0.6, <span class="html-italic">r</span><sub>2</sub> = 1). (<b>b</b>) Selection of the best solution from the Pareto set for MODRO (<span class="html-italic">delta</span> = 1200).</p>
Full article ">
14 pages, 2545 KiB  
Article
Spatial Variations of Physical Characteristics of Soil and Their Role in Creating a Model of a Geogenic Radon Hazard Index (GRHI) in the Kuznetsk Coal Basin
by Timofey Leshukov, Konstantin Legoshchin, Maria Savkina, Elizaveta Baranova, Kirill Avdeev and Aleksey Larionov
GeoHazards 2024, 5(4), 1294-1307; https://doi.org/10.3390/geohazards5040061 - 3 Dec 2024
Cited by 2 | Viewed by 719
Abstract
Geographic patterns determine geogenic radon factors that, changing over the territory, form spatial structures of different scales associated with regional and local variations. The study of these structures is important for assessing the possibility of using limited data to predict geogenic radon potential. [...] Read more.
Geographic patterns determine geogenic radon factors that, changing over the territory, form spatial structures of different scales associated with regional and local variations. The study of these structures is important for assessing the possibility of using limited data to predict geogenic radon potential. Our research focuses on the study of the physical properties of soils (moisture, soil density, porosity and void ratio) in the Kuznetsk coal basin. Their variations are studied using statistical methods, a variogram cloud and spatial autocorrelation of data. Soil moisture and porosity have the greatest variability in space and with depth. We conclude that the assessment of geogenic radon predictors requires consideration of the variation coefficient and autocorrelation indices at different scales. Based on the variability of humidity and the fairly homogeneous nature of the studied soils (loams), to assess the radon hazard, it is necessary to study the influence of climatic conditions, since the permeability of the environment for radon will be determined by soil moisture. With the predominance of substantially clayey soils, it is necessary to study the content of 226Ra in the upper horizons, since it is assumed that radon is predominantly diffusely transferred, in which its role is dominant. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the locations of sites for studying the physical properties of soils in the Kuzbass coal basin.</p>
Full article ">Figure 2
<p>Variograms of physical properties of soils at the regional level. (<b>A</b>) density, (<b>B</b>) moisture, (<b>C</b>) porosity coefficient and (<b>D</b>) soil porosity.</p>
Full article ">
19 pages, 6224 KiB  
Article
Implications of Tropical Cyclone Rainfall Spatial–Temporal Variability on Flood Hazard Assessments in the Caribbean Lesser Antilles
by Catherine Nabukulu, Victor. G. Jetten and Janneke Ettema
GeoHazards 2024, 5(4), 1275-1293; https://doi.org/10.3390/geohazards5040060 - 29 Nov 2024
Viewed by 698
Abstract
Tropical cyclones (TCs) significantly impact the Caribbean Lesser Antilles, often causing severe wind and water damage. Traditional flood hazard assessments simplify TC rainfall as single-peak, short-duration events tied to specific return periods, overlooking the spatial–temporal variability in rainfall that TCs introduce. To address [...] Read more.
Tropical cyclones (TCs) significantly impact the Caribbean Lesser Antilles, often causing severe wind and water damage. Traditional flood hazard assessments simplify TC rainfall as single-peak, short-duration events tied to specific return periods, overlooking the spatial–temporal variability in rainfall that TCs introduce. To address this limitation, a new user-friendly tool incorporates spatial–temporal rainfall variability into TC-related flood hazard assessments. The tool utilizes satellite precipitation data to break down TC-associated rainfall into distinct pathways/scenarios, mapping them to ground locations and linking them to specific sections of the storm’s rainfall footprint. This approach demonstrates how different areas can be affected differently by the same TC. In this study, we apply the tool to evaluate rainfall patterns and flood hazards in St. George’s, Grenada, during Hurricane Beryl in 2024. The scenario representing the 75th quantile in Spatial Region 2 (S2-Q0.75) closely matched the actual rainfall observed in the study area. By generating multiple hazard maps based on various rainfall scenarios, the tool provides decision-makers with valuable insights into the multifaced flood hazard risks posed by a single TC. Ultimately, island communities can enhance their early warning and mitigation strategies for TC impacts. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area in St. George’s Parish, Grenada, over which flood characteristics are modelled. The top left inset shows locations impacted by Hurricane Berl’s rainfall and the domain (of 250 km radius) centred on mainland Grenada, within which GPM-IMERG rainfall is analyzed for Hurricane Beryl. The right column shows the 2024 land cover maps of the regions analyzed for flood characteristics. The 2024 land cover is updated from the 2009 land cover map on page 6 in Roberts [<a href="#B31-geohazards-05-00060" class="html-bibr">31</a>] based on high-resolution satellite data and Google Maps.</p>
Full article ">Figure 2
<p>(<b>a</b>) Order of flow processes. Arrows represent ① baseflow (horizontal), interception and infiltration (vertical); ② overland flow (surface); ③ rising channel water levels breaking banks (bidirectional); ④ surface runoff contributing to pluvial, flash and fluvial flooding; ⑤ flood water receding to channel (inward). (<b>b</b>) Information layers for flood modelling. The arrow indicates that for each grid cell, OpenLISEM reads vertically through the information layers. Sourced from Jetten [<a href="#B47-geohazards-05-00060" class="html-bibr">47</a>].</p>
Full article ">Figure 3
<p>Hurricane Beryl’s average rainfall in the study area (equivalent to the DEM extent) as measured by GPM-IMERG Early run from 06:00 UTC 1 July to 00:00 UTC 2 July. The black vertical lines indicate when rainfall from the hurricane’s core impacted the island from 12:00 to 18:00 UTC. The orange line is the moment of landfall on Carricou.</p>
Full article ">Figure 4
<p>Spatial aggregation of Beryl’s rainfall within the defined domain of a 250 km radius centred on mainland Grenada. The analyzed duration is from 06:00 1 July to 00:00 2 July. (<b>a</b>) The spatial regions are mapped on top of the rainfall totals. (<b>b</b>) Visualizations of locations as covered by the rainfall spatial regions S1 to S4.</p>
Full article ">Figure 5
<p>Intensity–duration plots of the rainfall pathways/scenarios derived for the spatial regions at quantile positions 0.5, 0.75, and 0.9. The associated rainfall totals are in bold. (<b>a</b>) S1 mainly poured over the open waters. (<b>b</b>) The rainfall of Spatial Region 2 is what was experienced in the study area. Rainfall scenarios in plots (<b>c</b>,<b>d</b>) were mainly in the outer regions far away from Beryl’s track.</p>
Full article ">Figure 6
<p>Percentage of flood extent due to rainfall of S2 (<b>a</b>) and S1 (<b>b</b>) per flood depth class.</p>
Full article ">Figure 7
<p>Variation in the flooding depth distribution due to the 75th quantiles of the rainfall of S2 (first column), which actually reached Grenada, and S1 (the second column) shows flooding when the region of the highest rainfall hypothetically reaches the island. The analysis is for the port region (<b>a</b>,<b>b</b>), airport region (<b>c</b>,<b>d</b>), and hotel region (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 8
<p>The total number of buildings about the average size of 100 m<sup>2</sup> that are flooded due to S2 (<b>a</b>) and S1 (<b>b</b>) rainfall.</p>
Full article ">Figure A1
<p>Pond adjacent to Maurice Bishop International Airport. Source: The Government Information Service of Grenada (GIS), @ GIS Grenada [<a href="#B61-geohazards-05-00060" class="html-bibr">61</a>].</p>
Full article ">
28 pages, 9654 KiB  
Article
Time Series Foundation Models and Deep Learning Architectures for Earthquake Temporal and Spatial Nowcasting
by Alireza Jafari, Geoffrey Fox, John B. Rundle, Andrea Donnellan and Lisa Grant Ludwig
GeoHazards 2024, 5(4), 1247-1274; https://doi.org/10.3390/geohazards5040059 - 21 Nov 2024
Viewed by 1100
Abstract
Advancing the capabilities of earthquake nowcasting, the real-time forecasting of seismic activities, remains crucial for reducing casualties. This multifaceted challenge has recently gained attention within the deep learning domain, facilitated by the availability of extensive earthquake datasets. Despite significant advancements, the existing literature [...] Read more.
Advancing the capabilities of earthquake nowcasting, the real-time forecasting of seismic activities, remains crucial for reducing casualties. This multifaceted challenge has recently gained attention within the deep learning domain, facilitated by the availability of extensive earthquake datasets. Despite significant advancements, the existing literature on earthquake nowcasting lacks comprehensive evaluations of pre-trained foundation models and modern deep learning architectures; each focuses on a different aspect of data, such as spatial relationships, temporal patterns, and multi-scale dependencies. This paper addresses the mentioned gap by analyzing different architectures and introducing two innovative approaches called Multi Foundation Quake and GNNCoder. We formulate earthquake nowcasting as a time series forecasting problem for the next 14 days within 0.1-degree spatial bins in Southern California. Earthquake time series are generated using the logarithm energy released by quakes, spanning 1986 to 2024. Our comprehensive evaluations demonstrate that our introduced models outperform other custom architectures by effectively capturing temporal-spatial relationships inherent in seismic data. The performance of existing foundation models varies significantly based on the pre-training datasets, emphasizing the need for careful dataset selection. However, we introduce a novel method, Multi Foundation Quake, that achieves the best overall performance by combining a bespoke pattern with Foundation model results handled as auxiliary streams. Full article
Show Figures

