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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (18)

Search Parameters:
Keywords = surficial geology

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 13502 KiB  
Article
Assessment of Radon and Naturally Occurring Radionuclides in the Vredefort Meteorite Crater in South Africa
by Rikus Le Roux and Jacques Bezuidenhout
Atmosphere 2023, 14(12), 1826; https://doi.org/10.3390/atmos14121826 - 15 Dec 2023
Viewed by 1334
Abstract
The concentric impact rings of the Vredefort Crater contain rocks with elevated uranium concentrations resulting from the geological signature of a meteoric impact. The decay of this uranium was estimated to lead to elevated indoor radon concentrations in the Crater, but such a [...] Read more.
The concentric impact rings of the Vredefort Crater contain rocks with elevated uranium concentrations resulting from the geological signature of a meteoric impact. The decay of this uranium was estimated to lead to elevated indoor radon concentrations in the Crater, but such a study has never been carried out. This study explores the relationship between the natural radionuclides found in the geology of the Vredefort Crater and indoor radon concentrations. This was achieved through soil sampling and radionuclide surveys conducted on three impact rings, supplemented by indoor radon measurements in dwellings found in the area. In situ measurements revealed that one impact ring had higher-than-average uranium concentrations at 50 Bq/kg. Surprisingly, the measured indoor radon levels were lower than expected (113 Bq/m3). These measurements were taken during the COVID-19 pandemic and colder months, conditions that would typically result in elevated indoor radon levels. Soil samples indicated uranium activity of 30 Bq/kg, comparable to the world average of 35 Bq/kg. However, defunct mine tunnels in the area exhibited elevated radon concentrations, averaging 364 Bq/m3. The disparity between expected and measured indoor radon levels was attributed to the composition of surficial deposits, bedrock, and architectural features of the dwellings preventing radon accumulation. Full article
(This article belongs to the Special Issue Atmospheric Radon Concentration Monitoring and Measurements)
Show Figures

Figure 1

Figure 1
<p>A photo of two hills from different geological units (Dome rings) that were measured during the in situ survey.</p>
Full article ">Figure 2
<p>A geological map of the Vredefort meteor impact crater. The town of Vredefort is shown as a black dot and the study area is indicated by a red rectangle. (Modified from the RSA geological map 1997 by the Council of Geoscience of South Africa, obtained from <a href="https://www.geoscience.org.za" target="_blank">https://www.geoscience.org.za</a>, assessed on 1 October 2023).</p>
Full article ">Figure 3
<p>A Google Earth image showing the location of the four guest farms, indicated by the numbered regions, where indoor radon measurements were carried out.</p>
Full article ">Figure 4
<p>A geological map of the northeastern section of the Vredefort impact crater with a color-graded overlay that shows the in situ and on-foot measured uranium concentrations (Bq/kg). The red square demarcates the part of the survey that was done on foot (Modified from the RSA geological map, 1997, by the Council of Geoscience of South Africa, obtained from <a href="https://www.geoscience.org.za" target="_blank">https://www.geoscience.org.za</a>, assessed on 1 October 2023).</p>
Full article ">Figure 5
<p>A Google Earth image of the northeastern section of the Vredefort impact crater with a color-graded overlay that displays measured thorium concentrations (Bq/kg) with the borders of geological units, which are indicated as white lines.</p>
Full article ">Figure 6
<p>A Google Earth image showing measurements that were done on foot covering two geological units of the Witwatersrand Supergroup. A color-graded overlay displays the measured uranium concentrations (Bq/kg) and the borders of geological units are indicated as white lines. This image presents an excerpt of the measurements shown in <a href="#atmosphere-14-01826-f004" class="html-fig">Figure 4</a>, illustrating the characteristics of the terrain.</p>
Full article ">Figure 7
<p>An elevation graph with an overlay of the uranium concentrations of the measurements that were done on foot. The uranium activity (Bq/kg) is scaled by a factor of 10 for visual purposes.</p>
Full article ">Figure 8
<p>Google terrain map showing the location and indoor radon concentrations measured on the four farms, indicated by the numbered regions, in the Vredefort Dome.</p>
Full article ">Figure 9
<p>Histogram of the indoor radon concentrations measured on four guest farms in the Vredefort Dome.</p>
Full article ">Figure 10
<p>A Google terrain map showing the locations where the ground samples were taken, along with their uranium activity. The locations of the mines and the measured radon concentration are also shown. The inlay shows an enlarged view of the mining area.</p>
Full article ">Figure 11
<p>The natural logarithm of the indoor radon concentrations measured in the Vredefort Dome (indicated by the bar graph) compared to a normalized normal distribution with a mean of 4.6 and a standard deviation of 0.49 (indicated by the solid line).</p>
Full article ">
30 pages, 25207 KiB  
Article
Satellite Imagery for Rapid Detection of Liquefaction Surface Manifestations: The Case Study of Türkiye–Syria 2023 Earthquakes
by Maria Taftsoglou, Sotiris Valkaniotis, George Papathanassiou and Efstratios Karantanellis
Remote Sens. 2023, 15(17), 4190; https://doi.org/10.3390/rs15174190 - 25 Aug 2023
Cited by 23 | Viewed by 7747
Abstract
The 6 February 2023 earthquake doublet (Mw 7.7 and Mw 7.6) that occurred on the East Anatolian Fault Zone (EAFZ) triggered a significant amount of soil liquefaction phenomena in SE Türkiye and NW Syria. The great areal extent of the affected area and [...] Read more.