Figure 1

Figure 1
<p>Illustration of the construction of a nowcast model for California. The nowcast is a 2-parameter filter on the small earthquake seismicity [<a href="#B42-geohazards-05-00059" class="html-bibr">42</a>,<a href="#B43-geohazards-05-00059" class="html-bibr">43</a>]. (<b>a</b>) Seismicity in the Los Angeles region since 1960, M &gt; 3.29. (<b>b</b>) Monthly rate of small earthquakes as cyan vertical bars. The blue curve is the 36-month exponential moving average (EMA). (<b>c</b>) Mean rate of small earthquakes since 1970. (<b>d</b>) Nowcast curve that is the result of applying the optimized EMA and corrections for the time-varying small earthquake rate to the small earthquake seismicity. (<b>e</b>) Optimized receiver operating characteristic (ROC) curve (red line) used in the machine learning algorithm. Skill is the area under the ROC curve and is used in the optimization. Skill trade-off diagram shows the range of models used in the optimization.</p>
Full article ">Figure 2
<p>Image showing the application of the trained QuakeGPT transformer to an independent, scaled nowcast validation curve (green shading), followed by prediction of future values beyond the end of the nowcast curve (magenta shading). In this model, 36 previous values are used to predict the next value. Dots show the predictions and the solid line shows the nowcast curve whose values are to be predicted. Green dots show the predictions of the transformer up to the last 37 values. The 36 blue dots are predictions that were made and then fed back into the transformer to predict the final point (red dot). In this model, 50 members of an ensemble of runs were used to make the predictions. The dots represent the mean predictions. Brown areas represent the 1-sigma standard deviations to the mean values. In this model, 2021 years of simulation data were used to train the model.</p>
Full article ">Figure 3
<p>Distribution of earthquake epicenters in Southern California (32° N to 36° N, −120° to −114°) from USGS data (1986–2024). The scatter plot shows the spatial density of seismic events used to analyze and optimize spatial bins for earthquake nowcasting.</p>
Full article ">Figure 4
<p>The 500 most active and vulnerable spatial bins, marked in blue, selected for analysis out of the total 2400, based on the frequency of earthquakes from 1986 to 2024. This selection focuses on high-risk areas.</p>
Full article ">Figure 5
<p>Six time series from randomly selected spatial bins, highlighting earthquakes of magnitude greater than 5.</p>
Full article ">Figure 6
<p>The final graph structure representing the 500 most active bins, created using an epsilon of 0.15 degrees. Initially forming a multi-component graph, components are linked to ensure full connectivity.</p>
Full article ">Figure 7
<p>Released energy time series plots for six randomly selected spatial bins, comparing model predictions (GNNCoder one-layer, DilatedRNN, TiDE, iTransformer-M4) against actual observed seismic activities. The brown line represents our GNN approach, which shows a closer match with the actual time series, capturing crucial upward slopes that may signal an impending earthquake. The green and red lines occasionally miss these trends, making more errors where even slight changes in seismic activity are critical. The purple line from the iTransformer-M4 model fails to accurately capture the time series values and exhibits excessive fluctuations.</p>
Full article ">Figure 8
<p>This plot illustrates the spatial bins overlaid on the fault lines to assess the extent to which the fault lines are captured by the bins (graph nodes). It highlights the limitations of the current graph, where some critical fault lines fall outside the spatial bins, impacting the performance of deeper GNN models like the GNNCoder 3-layer model.</p>
Full article ">
27 pages, 66434 KiB  
Article
Floods and Structural Anthropogenic Barriers (Roads and Waterworks) Affecting the Natural Flow of Waters: Hydraulic Modelling and Proposals for the Final Section of the River Segura (Spain)
by Antonio Oliva and Jorge Olcina
GeoHazards 2024, 5(4), 1220-1246; https://doi.org/10.3390/geohazards5040058 - 7 Nov 2024
Viewed by 997
Abstract
Floods are the climate hazard that has the greatest socio-economic and territorial impact on the world. The root causes of these events are atmospheric and hydrological phenomena. However, human action usually aggravates their effects, as it alters the normal functioning of the river [...] Read more.
Floods are the climate hazard that has the greatest socio-economic and territorial impact on the world. The root causes of these events are atmospheric and hydrological phenomena. However, human action usually aggravates their effects, as it alters the normal functioning of the river courses and water flows. The installation of road, rail and hydraulic infrastructures in a floodplain with no prior calculation or appropriate adaptation exacerbates the negative consequences of floods, increasing the extension of the flooded area and the height of the flood waters. This study addresses the problem of the barrier effect generated, on the one hand, by the layout of the N-332 road, as it is built at the same level as the ground, hindering the flow of overflowing water during episodes of flooding, and on the other hand, the channelling wall of the Segura River in the final stretch of its mouth, in the towns of San Fulgencio and Guardamar del Segura. These elements have aggravated flooding in this area. In order to analyse the consequences of the flood, IBER (v.3.3) software has been used to model a flood with similar effects to that of the episode of September 2019. The current situation has also been analysed with two openings in order to determine the effects that a future flood would have. After analysing the results, a proposal to correct the barrier effect of the N-332 road and the new channelling wall of the River Segura has been elaborated upon and then modelled. The results are positive and effective in reducing the negative effects of floods in the lower basin of the River Segura. Full article
Show Figures

Figure 1

Figure 1
<p>Study area: final stretch and mouth of the Segura River in the Segura River basin. Source: National Geographic Institute (IGN) and the Segura Hydrographic Confederation (CHS). Own elaboration. The square with yellow dashed lines represents the study area. The continuous green lines correspond to the limits of the nearest localities in the study area, especially San Fulgencio and Guardamar del Segura.</p>
Full article ">Figure 2
<p>Maximum extent of the former Sinus Ilicitanus reservoir and other lake areas in the lower Segura basin. Source: National Geographic Institute (IGN) and the Segura Hydrographic Confederation (CHS). Own elaboration.</p>
Full article ">Figure 3
<p>An azarbe network in the flood plain of the Segura River in the lower Segura basin. Source: National Geographic Institute (IGN) and the Segura Hydrographic Confederation (CHS). Own elaboration.</p>
Full article ">Figure 4
<p>Behaviour of the overflowing water flows on impact with the N-332 road and the breakwater that divides the two channels of the Segura River: Source: National Geographic Institute (IGN) and the Segura Hydrographic Confederation (CHS). Own elaboration.</p>
Full article ">Figure 5
<p>Barrier effect of the N-332 in the flooding event in September 2019. Source: Diario Información 14 September 2019. Note: The blue arrows mark the natural direction of the flows. You can see when the waters are overflowing the N-332 at a given point. It can also be seen how the channels are completely saturated with water and overflowed, flooding the adjacent land.</p>
Full article ">Figure 6
<p>Wall or breakwater separating the old channel from the new channel. Source: Antonio Oliva Cañizares. Note: In the image on the left, we can see the point where the watercourses drain into the old Segura riverbed. This riverbed has become clogged, and the presence of reeds is reducing its capacity. To the right of the riverbed, we can see the wall separating the old riverbed from the new one. In fact, the old riverbed is higher than the new one. The image on the right shows the width of the new Segura riverbed and the wall or breakwater separating the two riverbeds from the opposite side.</p>
Full article ">Figure 7
<p>Study area flooded in September 2019. Source: Diario Información 14 September 2019. Note: This image clearly shows the barrier effect of the N-332 road and the breakwater that divides the old channel from the new one, resulting in the whole area being flooded by the overflowing waters (the yellow arrows). The overflowing of the azarbes can also be seen.</p>
Full article ">Figure 8
<p>Digital terrain model obtained for modelling. Source: Own elaboration. Note: Note how the heights of the buildings stand out. This is evidence of the combination of the ground with the buildings thanks to the LIDAR points.</p>
Full article ">Figure 9
<p>Result of the geometry carried out in the QGIS programme. Source: Own elaboration.</p>
Full article ">Figure 10
<p>Comparison between the real state of the mouth of the Segura River (<b>left</b>) and the result of the modifications made to the Segura riverbed: elimination of breakwaters and separating wall and joining of both riverbeds (<b>right</b>) (red arrows). Source: IBER. Own elaboration.</p>
Full article ">Figure 11
<p>Elevation of the N-332 road (<b>left</b>) and culvert including the space between the piers of the raised bridge and the general view of the mesh developed in the study area (<b>right</b>). Source: Own elaboration.</p>
Full article ">Figure 12
<p>Rural roads and cultivated areas modified to provide a natural outlet for floodwaters. Source: National Geographic Institute (IGN) and the Segura Hydrographic Confederation (CHS). Own elaboration. Note: Arrows indicate modified paths to remove barriers to water flow.</p>
Full article ">Figure 13
<p>IBER simulation (6 h) of the September 2019 flood in the final stretch of the Segura River. Source: Own elaboration. Note: please see the <a href="#app1-geohazards-05-00058" class="html-app">Video S1</a> time of simulation 21,600 (s).</p>
Full article ">Figure 14
<p>IBER simulation (12 h) of the September 2019 flood in the final stretch of the Segura River. Source: Own elaboration. Note: please see the <a href="#app1-geohazards-05-00058" class="html-app">Video S1</a> time of simulation 43,200 (s).</p>
Full article ">Figure 15
<p>IBER simulation (12 h) of the September 2019 flood in the final stretch of the Segura River with the two current openings. Source: Own elaboration. Note: please see the <a href="#app1-geohazards-05-00058" class="html-app">Video S2</a> time of simulation 43,200 (s).</p>
Full article ">Figure 16
<p>Hydraulic simulation of the behaviour of a flood having carried out the actions of creating a single channel and raising the N-332 road (6 h) (IBER) Source: Own elaboration. Note: please see the <a href="#app1-geohazards-05-00058" class="html-app">Video S3</a> time of simulation 21,600 (s).</p>
Full article ">Figure 17
<p>Hydraulic simulation of the behaviour of a flood having carried out the actions of creating a single channel and raising the N-332 road (12 h) (IBER) Source: Own elaboration. Note: please see the <a href="#app1-geohazards-05-00058" class="html-app">Video S3</a> time of simulation 43,200 (s).</p>
Full article ">Figure 18
<p>Hydraulic simulation of the behaviour of a flood having carried out the actions of creating a single channel and raising the N-332 road (24 h) (IBER) Source: Own elaboration. Note: please see the <a href="#app1-geohazards-05-00058" class="html-app">Video S3</a> time of simulation 86,400 (s).</p>
Full article ">Figure 19
<p>Current state of the N-332 road (<b>left</b>) and sketch of the proposed raising of the road in pilasters with sufficient space for the overflowing water to circulate without difficulty (<b>right</b>). Source: Photograph taken by Antonio Oliva Cañizares (20 June 2023) (<b>left</b>) and sketch prepared by artificial intelligence (IA) (ChatGPT) (<b>right</b>).</p>
Full article ">
13 pages, 2809 KiB  
Article
Topographic–Vegetation Interactions on an Incipient Foredune Field Post-Tropical Storm
by Jean T. Ellis, Michelle E. Harris and Brianna F. Barrineau
GeoHazards 2024, 5(4), 1207-1219; https://doi.org/10.3390/geohazards5040057 - 4 Nov 2024
Viewed by 825
Abstract
Sand dunes protect the most important economic and ecologically critical landscapes from coastal hazards (storms and high-tide flooding). The characteristics of the dune affect their protective ability. This paper qualitatively and quantitatively assesses the relationships between pre- and post-storm conditions for vegetation and [...] Read more.
Sand dunes protect the most important economic and ecologically critical landscapes from coastal hazards (storms and high-tide flooding). The characteristics of the dune affect their protective ability. This paper qualitatively and quantitatively assesses the relationships between pre- and post-storm conditions for vegetation and the morphology of an incipient dune system along the South Carolina coast. Field-based dune vegetation and morphology measurements were obtained before and after tropical storm Dorian (2019). Vegetation is assessed with respect to distribution and functional type, and subgroups are introduced to categorize land cover transitions. At the quadrat scale (0.2 m2) following the storm, there was a shift from stabilizer to builder, a decrease of sand (2%), and the vegetation remained consistent at around 61% of the land cover. Transect-level analysis (0.2 m × 1.0 m) revealed distinct variability concerning post-storm morphology change in the extreme study site extents. Dorian resulted in approximately 10% volumetric loss over the entire study site (101 m2). This study demonstrated changes to a dune system following a tropical storm with wind as the dominant forcing factor. This study revealed that vegetation presence is not broadly correlated with reduced levels of post-storm erosion. Full article
Show Figures

Figure 1

Figure 1
<p>Study site location at 53rd Avenue on Isle of Palms, South Carolina, U.S.A. The topleft panel insert is the State of South Carolina, U.S.A., where the yellow star denotes the location of Isle of Palms (primary image), which is bounded on the northeast and southwest by Dewees and Breach Inlets, respectively. The white dashed line marks the spatial extent of the 2018 beach nourishment. The yellow inset panel denotes the location of 53rd Avenue, where cross-shore transects are presented. The dashed black line is the seaward extent of the developed secondary foredune system. Base imagery: Google Earth Pro (2019), Maxar Technologies.</p>
Full article ">Figure 2
<p>Data processing methods where (<b>a</b>) shows the combined pre- and post-Dorian classified LC sections to create 0.2 m × 1.0 m classified polygons; (<b>b</b>) is spatially reduced DEMs; and (<b>c</b>) shows the polygons considered for analysis. In panels (<b>a</b>,<b>c</b>): Sa is sand, S is stabilizer, B is builder, and O is other.</p>
Full article ">Figure 3
<p>Pre- and post-<span class="html-italic">Dorian</span> ((<b>a</b>,<b>b</b>), respectively) percentages of sand, stabilizers, and builders along each transect. Land covers were normalized to 100% for each transect for the pre- and post-storm conditions.</p>
Full article ">Figure 4
<p>Classified 0.2 m × 1.0 m LC polygons for pre- (<b>a</b>) and post- (<b>b</b>) Dorian. The pre-and post-<span class="html-italic">Dorian</span> DEMs are shown in (<b>c</b>,<b>d</b>). These results have been clipped to correspond with the methods associated with <a href="#geohazards-05-00057-f002" class="html-fig">Figure 2</a>. The offshore and onshore locations shown in (<b>a</b>) are the same in (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 4 Cont.
<p>Classified 0.2 m × 1.0 m LC polygons for pre- (<b>a</b>) and post- (<b>b</b>) Dorian. The pre-and post-<span class="html-italic">Dorian</span> DEMs are shown in (<b>c</b>,<b>d</b>). These results have been clipped to correspond with the methods associated with <a href="#geohazards-05-00057-f002" class="html-fig">Figure 2</a>. The offshore and onshore locations shown in (<b>a</b>) are the same in (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 5
<p>LC Subgroups (left stripes) and DEM (right stripes) change maps for T10 to T1. Numbers indicate the elevation percent change for each transect line.</p>
Full article ">Figure 6
<p>Elevation change and count according to subgroup.</p>
Full article ">
17 pages, 4479 KiB  
Article
Climate Change Impact on the Stability of Soil Slopes from a Hydrological and Geotechnical Perspective
by Prodromos N. Psarropoulos, Nikolaos Makrakis and Yiannis Tsompanakis
GeoHazards 2024, 5(4), 1190-1206; https://doi.org/10.3390/geohazards5040056 - 1 Nov 2024
Viewed by 1684
Abstract
Climate change (CC) is expected to cause significant changes in weather patterns, leading to extreme phenomena. Specifically, the intensity of precipitation extremes is continuously escalating, even in regions with decreasing average precipitation levels. Given that CC leads to long-term shifts in [...] Read more.
Climate change (CC) is expected to cause significant changes in weather patterns, leading to extreme phenomena. Specifically, the intensity of precipitation extremes is continuously escalating, even in regions with decreasing average precipitation levels. Given that CC leads to long-term shifts in weather patterns and may affect the precipitation characteristics (i.e., frequency, duration, and intensity) directly related to groundwater table fluctuations and soil erosion phenomena, it has the potential to significantly affect soil slope instabilities. In turn, slope stability and the structural integrity of nearby structures and infrastructure will be affected. Accordingly, the present paper focuses on the impact of CC on the geohazard of soil slope instability by considering both hydrological aspects, i.e., the impact on rainfall intensity on the groundwater table and the geotechnical aspects of this complex problem. The findings reveal that the impact of CC on potential slope instabilities can be detrimental or even beneficial, depending on the specific site and water conditions. Therefore, it is essential to do the following: (a) collect all the available data of the area of interest, (b) assess their variations over time, and (c) examine each potentially unstable slope on a case-by-case basis to properly mitigate this geohazard. Full article
(This article belongs to the Special Issue Landslide Research: State of the Art and Innovations)
Show Figures