The 6 February 2023 earthquake doublet (Mw 7.7 and Mw 7.6) that occurred on the East Anatolian Fault Zone (EAFZ) triggered a significant amount of soil liquefaction phenomena in SE Türkiye and NW Syria. The great areal extent of the affected area and the necessity of rapid response led to the adoption and improvement of a workflow for mapping liquefaction phenomena based on remote sensing data. Using satellite imagery, we identified 1850 sites with liquefaction manifestation and lateral spreading deformation. We acquired a thorough map of earthquake-triggered liquefaction based on visual mapping with optical satellite imagery (high and very high-resolution) and the aid of radar satellite imagery and interferometry. The majority of sites are found along meandering sections of river valleys, coastal plains, drained lakes, swamps, and lacustrine basins along the East Anatolian Fault, highlighting once again the influence of geomorphology/surficial geology on the distribution of liquefaction phenomena. A total of 95% of the liquefaction occurrences were mapped within 25 km from the surface trace of the fault, confirming the distance from fault rupture as a more effective tool for predicting the distribution of liquefaction than epicentral distance. Thus, taking into consideration the rapid documentation of these phenomena without the limitations in terms of time, cost, and accessibility of the field investigation techniques, this desktop-based approach can result in a rapid and comprehensive map of liquefaction from a strong earthquake, and can also be used as a future guide for subsequent field investigations for liquefaction hazard mapping. Full article
Show Figures

Figure 1

Figure 1
<p>Aftershocks for the M7.7 and M7.6 earthquakes. Epicenter locations derived from AFAD catalog (6 February 2023–2 May 2023).</p>
Full article ">Figure 2
<p>Active fault map of Eastern Anatolia showing the East Anatolia Fault segments (black lines, from [<a href="#B54-remotesensing-15-04190" class="html-bibr">54</a>]). Holocene and Quaternary deposits from [<a href="#B63-remotesensing-15-04190" class="html-bibr">63</a>]. Inset box at lower right shows the regional tectonic setting modified after [<a href="#B64-remotesensing-15-04190" class="html-bibr">64</a>]. KFS: Karliova Fault segment, IFS: Ilica Fault segment, PaFS: Palu Fault segment, DSFZ: Dead Sea Fault Zone, CA: Cyprus Arc. PFS: Pütürge Fault segment, EFS: Erkenek Fault segment, PAFS: Pazarcik Fault segments, AFS: Amanos Fault segments, NFZ: Narli Fault zone, YF: Yesemek Fault, SUFS: Sürgü Fault segment, CAFS: Çardak Fault segment, SFS: Savrun Fault segment, EFZ: Enginek Fault zone, TFS: Toprakkale Fault segment, KFS: Karataş Fault segment, MFZ: Maraş Fault zone, DIFZ: Düziçi–İskenderun Fault zone.</p>
Full article ">Figure 3
<p>Comparison of liquefaction features in different optical sensors and resolutions. From left to right; (<b>Sentinel-2</b>) (10 m), (<b>Planetscope</b>) (3–4 m), (<b>Maxar WorldView-02</b>) (0.46 m).</p>
Full article ">Figure 4
<p>Analytical flow chart of steps for identification and mapping of liquefaction-related phenomena using remote sensing data.</p>
Full article ">Figure 5
<p>Co-seismic phase interferogram (Sentinel-1 ascending track, 28 January 2023–9 February 2023) over Amik Plain, Hatay Province; irregular and isolated fringe patterns in the eastern part of the basin away from the main fault deformation (western part) correspond to spots with significant liquefaction deformation such as subsidence, lateral spreading, or ground oscillation across the former Amik Lake floor (dotted line).</p>
Full article ">Figure 6
<p>Interferometric coherence over Amik Plain, Hatay Province; pre-earthquake pair (<b>a</b>) in comparison with co-seismic pair (<b>b</b>). Dark areas mark significant coherence loss. Coherence change is shown in (<b>c</b>). Concentrations of coherence loss coincide with areas with significant liquefaction manifestations (<b>d</b>) like sand blows, lateral spreading, and ground oscillation across the former Amik Lake floor.</p>
Full article ">Figure 7
<p>Examples of different types of earthquake-induced liquefaction phenomena identified and mapped using optical imagery.</p>