Figure 1

Figure 1
<p>Flowchart depicting the indirect relationship between climate and structural integrity of any structure and infrastructure located on soil slopes.</p>
Full article ">Figure 2
<p>Water balance on a soil slope during rainfall (adopted from [<a href="#B52-geohazards-05-00056" class="html-bibr">52</a>]).</p>
Full article ">Figure 3
<p>Potential changes in factor of safety (modified from [<a href="#B53-geohazards-05-00056" class="html-bibr">53</a>]). A factor of safety &lt; 1.0 indicates an unstable slope.</p>
Full article ">Figure 4
<p>(<b>a</b>) Slope instability phenomena and (<b>b</b>) soil erosion phenomena close to the routing of <span class="html-italic">TAP</span> (photos taken by P.N. Psarropoulos).</p>
Full article ">Figure 5
<p>Sketch of a planar translational slide under dry conditions.</p>
Full article ">Figure 6
<p>Sketch of a planar translational slide with inclined groundwater table.</p>
Full article ">Figure 7
<p>Time histories of: (<b>a</b>) the cumulative rainfall and (<b>b</b>) the absolute rainfall of the two examined rainfall scenarios.</p>
Full article ">Figure 8
<p>Time histories of infiltration and runoff for the three examined soil permeability levels: (<b>a</b>) high, (<b>b</b>) medium, and (<b>c</b>) low.</p>
Full article ">Figure 9
<p>Variation of <span class="html-italic">FS</span> with respect to slope inclination, <span class="html-italic">θ</span>, without groundwater table for the three examined soil types, as compared to the critical <span class="html-italic">FS</span> = 1: (<b>a</b>) c = 20 kPa, φ = 30°, (<b>b</b>) c = 0 kPa, φ = 30° and (<b>c</b>) c = 20 kPa, φ = 0°.</p>
Full article ">Figure 10
<p>Variation of FS′ with respect to slope inclination, θ, with inclined groundwater table for the three examined soil types. Cases (<b>a</b>–<b>c</b>) refer to <span class="html-italic">z<sub>w</sub></span> = <span class="html-italic">z</span>/2, while cases (<b>d</b>–<b>f</b>) to <span class="html-italic">z<sub>w</sub></span> = 0. A direct comparison with the critical <span class="html-italic">FS</span> = 1 can be performed for all cases.</p>
Full article ">
18 pages, 8479 KiB  
Article
Tectonic Control of Aseismic Creep and Potential for Induced Seismicity Along the West Valley Fault in Southeastern Metro Manila, Philippines
by Rolly E. Rimando and Peter L. K. Knuepfer
GeoHazards 2024, 5(4), 1172-1189; https://doi.org/10.3390/geohazards5040055 - 30 Oct 2024
Viewed by 1195
Abstract
Vertical creep along 15 ground ruptures within a 15 km long and 1.5 km wide zone has been occurring along the southeastern part of Metro Manila. Though the unusually high rates of vertical slip point to excessive groundwater withdrawal as the trigger, the [...] Read more.
Vertical creep along 15 ground ruptures within a 15 km long and 1.5 km wide zone has been occurring along the southeastern part of Metro Manila. Though the unusually high rates of vertical slip point to excessive groundwater withdrawal as the trigger, the evidence presented herein indicates that these may not be simple irregular subsidence fissures. Tectonic control of creep along these traces is suggested by the following: the occurrence of some of these ground ruptures along pre-existing scarps that coincide with topographic and lithologic boundaries, the left-stepping en echelon pattern of surface rupturing, and the distribution of the creeping zone within the dilational gap of the dextral strike-slip West Valley Fault (WVF). Furthermore, interpretation of an exposure across one of the creeping faults indicates reactivation by creep of a pre-existing tectonic fault zone. The paleoseismic evidence also suggests that the pre-creep slips are coseismic and dominantly strike-slip. Recognizing the occurrence of coseismic slip preceding aseismic creep is a primary consideration in assessing the potential of the WVF’s creeping segment and its adjacent segments in generating earthquakes. Tighter groundwater extraction regulations may be necessary to avoid exacerbating the effects of vertical ground deformation and the occurrence of induced seismicity. Full article
Show Figures

Figure 1

Figure 1
<p>The creeping zone corresponds to a segment (Sucat–Biñan or segment II [<a href="#B3-geohazards-05-00055" class="html-bibr">3</a>]) of the Valley Fault System (VFS) or Marikina Valley Fault System (MVFS) on the southeastern part of Metro Manila. The VFS belongs to a system of faults and subduction zones that accommodates part of the deformation in central Luzon due to the northwestward drift of the Philippine Sea Plate (PSP) towards the Sunda Plate (SP). (<b>a</b>) The VFS region is bounded by the Philippine fault zone (PFZ) on the east, the East Zambales Fault (EZF) on the west, and the Lubang Fault (LF) on the south. Also shown are the East Luzon Trough (ELT) and the Philippine Trench (PT) to the east and the Manila Trench to the west. Direction and rate of PSP motion after Seno et al. [<a href="#B5-geohazards-05-00055" class="html-bibr">5</a>]. The active faults are from the Philippine Institute of Volcanology and Seismology [<a href="#B6-geohazards-05-00055" class="html-bibr">6</a>], Rimando [<a href="#B2-geohazards-05-00055" class="html-bibr">2</a>], Rimando and Knuepfer [<a href="#B3-geohazards-05-00055" class="html-bibr">3</a>], and references therein. (<b>b</b>) The VFS extends for 135 km from the Sierra Madre to the eastern part of the Tagaytay Ridge. White dashed lines indicate approximate boundaries of the zone of volcanoes (Macolod Corridor) to the south of the VFS. White solid arrows indicate direction of extension within the Macolod Corridor. The northern part of the Marikina Valley (MV) is a pull-apart basin formed within the gap between the two major segments of the VFS (the West Valley Fault, or WVF, and the East Valley Fault, or EVF). The minor structural/geometric segments of the VFS are indicated by Roman numerals I to X. The creeping segment corresponds to segment II. Segmentation after Rimando [<a href="#B2-geohazards-05-00055" class="html-bibr">2</a>] and Rimando and Knuepfer [<a href="#B3-geohazards-05-00055" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) The WVF creeping segment (segment II in <a href="#geohazards-05-00055-f001" class="html-fig">Figure 1</a>). The creeping zone consists of lithologic units derived from nearby volcanic centers. Most of the creeping faults occur along scarps identified in the 1965 aerial photos and in topographic maps based on older aerial photos. (<b>b</b>–<b>h</b>) Photos of vertical displacement and damages to ground and man-made structures along creep scarps. Locations and dates when photos were taken are indicated under each photo. For variation in displacement and slip/displacement rates, refer to Rimando et al. [<a href="#B4-geohazards-05-00055" class="html-bibr">4</a>]. The encircled 6a to 6f and the symbol <span class="html-fig-inline" id="geohazards-05-00055-i001"><img alt="Geohazards 05 00055 i001" src="/geohazards/geohazards-05-00055/article_deploy/html/images/geohazards-05-00055-i001.png"/></span> refer to the locations of scarp profiles used for height estimation. NPC, GRV, ADL, VOS, and JUA refer to deformation survey sites where displacements have been regularly measured (2–3 times/yr) since 1999. Slip rates derived from displacement measurement at these sites are in <a href="#geohazards-05-00055-f003" class="html-fig">Figure 3</a>. Source of <span class="html-italic">geology</span>: Bureau of Mines and Geosciences [<a href="#B17-geohazards-05-00055" class="html-bibr">17</a>].</p>
Full article ">Figure 3
<p>Plot of cumulative vertical displacement estimates from various sites within the creeping segment since the 1990s. The numbers are slip rates in cm/y. Displacements in the 1990s (left side of dashed vertical line) were measured using a hand level and a graduated telescoping rod. The 1999–2019 displacement data used are from Rimando et al. [<a href="#B4-geohazards-05-00055" class="html-bibr">4</a>] and references therein. A more precise method using an electronic digital level, a bar-code leveling staff, and several fixed benchmarks was employed to generate the 1999–2019 displacement data by the joint Philippine Institute of Volcanology and Seismology–Tokyo Institute of Technology (PHIVOLCS-TIT) team. The 2019–2023 data were generated using the same method.</p>
Full article ">Figure 4
<p>Segments I, II, and III of the WVF. The mapping of segments I and III and the rest of the VFS was based mainly on the recognition of offset features (e.g., streams and spurs) and landforms associated with active faults. The mapping of segment II was initially based on the recognition of old scarps and creep scarps, many of which followed the pre-existing scarps. The vertical displacements along the creeping segments were based on field measurements carried out mostly in 1996. The left portion of the plot in <a href="#geohazards-05-00055-f003" class="html-fig">Figure 3</a> was constructed based on these measurements, while the right portion of the plot was based on the results of precise leveling that started in 1999 at sites labeled NPC, GRV, ADL, VOS, and JUA. The inset indicates the location of the segments and shows NW-SE extension (open arrows) within the dilation gap between segments I and III. The encircled 7a to e refer to the locations of scarp profiles derived from precise leveling data for height estimation. Site of paleoseismic section along segment II is also shown. Modified from Rimando [<a href="#B2-geohazards-05-00055" class="html-bibr">2</a>] and Rimando and Knuepfer [<a href="#B3-geohazards-05-00055" class="html-bibr">3</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Ground rupture along the southernmost segment of the creeping zone. Creep scarp occurs near the foot of an old scarp at Juana Subdivision, Biñan, Laguna. (<b>b</b>) Evidence of pre-creep scarp coinciding with the creeping zone’s southernmost segment (see <a href="#geohazards-05-00055-f002" class="html-fig">Figure 2</a>). The scarp is indicated by hachures (location marked by arrows) on a 1:50,000-scale topographic map that was drawn based on 1947–1952 aerial photos. North points to the left side of the figure and is parallel to the gridline. Ground ruptures in (<b>a</b>) occur near the southernmost part of the segment. (<b>c</b>) Stereopair of the general area from 1966 aerial photos showing the same scarp in (<b>b</b>). (<b>d</b>) The active creep scarp at Villa Olympia Subdivision (VOS) (San Pedro, Laguna) along the southernmost creeping segment at VOS, which also occurs along the scarp in the topographic map in (<b>b</b>) and in the 1966 aerial photo (<b>c</b>). Photo taken in September, 2023. Part of the road across the scarp had been resurfaced. The creep scarp, which appears to be superimposed on a pre-existing scarp, was first noticed in 1996 when it was only ~0.15 m high, but it reached a height of 0.725 m by 2022. (<b>e</b>) Profiles of the larger scarp hosting the active scarp were based on the precise leveling surveys conducted in 1999 and 2023. Profiles were drawn along the road, which is oblique to the NE-trending scarp. The estimated total scarp heights were 6.48 m and ~7 m in 1999 and 2023, respectively. <a href="#geohazards-05-00055-f006" class="html-fig">Figure 6</a>a shows a profile taken along a line close to and perpendicular to the scarp in (<b>d</b>).</p>
Full article ">Figure 6
<p>Profiles drawn perpendicular to the segments on the western part of the WVF’s creeping zone (segment II). The arrows indicate the location of the scarps due to the current creep. The locations of the profiles (<b>a</b>–<b>f</b>) are indicated in <a href="#geohazards-05-00055-f002" class="html-fig">Figure 2</a>a. Scarp height estimates are indicated.</p>
Full article ">Figure 7
<p>Profiles across selected creeping segments in 5 precise leveling sites. The locations of the profiles (<b>a</b>–<b>e</b>) are indicated in <a href="#geohazards-05-00055-f004" class="html-fig">Figure 4</a>. Scarp height estimates are shown for each profile.</p>
Full article ">Figure 8
<p>(<b>a</b>) Section of stream exposure across one of the creeping traces at Magdaong River, Muntinlupa City. Exposure shows prior surface-rupturing events, which were, most likely, generated by strike-slip faulting. (<b>b</b>) Simplified log of stream exposure across one of the creeping traces at Magdaong River, Muntinlupa City. The location of the site is indicated in <a href="#geohazards-05-00055-f002" class="html-fig">Figure 2</a> and <a href="#geohazards-05-00055-f004" class="html-fig">Figure 4</a>. Older strike-slip fault ruptures cut through buried sediments only, while modern creep has vertically displaced the present surface. Fault contact between units 2 and 3 at the bottom part of the section is indistinct, so this part of the log is quite interpretive. The section was originally logged at a 1:20 scale. Descriptions of stratigraphic units appear in <a href="#geohazards-05-00055-t001" class="html-table">Table 1</a>. The numbered letters refer to fault strands. The upper right part of the figure also shows the location of scarp profiles c and d in <a href="#geohazards-05-00055-f006" class="html-fig">Figure 6</a> and of the Magdaong River exposure. (<b>c</b>) The upward termination of fault strands c<sub>3</sub> to c<sub>6</sub> at the base of unit 6 represents one pre-creep event, interpreted as a sudden coseismic slip. Location of photo shown in (<b>b</b>).</p>
Full article ">
20 pages, 1893 KiB  
Article
Numerical Modeling of Tsunamis Generated by Subaerial, Partially Submerged, and Submarine Landslides
by Tomoyuki Takabatake and Ryosei Takemoto
GeoHazards 2024, 5(4), 1152-1171; https://doi.org/10.3390/geohazards5040054 - 21 Oct 2024
Viewed by 1129
Abstract
Using the existing two-dimensional experimental data and Open-source Fields Operation and Manipulation (OpenFOAM) software, this study performs a comprehensive comparative analysis of three types of landslide-generated tsunamis (subaerial, partially submerged, and submarine). The primary objective was to assess whether numerical simulations can accurately [...] Read more.
Using the existing two-dimensional experimental data and Open-source Fields Operation and Manipulation (OpenFOAM) software, this study performs a comprehensive comparative analysis of three types of landslide-generated tsunamis (subaerial, partially submerged, and submarine). The primary objective was to assess whether numerical simulations can accurately reproduce the experimental results of each type and to compare the predictive equations of the tsunami amplitudes derived from experimental and simulated data. The mesh size and dynamic viscosity parameters were initially optimized for a specific partially submerged landslide tsunami scenario and then applied across a broader range of experimental scenarios. Most of the simulated wave amplitudes remained within the 50% error margin, although significant discrepancies were observed between landslide types. When focusing on the crest amplitude of the first wave, the simulations of subaerial landslides least deviated from the experimental data, with a mean absolute percentage error of approximately 20%, versus approximately 40% for the partially submerged and submarine landslides. The predictive equations derived from the simulations closely matched those from the experimental data, confirming that OpenFOAM can effectively capture complex landslide–tsunami dynamics. Nonetheless, variations in the coefficients related to slope angles highlight the need for further calibration to enhance the simulation fidelity. Full article
(This article belongs to the Topic Natural Hazards and Disaster Risks Reduction, 2nd Volume)
Show Figures

Figure 1

Figure 1
<p>Experimental layout of the benchmark experiments [<a href="#B32-geohazards-05-00054" class="html-bibr">32</a>].</p>
Full article ">Figure 2
<p>Geometric definition of the experimental parameters.</p>
Full article ">Figure 3
<p>Numerical setup displaying the overall layout (<b>left</b>) and a zoomed-in view of the area near the slope (<b>right</b>) with a mesh size of 0.01 m.</p>
Full article ">Figure 4
<p>Comparison of time histories of water surface elevations, observed in the sensitivity studies with different mesh sizes and dynamic viscosity levels.</p>
Full article ">Figure 5
<p>Snapshots of the interaction between the granular landslide and the water body during a partially submerged landslide tsunami under the optimized numerical settings (mesh size = 0.01 m; dynamic viscosity = 1.0 × 10<sup>−3</sup> m<sup>2</sup>/s). The color map represents the velocity magnitude of the water body and the vectors indicate the flow direction. The landslide is colored in yellow. (<b>a</b>) Initial movement of the landslide mass; (<b>b</b>) First wave propagating offshore; (<b>c</b>) Second wave propagating offshore; (<b>d</b>) The point at which the landslide mass reaches the bottom surface.</p>
Full article ">Figure 6
<p>Comparison of the experimental [<a href="#B32-geohazards-05-00054" class="html-bibr">32</a>] and simulated (in this study) wave amplitudes. In each panel, the blue circles, red diamonds, and green squares are the comparison results of subaerial landslides, partially submerged landslides, and submarine landslides, respectively; the solid line is the identity line and the dotted lines are the 50% error margins.</p>
Full article ">Figure 7
<p>Comparison of the wave periods (left column), celerities (middle column), and wavelengths (right column) between the experiments in [<a href="#B32-geohazards-05-00054" class="html-bibr">32</a>] and the simulations of the present study. The solid line is the identity line, and the dotted lines represent the 50% error bounds.</p>
Full article ">Figure 8
<p>Comparisons of leading-wave amplitudes between (<b>a</b>) the experimental values and the values estimated using Equation (6), (<b>b</b>) the simulated values and those estimated using Equation (7), and (<b>c</b>) the predicted values of Equation (7) and the predicted values of Equation (6). The solid line is the identity line and the dotted lines represent the 50% error bounds.</p>
Full article ">
27 pages, 8508 KiB  
Article
Towards a Modern and Sustainable Sediment Management Plan in Mountain Catchment
by Alessio Cislaghi, Emanuele Morlotti, Vito Giuseppe Sacchetti, Dario Bellingeri and Gian Battista Bischetti
GeoHazards 2024, 5(4), 1125-1151; https://doi.org/10.3390/geohazards5040053 - 17 Oct 2024
Viewed by 1056
Abstract
Sediment management is fundamental for managing mountain watercourses and their upslope catchment. A multidisciplinary approach—not limited to the discipline of hydraulics—is necessary for investigating the alterations in sediment transport along the watercourse by detecting those reaches dominated by erosion and deposition processes, by [...] Read more.
Sediment management is fundamental for managing mountain watercourses and their upslope catchment. A multidisciplinary approach—not limited to the discipline of hydraulics—is necessary for investigating the alterations in sediment transport along the watercourse by detecting those reaches dominated by erosion and deposition processes, by quantifying the sediment volume change, by assessing the functionality of the existing torrent control structures, and by delimitating the riparian vegetation patches. To pursue these goals, specific continuous monitoring is essential, despite being extremely rare in mountain catchments. The present study proposed an integrated approach to determine the hydro-morphological–sedimentological–ecological state of a mountain watercourse though field- and desk-based analyses. Such an integral approach includes a rainfall–runoff model, a morphological change analysis and the application of empirical formulations for estimating peak discharge, mobilizable sediment/large wood volume and watercourse hydraulic capacity, at reach and catchment scales. The procedure was tested on the Upper Adda River catchment (North Italy). The results identified where and with what priority maintenance and monitoring activities must be carried out, considering sediment regime, torrent control structures and vegetation. This study is an example of how it is possible to enhance all existing information through successive qualitative and quantitative approximations and to concentrate new resources (human and economic) on specific gaps, for drafting a scientifically robust and practical sediment management plan. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the surveyed 12 km of Adda River flowing north to south, in the Upper Valtellina (Lombardy, North Italy).</p>
Full article ">Figure 2
<p>Framework of the hydro–geo-morphological, sedimentological and ecological integrated analysis (HySEcA), on which basis operational and monitoring measures are proposed in the sediment management plan.</p>
Full article ">Figure 3
<p>Photographs of transverse torrent control structures that show the four conditions of the Loss of Functionality Index (<span class="html-italic">LoFI</span>) according to the percentage of spillway occupied by the sediment. Value 1, or Low, indicates a spillway covered for less than 50%; Value 2, or Medium-low, indicates a coverage between 50 and 75%; Value 3, or Medium-high, indicates a coverage between 75 and 90%; and Value 4, or High, indicates a coverage more than 90%.</p>
Full article ">Figure 4
<p>Photographs of riparian vegetation according to colonization density (negligible, low, medium and high).</p>
Full article ">Figure 5
<p>The identification of 14 reaches (from 4A to 8A) and 28 subcatchments (from 4A_1 to 8A_7) for the study area.</p>
Full article ">Figure 6
<p>The locations and functionality assessment of the inspected transverse torrent control structures along the surveyed watercourse.</p>
Full article ">Figure 7
<p>The proposed framework to draft a modern and sustainable sediment management plan.</p>
Full article ">Figure 8
<p>Locations of retention check dams and retention basins in the study area.</p>
Full article ">Figure 9
<p>Catchment classification based on the hydrological and sediment transport processes using the discriminating limits of the different categories according to Wilford et al. (2004) [<a href="#B97-geohazards-05-00053" class="html-bibr">97</a>].</p>
Full article ">Figure 10
<p>Relationships between the Melton Ratio and the mean value of connectivity index of all subcatchments of the study area.</p>
Full article ">Figure 11
<p>Flowchart of appropriate monitoring activities for integrating the sediment management plan.</p>
Full article ">
32 pages, 18414 KiB  
Article
La Palma 2021 Eruption (Canary Islands): Measurements and Modelling of Lava Flow Cooling Rates and Applications for Infrastructure Reconstruction and Risk Mitigation
by Luis González-de-Vallejo, Aaron Álvarez-Hernández, Mercedes Ferrer, John P. Lockwood, Nemesio M. Pérez, Pedro A. Hernández, Ana Miranda-Hardisson, José A. Rodríguez-Losada, David Afonso-Falcón, Héctor de-los-Ríos, Javier Páez-Padilla and Luis E. Hernández-Gutiérrez
GeoHazards 2024, 5(4), 1093-1124; https://doi.org/10.3390/geohazards5040052 - 4 Oct 2024
Viewed by 2731
Abstract
On 19 September 2021, a strombolian volcanic eruption began on the island of La Palma in the Canary Islands. This event resulted in the destruction of 73 km of roads, urban infrastructure, numerous houses, and agricultural crops, affecting approximately 7200 people and causing [...] Read more.
On 19 September 2021, a strombolian volcanic eruption began on the island of La Palma in the Canary Islands. This event resulted in the destruction of 73 km of roads, urban infrastructure, numerous houses, and agricultural crops, affecting approximately 7200 people and causing losses exceeding 1.2 billion euros. Around 12 km2 were covered by aa and pahoehoe lava flows, which reached thicknesses of over 70 m. Following the end of the eruption, thermal, geological, and geotechnical site investigations were carried out for the reconstruction and territorial and urban planning, with the main objectives focused on opening roads through hot lava, constructing new urban settlements in areas covered by lava flows, and facilitating the agricultural recovery. The primary challenges to reconstruction included the very slow cooling rate of the lava, resulting in persistent high temperatures, exceeding 500 °C, its highly heterogeneous geotechnical properties with numerous cavities and lava caves, and the presence of toxic gases. Site investigations included geotechnical boreholes, seismic geophysics and ground-penetration radar, and temperature measurements of lava flows using drones and thermocouple devices inside boreholes. To estimate the cooling rates of the lava flows, two physical cooling models were developed based on thermal behavior and geological–geotechnical data. The results indicated that lava cooling durations in some areas exceed practical waiting times for commencing reconstruction. This led to the development of geological engineering solutions that permit road construction and urban and agricultural reconstruction to begin sooner than estimated by the cooling models. On the other hand, potential hazards arising from the eruption process have also been taken into account. Stability analyses of the 200 m high volcanic cone formed during the eruption indicate the possibility of failure in the event of heavy rain and consequently lahar hazards. The results of the investigations carried out and their applications to post-disaster reconstruction may be useful for other volcanic regions, contributing to minimizing risk to infrastructure and urban settlements. Full article
Show Figures