Full article ">Figure 8
<p>Overview map of liquefaction and lateral spreading sites identified and mapped using satellite imagery. Fault rupture is shown with black lines, as mapped from Sentinel-2 imagery and [<a href="#B52-remotesensing-15-04190" class="html-bibr">52</a>]. Epicenters of the February 6th seismic events are marked with yellow stars (AFAD). Quaternary formations along the fault rupture from [<a href="#B63-remotesensing-15-04190" class="html-bibr">63</a>]. Inset maps 1–3 are described in <a href="#remotesensing-15-04190-f009" class="html-fig">Figure 9</a>.</p>
Full article ">Figure 9
<p>Detailed inset maps (<b>1</b>–<b>3</b>) along the East Anatolian Fault Zone, where liquefaction and lateral spreading phenomena were identified for the 6 February 2023 earthquakes. Fault rupture is shown with black lines, as mapped from Sentinel-2 imagery [<a href="#B52-remotesensing-15-04190" class="html-bibr">52</a>]. Epicenters of the 6 February seismic events are marked with yellow stars and orange stars for the 20 February M6.4 aftershock near Antakya (AFAD). Quaternary formations (yellow) along the fault rupture from [<a href="#B63-remotesensing-15-04190" class="html-bibr">63</a>]. Location of maps (<b>1</b>–<b>3</b>) is shown in previous <a href="#remotesensing-15-04190-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 10
<p>(<b>a</b>) Relative elevation model (REM) of Orontes River Valley, south of the former Amik Lake. Liquefaction sites are marked with red dots. Elevation source: Copernicus DEM. (<b>b</b>) Detection of liquefaction ejecta material (light colors) and lateral spreading phenomena along a meandering section of Orontes River in Maxar VHR satellite imagery.</p>
Full article ">Figure 11
<p>(<b>Above</b>) KH-4 Corona declassified satellite imagery from 1969, showing the extent of landfill and coastal front expansion during the past decades (current shoreline with white line), (<b>Bottom</b>): Post-earthquake VHR optical satellite imagery by Maxar showing the current status in Iskenderun. Liquefaction-related phenomena are marked with red.</p>
Full article ">Figure 12
<p>Pre-earthquake (<b>a</b>,<b>c</b>) and post-earthquake VHR satellite imagery (Maxar) showing liquefaction ejecta on the piers of Iskenderun port (<b>b</b>) and submersion of specific sections due to lateral spreading (<b>d</b>).</p>
Full article ">Figure 13
<p>(<b>a</b>) Distribution of liquefaction phenomena along Gölbasi, Azapli, and Inekli Lakes in the Gölbasi Basin (red markers) with the majority found along the 6 February fault rupture and around the lakes/marshes. (<b>b</b>) Lateral spreading deformation and cracks (yellow lines) along the coast of Gölbasi Lake and liquefaction phenomena in the urban area of Gölbasi City. (<b>c</b>) Post-earthquake VHR optical satellite imagery by Maxar depicting in detail a submerged section of Gölbasi Lake coast due to lateral spreading.</p>
Full article ">Figure 14
<p>Pre-earthquake satellite imagery (<b>a</b>,<b>c</b>) and post-earthquake satellite imagery (Maxar) showing the lateral spreading phenomena (<b>b</b>) and the subsidence of specific areas in the coastline zone of Gölbasi Lake (<b>d</b>).</p>
Full article ">Figure 15
<p>Distribution of liquefaction sites in comparison (<b>a</b>) with their distance from the epicenter of the 6 February 2023 M7.7 earthquake (yellow star) and (<b>b</b>) the fault rupture (black line).</p>
Full article ">Figure 16
<p>Cumulative frequency diagram of distance from the fault rupture for the liquefaction sites.</p>
Full article ">Figure 17
<p>Distribution of liquefaction sites in comparison (<b>a</b>) with peak ground acceleration ShakeMap and (<b>b</b>) peak ground velocity ShakeMap. Cumulative frequency diagrams for PGA (<b>c</b>) and PGV (<b>d</b>). Contour of 0.14 g PGA and contour of 12 cm/s PGV that represent the 95th percentile of liquefaction sites are shown in (<b>a</b>,<b>b</b>) as a dotted line. Yellow stars show the epicenters of the two mainshocks (M7.7 and M7.6) and orange stars the location of major aftershocks (M6.7 and M6.4).</p>
Full article ">Figure 18
<p>Liquefaction and lateral spreading sites (red dots) in Aksu River bend valleys area, near the epicenter of the Mw 7.7 earthquake (yellow star). Ground cracks and earthquake surface ruptures shown as red lines. Insets show details of liquefaction ejecta over paleo-meanders of Aksu River (Maxar Open Data VHR satellite imagery). USGS near real-time liquefaction hazard map as a reference layer.</p>
Full article ">Figure 19
<p>Topography of Amik (Amuq) Plain. White continuous line (1) marks the 19th–20th century extent of the now-drained Amik Lake, while white dashed line (2) shows the estimated furthest extent of the historical lake coastline. The 6 February 2023 liquefaction and lateral spreading sites shown as orange dots, and earthquake fault rupture with red lines. Contour line interval is 2 m. Elevation source: Copernicus DEM.</p>
Full article ">Figure 20
<p>(<b>a</b>) Late Holocene geomorphology of Amik (Amuq) Plain from [<a href="#B74-remotesensing-15-04190" class="html-bibr">74</a>], with liquefaction and lateral spreading sites (red dots) and surface ruptures/ground cracks (purple lines). (<b>b</b>) Liquefaction and lateral spreading sites (red dots) over Amik (Amuq) Plain over a declassified KH-4 Corona satellite imagery from 1969. Black dotted line marks the estimated furthest extent of historical Amik Lake.</p>
Full article ">Figure 21
<p>Recurrent liquefaction from the 20 February Mw 6.4 strong aftershock in Antakya area. Irregular features (isolated fringe patterns and spotted areas with loss of coherence) that are persistent in both ascending and descending interferograms (<b>a</b>,<b>b</b>) and not related to the larger scale fault deformation patterns (larger fringes in (<b>a</b>,<b>b</b>)) correspond to areas with widespread landslides and slow-moving landslide deformations (northern dashed ellipse) across the hills opposite of Antakya City, and widespread liquefaction and lateral spreading deformation (southern dashed ellipse) across the meandering valley of Orontes River just north of Antakya (<b>c</b>).</p>
Full article ">
18 pages, 7040 KiB  
Article
A Detailed Liquefaction Susceptibility Map of Nestos River Delta, Thrace, Greece Based on Surficial Geology and Geomorphology
by Maria Taftsoglou, Sotirios Valkaniotis, George Papathanassiou, Nikos Klimis and Ioannis Dokas
Geosciences 2022, 12(10), 361; https://doi.org/10.3390/geosciences12100361 - 29 Sep 2022
Cited by 2 | Viewed by 3004
Abstract
The existence of high potential onshore and offshore active faults capable to trigger large earthquakes in the broader area of Thrace, Greece in correlation with the critical infrastructures constructed on the recent and Holocene sediments of Nestos river delta plain, was the motivation [...] Read more.
The existence of high potential onshore and offshore active faults capable to trigger large earthquakes in the broader area of Thrace, Greece in correlation with the critical infrastructures constructed on the recent and Holocene sediments of Nestos river delta plain, was the motivation for this research. The goal of this study is twofold; compilation of a new geomorphological map of the study area and the assessment of the liquefaction susceptibility of the surficial geological units. Liquefaction susceptibility at regional scale is assessed by taking into account information dealing with the depositional environment and age of the surficial geological units. In our case, available geological mapping shows a deficient depiction of Pleistocene and Holocene deposits. Taking into consideration the heterogeneously behavior of active floodplains and deltas in terms of liquefaction, a detailed classification of geological units was mandatory. Using data provided by satellite and aerial imagery, and topographic maps, dated before the 1970’s when extensive modifications and land reclamation occurred in the area, we were able to trace fluvial and coastal geomorphological features like abandoned stream/meanders, estuaries, dunes, lagoons and ox-bow lakes. This geomorphological-oriented approach clearly classified the geological units according to their depositional environment and resulted in a more reliable liquefaction susceptibility map of 4 classes of susceptibility; Low, Moderate, High and Very High. The sediments classified as very high