Figure 1

Figure 1
<p>La Palma Island and lava flow field location.</p>
Full article ">Figure 2
<p>Aerial views before and after the 2021 volcanic eruption (Source: Google Earth).</p>
Full article ">Figure 3
<p>Type of lava flows’ distribution from the La Palma 2021 eruption.</p>
Full article ">Figure 4
<p>Aa lava flow section showing upper and lower scoriaceous crusts (dark brown) surrounding dense aa core (gray). The black point inside the circle indicates the site where the photo was taken on the lava field (see <a href="#geohazards-05-00052-f003" class="html-fig">Figure 3</a>).</p>
Full article ">Figure 5
<p>Pahoehoe lava flows showing cooling ropes from the La Palma 2021 eruption. The black point inside the circle indicates the site where the photo was taken on the lava field (see <a href="#geohazards-05-00052-f003" class="html-fig">Figure 3</a>).</p>
Full article ">Figure 6
<p>Thickness map of lava flows and borehole’s location.</p>
Full article ">Figure 7
<p>Borehole drilling in hot lavas, August 2022 (<b>left</b>), and Georadar survey to detect cavities, October 2022 (<b>right</b>).</p>
Full article ">Figure 8
<p>Geophysical cross section of the volcanic cone slope obtained through seismic spectral analysis of surface waves techniques (SASW).</p>
Full article ">Figure 9
<p>Lava flow section using ground-penetration radar techniques to identify cavities and volcanic caves in pahoehoe lavas, depicted in blue. The red color corresponds to materials with high electromagnetic conductivity.</p>
Full article ">Figure 10
<p>Temperatures measured inside the boreholes at different depths between 2 August 2022 and 4 July 2023.</p>
Full article ">Figure 11
<p>Fitting curve obtained from Expression (9) as a function of depth and cooling coefficient (<a href="#geohazards-05-00052-t002" class="html-table">Table 2</a>) in aa lava flows for homogeneous conditions. The dashed lines indicate the margin of error.</p>
Full article ">Figure 12
<p>Cooling curves under homogeneous conditions as a function of depth in aa lava flows. ☆: initial cooling temperature. The dashed line marks a temperature of 50 °C.</p>
Full article ">Figure 13
<p>Fit curve obtained from Expression (19), as a function of the cooling factor and cooling coefficient (<a href="#geohazards-05-00052-t002" class="html-table">Table 2</a>) for heterogeneous conditions. The dashed lines indicate the margin of error. <b>X</b>: theoretical data.</p>
Full article ">Figure 14
<p>Cooling curves at different depths for heterogeneous conditions in aa lava flows, borehole S1. ☆: initial cooling temperature. The dashed line marks a temperature of 50 °C.</p>
Full article ">Figure 15
<p>Cooling curves at different depths for heterogeneous conditions in aa lava flows, borehole S3. ☆: initial cooling temperature. The dashed line marks a temperature of 50 °C.</p>
Full article ">Figure 16
<p>Cooling coefficients and cooling curves calculation flowchart.</p>
Full article ">Figure 17
<p>Mean values of lava flows temperatures obtained 2 years after the volcanic eruption at different distances from the roadsides.</p>
Full article ">Figure 18
<p>Map of surface lava flow temperatures measured in situ by drones in the vicinity of a new road built two years after the eruption. The area is dominated by pahoehoe-type lava flows.</p>
Full article ">Figure 19
<p>Thermal and geomechanical investigations applied to the reconstruction and territorial recovery of La Palma.</p>
Full article ">Figure 20
<p>Temperature distribution as a function of the thickness of aa and pahoehoe lava flows. The solid line represents a quadratic adjustment of the data, illustrating the overall trend. The dashed lines indicate the error margins, providing a visual representation of the uncertainty in the data fitting.</p>
Full article ">Figure 21
<p>(<b>A</b>) Thickness map of lava flows for agricultural recovery purposes. (<b>B</b>) Map of surface temperature measured during 2023. (<b>C</b>,<b>D</b>). Maps of surface temperature estimated for 2024 and 2025 respectively according to cooling rate estimations.</p>
Full article ">Figure 22
<p>Road construction on hot lava flows 6 months after volcanic eruption of La Palma 2021, according to the proposed solution (<b>left</b>). Road excavation of lava flows and related hazard warning signs for high temperatures and gases at a site where the proposed solution was not implemented (<b>right</b>). The black points inside the circle indicate the sites where the photos were taken on the lava field (see <a href="#geohazards-05-00052-f003" class="html-fig">Figure 3</a>).</p>
Full article ">Figure 23
<p>Cooling processes of the lava flows.</p>
Full article ">Figure 24
<p>Measuring gases in a basement located in Puerto Naos, La Palma (<b>left</b>). Accumulated ash deposit near the volcanic cone (<b>right</b>).</p>
Full article ">Figure A1
<p>Representation of the correlation matrix between λ and CF. ☆: maximum correlation value.</p>
Full article ">
19 pages, 5225 KiB  
Article
Seismic Response of a Cable-Stayed Bridge with Concrete-Filled Steel Tube (CFST) Pylons Equipped with the Seesaw System
by Panagiota Katsimpini, George Papagiannopoulos and George Hatzigeorgiou
GeoHazards 2024, 5(4), 1074-1092; https://doi.org/10.3390/geohazards5040051 - 4 Oct 2024
Cited by 1 | Viewed by 1235
Abstract
This research examines the seismic behavior of a cable-stayed bridge featuring concrete-filled steel tube (CFST) pylons, which includes the seesaw system. The objective of the study is to assess the efficacy of the seesaw system in mitigating the seismic response of the bridge [...] Read more.
This research examines the seismic behavior of a cable-stayed bridge featuring concrete-filled steel tube (CFST) pylons, which includes the seesaw system. The objective of the study is to assess the efficacy of the seesaw system in mitigating the seismic response of the bridge across various earthquake scenarios, while also accounting for the implications of soil–structure interaction (SSI). A comprehensive finite element model of the bridge is constructed, incorporating the CFST pylons, cable system, and the novel seesaw energy dissipation system. This model is tested against a range of ground motions that reflect different seismic hazard levels and characteristics. The impact of SSI is analyzed through a series of parametric studies that explore various soil conditions and foundation types. The findings indicate that the implementation of the seesaw system markedly decreases the seismic demands placed on the bridge structure, particularly regarding deck displacements, pylon base shear, and cable forces. Furthermore, the study underscores the significant influence of SSI on the dynamic behavior of the bridge system, emphasizing the necessity of its inclusion in seismic design and analysis. This research enhances the understanding of seismic protection strategies for cable-stayed bridges, providing valuable insights into the advantages of integrating energy dissipation systems and recognizing the importance of SSI effects in evaluating seismic performance. Full article
Show Figures

Figure 1

Figure 1
<p>The seesaw system.</p>
Full article ">Figure 2
<p>View of the bridge [<a href="#B37-geohazards-05-00051" class="html-bibr">37</a>].</p>
Full article ">Figure 3
<p>View of the bridge equipped with the seesaw system.</p>
Full article ">Figure 4
<p>Three-dimensional view of the bridge.</p>
Full article ">Figure 5
<p>Three-dimensional view of the bridge equipped with the seesaw system.</p>
Full article ">Figure 6
<p>Three-dimensional view of the bridge considering SSI.</p>
Full article ">Figure 7
<p>Three-dimensional view of the bridge equipped with the seesaw system considering SSI.</p>
Full article ">Figure 8
<p>Base shear of the tower of the bridge.</p>
Full article ">Figure 9
<p>Base moment of the tower of the bridge.</p>
Full article ">Figure 10
<p>Peak acceleration of a joint of the deck of the bridge.</p>
Full article ">Figure 11
<p>Maximum displacement of a joint of the deck of the bridge.</p>
Full article ">Figure 12
<p>Time history of displacement of a joint on the deck of the bridge with and without the seesaw system (stiff soil).</p>
Full article ">Figure 13
<p>Time history of displacement of a joint on the deck of the bridge with and without the seesaw system (compliant soil).</p>
Full article ">Figure 14
<p>Time history of displacement of a joint on the deck of the bridge based on fixed and on compliant soil (without the seesaw system).</p>
Full article ">Figure 15
<p>Time history of displacement of a joint on the deck of the bridge based on fixed and on compliant soil (with the seesaw system).</p>
Full article ">Figure 16
<p>Ductility demand and displacement demand with and without the seesaw system.</p>
Full article ">
34 pages, 4983 KiB  
Article
GIS-Based Risk Assessment of Building Vulnerability in Flood Zones of Naic, Cavite, Philippines Using AHP and TOPSIS
by Shashi Rani Singh, Ehsan Harirchian, Cris Edward F. Monjardin and Tom Lahmer
GeoHazards 2024, 5(4), 1040-1073; https://doi.org/10.3390/geohazards5040050 - 2 Oct 2024
Viewed by 2528
Abstract
Floods pose significant challenges globally, particularly in coastal regions like the Philippines, which are vulnerable to typhoons and subsequent inundations. This study focuses on Naic city in Cavite, Philippines, using Geographic Information Systems (GIS) to develop flood risk maps employing two Multi-Criteria Decision-Making [...] Read more.
Floods pose significant challenges globally, particularly in coastal regions like the Philippines, which are vulnerable to typhoons and subsequent inundations. This study focuses on Naic city in Cavite, Philippines, using Geographic Information Systems (GIS) to develop flood risk maps employing two Multi-Criteria Decision-Making (MCDM) methods including Analytical Hierarchy Process (AHP) and Technique for Order Preference by Similarity to Ideal Solution (TOPSIS). These maps integrate hazard, vulnerability, and exposure assessments to identify structures most vulnerable to flooding. Weight assignments in the study are derived from a literature review and expert opinions, reflecting the Philippines’ flood-prone geography and historical data. Structural attributes, categorized on a low to very high scale, were assessed based on field survey data from 555 buildings. AHP categorized 91.3% of buildings as moderate to very high risk, whereas TOPSIS placed 68% in this category, underscoring methodological disparities in data handling and assumptions. This research enhances understanding of flood threats and offers a decision-making framework for resilient flood risk management strategies. Identifying vulnerable buildings aims to support informed urban planning and disaster preparedness in flood-prone areas, thereby mitigating potential property, infrastructure, and livelihood damage. Full article
Show Figures

Figure 1

Figure 1
<p>Study Area in Naic with its surveyed buildings.</p>
Full article ">Figure 2
<p>Map of used LiDAR Data for Naic Floodplain.</p>
Full article ">Figure 3
<p>Location and delineated watershed of the Labac river basin relative to the Philippines.</p>
Full article ">Figure 4
<p>Procedures to model flood hazard.</p>
Full article ">Figure 5
<p>Methodological framework of study in building risk assessment.</p>
Full article ">Figure 6
<p>Generated maps from ArcMap in building flood hazard parameters.</p>
Full article ">Figure 6 Cont.
<p>Generated maps from ArcMap in building flood hazard parameters.</p>
Full article ">Figure 6 Cont.
<p>Generated maps from ArcMap in building flood hazard parameters.</p>
Full article ">Figure 7
<p>Generated maps from ArcMap in building flood vulnerability parameters.</p>
Full article ">Figure 7 Cont.
<p>Generated maps from ArcMap in building flood vulnerability parameters.</p>
Full article ">Figure 7 Cont.
<p>Generated maps from ArcMap in building flood vulnerability parameters.</p>
Full article ">Figure 7 Cont.
<p>Generated maps from ArcMap in building flood vulnerability parameters.</p>
Full article ">Figure 7 Cont.
<p>Generated maps from ArcMap in building flood vulnerability parameters.</p>
Full article ">Figure 7 Cont.
<p>Generated maps from ArcMap in building flood vulnerability parameters.</p>
Full article ">Figure 8
<p>Result maps for building flood exposure parameters using ArcMap: (<b>a</b>) building density, (<b>b</b>) number of buildings per grid, and (<b>c</b>) type of building use.</p>
Full article ">Figure 8 Cont.
<p>Result maps for building flood exposure parameters using ArcMap: (<b>a</b>) building density, (<b>b</b>) number of buildings per grid, and (<b>c</b>) type of building use.</p>
Full article ">Figure 9
<p>Generated maps from the different parameters in using AHP: (<b>a</b>) Building Flood Hazard Map, (<b>b</b>) Building Flood Vulnerability Map, and (<b>c</b>) Building Flood Exposure Map.</p>
Full article ">Figure 9 Cont.
<p>Generated maps from the different parameters in using AHP: (<b>a</b>) Building Flood Hazard Map, (<b>b</b>) Building Flood Vulnerability Map, and (<b>c</b>) Building Flood Exposure Map.</p>
Full article ">Figure 10
<p>Generated maps from the different parameters in using TOPSIS: (<b>a</b>) Building Flood Hazard Map, (<b>b</b>) Building Flood Vulnerability Map, and (<b>c</b>) Building Flood Exposure Map.</p>
Full article ">Figure 10 Cont.
<p>Generated maps from the different parameters in using TOPSIS: (<b>a</b>) Building Flood Hazard Map, (<b>b</b>) Building Flood Vulnerability Map, and (<b>c</b>) Building Flood Exposure Map.</p>
Full article ">Figure 11
<p>Building Flood Risk Map Using AHP.</p>
Full article ">Figure 12
<p>Building Flood Risk Map Using TOPSIS.</p>
Full article ">Figure A1
<p>Study Area Naic in Philippines.</p>
Full article ">
22 pages, 15918 KiB  
Article
Exceptional Cluster of Simultaneous Shallow Landslides in Rwanda: Context, Triggering Factors, and Potential Warnings
by Fils-Vainqueur Byiringiro, Marc Jolivet, Olivier Dauteuil, Damien Arvor and Christine Hitimana Niyotwambaza
GeoHazards 2024, 5(4), 1018-1039; https://doi.org/10.3390/geohazards5040049 - 25 Sep 2024
Cited by 1 | Viewed by 1116
Abstract
Rwanda, in eastern tropical Africa, is a small, densely populated country where climatic disasters are often the cause of considerable damage and deaths. Landslides are among the most frequent hazards, linked to the country’s peculiar configuration including high relief with steep slopes, humid [...] Read more.
Rwanda, in eastern tropical Africa, is a small, densely populated country where climatic disasters are often the cause of considerable damage and deaths. Landslides are among the most frequent hazards, linked to the country’s peculiar configuration including high relief with steep slopes, humid tropical climate with heavy rainfall, intense deforestation over the past 60 years, and extensive use of the soil for agriculture. The Karongi region, in the west-central part of the country, was affected by an exceptional cluster of more than 700 landslides during a single night (6–7 May 2018) over an area of 100 km2. We analyse the causes of this spectacular event based on field geological and geomorphology investigation and CHIRPS and ERA5-Land climate data. We demonstrate that (1) the notably steep slopes favoured soil instability; (2) the layered soil and especially the gravelly, porous C horizon allowed water storage and served as a detachment level for the landslides; (3) relatively low intensity, almost continuous rainfall over the previous two months lead to soil water-logging; and (4) acoustic waves from thunder or mechanical shaking by strong wind destabilized the water-logged soil through thixotropy triggering the landslides. This analysis should serve as a guide for forecasting landslide-triggering conditions in Rwanda. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Topographic map of Eastern Africa (NASA GTOPO30 DEM) showing the two branches of the rift (the Western Rift and the Eastern Rift) fringing the Lake Victoria high plateau. The white rectangle corresponds to <a href="#geohazards-05-00049-f001" class="html-fig">Figure 1</a>b. (<b>b</b>) Rwanda elevation map based on the Copernicus 30DEM showing the high topography on both sides of Lake Kivu. The eastern rift shoulder (the Congo–Nile ridge) locally reaches 3000 m in altitude (<a href="#geohazards-05-00049-f001" class="html-fig">Figure 1</a>c) and is strongly affected by landslides. The Karongi area discussed in this work is indicated by the white square. The grey rectangle corresponds to the topographic data illustrated in <a href="#geohazards-05-00049-f001" class="html-fig">Figure 1</a>c. (<b>c</b>) Topographic swap (profiles are sampled every 20 m and separated by 1 km) across the Lake Kivu rift system showing the high altitudes of the Congo–Nile ridge in Rwanda and the progressive eastward flattening towards eastern lowlands of Rwanda and the Lake Victoria plateau. The black line represents the mean value and the grey shaded area indicate the dispersion of the values. (<b>d</b>) Mean annual rainfall pattern calculated using CHIRPS [<a href="#B39-geohazards-05-00049" class="html-bibr">39</a>] annual data from 1981 to 2022. Note the strong contrast between the relatively low rainfall (&lt;1000 mm/yr) in eastern Rwanda and the high rainfall (&gt;2000 mm/yr) in the eastern Congo Basin to the west. Mean rainfall on the Congo–Nile ridge ranges from 1250 to 1500 mm/yr.</p>
Full article ">Figure 2
<p>Mud–debris flows in the Karongi area. Pictures of the aftermath of landslides that occurred on the 7th of May 2018 [<a href="#B38-geohazards-05-00049" class="html-bibr">38</a>]. The images are located in Figure 3. <a href="#geohazards-05-00049-f002" class="html-fig">Figure 2</a>a is one of the numerous landslides that occurred along the steep Rwankuba crest and <a href="#geohazards-05-00049-f002" class="html-fig">Figure 2</a>b corresponds to the “Major landslide” northwest of Burega School. Note that the latest is also labelled on Figure 4.</p>
Full article ">Figure 3
<p>Geological map of the Karongi area adapted from [<a href="#B42-geohazards-05-00049" class="html-bibr">42</a>,<a href="#B46-geohazards-05-00049" class="html-bibr">46</a>], incorporating field observations and measurements of lithological and tectonic structures. The black rectangle corresponds to the area studied in more detail in this work.</p>
Full article ">Figure 4
<p>Topography of the Karongi area from COPERNICUS DEM [<a href="#B47-geohazards-05-00049" class="html-bibr">47</a>]. The topographic profile was constructed by sampling every 20 m along profiles separated by 100 m. The black line indicates the mean topography, while the grey area represents the dispersion of the values. The major landslide indicated is that shown in <a href="#geohazards-05-00049-f002" class="html-fig">Figure 2</a>b.</p>
Full article ">Figure 5
<p>Land use (based on the Google Earth satellite image available for the 1-January-2023) and occurrence of the landslides in the Karongi area based on satellite image analyses.</p>
Full article ">Figure 6
<p>October 2018 high-resolution CNES/Airbus image (via Google Earth) showing the major landslide and adjacent ones (green line contours) in the Karongi area developing on graphitic schists and meta-sandstone bedrock. Outside the landslides, the angular polygons of green and light brown colours are cultivated fields, and the dark green areas are trees (generally eucalyptus). See <a href="#geohazards-05-00049-f002" class="html-fig">Figure 2</a>b for an image of the landslide immediately after being triggered.</p>
Full article ">Figure 7
<p>Interpolated map of soil thickness in the Karongi area created by the Inverse Distance Weighted (IDW) method. Black dots and associated numbers show the location of the data points and their associated soil thicknesses (rounded to the upper 0.5 m value). As indicated in <a href="#sec3dot3-geohazards-05-00049" class="html-sec">Section 3.3</a>, thickness was measured on exposed complete soil profiles using a tape measure.</p>
Full article ">Figure 8
<p>Main soil characteristics of the Karongi area. (<b>a</b>)—General view of a soil profile and corresponding idealized section indicating the different horizons with their range of thicknesses. The letters a to g are related to the pictures. (<b>b</b>)—Undifferentiated E and B horizons developing from a bedrock composed of meta-sandstone. (<b>c</b>)—Large-scale view of the quartz gravel layer forming the C horizon on meta-sandstone bedrock. (<b>d</b>)—Close view of the C horizon. Note that the quartz gravels and pebbles are poorly rounded. (<b>e</b>)—Soil profile developing on graphitic schist bedrock. (<b>f</b>)—Large-scale view of the soil-barren meta-sandstone/quartzite basement (R horizon). (<b>g</b>)—Large-scale view of the soil-barren graphitic schist bedrock (R horizon).</p>
Full article ">Figure 9
<p>Landscape (<b>a</b>) and cross-section (<b>b</b>) of the main landslide in the Karongi area (see <a href="#geohazards-05-00049-f006" class="html-fig">Figure 6</a> for localization). The pictures illustrate the main elements and are reported in the section and general picture. (<b>c</b>)—Detachment zone of the landslide in the graphitic schists. Note the very steep slope (about 60°). The surface is covered with pebbles and blocks of schists from the underlying basement. (<b>d</b>)—Soil section at the top of the landslide showing dark strongly weathered graphitic schists (R horizon) overlaid by a thin layer of quartz pebbles (C horizon) that separate the R horizon from a very condensed E, B, and A horizon stack. (<b>e</b>)—Similar soil profile but developed in non-graphitic schists and sandstones. (<b>f</b>)—Lower part of the detachment zone: the break in topographic slope is sharp, and a large amount of colluvial material is deposited in that area, affected by open cracks. (<b>g</b>)—The thick colluvial layer collapsing downslope as metre-thick coherent rafts of clay–sandy material with few blocks. (<b>h</b>)—Blocky material of an older landslide that occurred on the SE slope of the system in a sandstone-rich basement unit. Note that for the landscape pictures (<b>c</b>,<b>f</b>–<b>h</b>), the scale varies with distance.</p>
Full article ">Figure 10
<p>Slope map of the Karongi area with the position of the landslides (yellow polygons). The histogram represents the density of landslides versus the landslide initiation angle.</p>
Full article ">Figure 11
<p>Rainfall in the Karongi area calculated from CHIRPS data. (<b>a</b>): Yearly rainfall from 1981 to 2022. The red line indicates 2018. (<b>b</b>): Monthly rainfall over the 42 years. The grey envelope includes the dispersion of the data. The blue line is the mean value for each month. The red line corresponds to 2018. (<b>c</b>): Calculated March–April Rainfall Anomaly Index (RAI) for the period 1981–2022. (<b>d</b>): Total number of rainy days in March–April over the 42 years. The red line indicates 2018. (<b>e</b>): Daily rainfall from 1 March to 6 May 2018.</p>
Full article ">Figure 12
<p>Detection of stormy events in the Karongi area (white rectangle on <a href="#geohazards-05-00049-f012" class="html-fig">Figure 12</a>a,b refers to the area covered by the geomorphology study) using the ERA5-Land reanalysis data. (<b>a</b>) Example of an hourly precipitation map showing a major storm cell positioned in SW Rwanda (the grey line indicates the border of the country). (<b>b</b>) Wind speed (red colours) and direction (black arrows) associated with the same cell. (<b>c</b>) Hourly precipitation (top) and windspeed (bottom) between the 1st of March and the 6th of May 2018. Numbers indicate the day of major events. Events including both exceptional rain and wind are considered to represent storms.</p>
Full article ">Figure 13
<p>Proposed model for the formation of the Karongi landslides. (<b>a</b>) The soil rests at equilibrium on the slope, limited from the basement by the gravel layer of the C horizon (note that for better representation, the thickness of the C layer as well as the size of the pebbles is over-exaggerated. φ<sub>1</sub>, C<sub>1</sub>, p<sub>e1</sub>, and φ<sub>2</sub>, C<sub>2</sub>, p<sub>e2</sub> are the internal frictional angle, cohesion, and interstitial pressure, respectively, in the clay-rich soil and gravely C horizon. q is the surface slope angle, Z<sub>p</sub> is the water table level, Z<sub>r</sub> and Z<sub>r’</sub> are the limits of the rupture layer, and τ’<sub>N</sub> is the normal effective constrain. The light blue arrows indicate water infiltration from the surface and preferential circulation within the C horizon. The brown layer represents the non-saturated part of the E + B horizon, and the blueish layer represents the progressively saturated part of that horizon with a higher water content at the base and in the C horizon. (<b>b</b>) Because of continuous rainfall, the water table level rises with time. When saturation reaches a near threshold, vibrations linked to sound waves during a thunderstorm event induce a rise in the interstitial fluid pressure p<sub>1</sub> and especially p<sub>2</sub> that become larger than the respective equilibrium pressures p<sub>e1</sub> and p<sub>e2</sub>, initiating soil movement and triggering thixotropy in the clay-rich layer. (<b>c</b>) Final situation after the landslide.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop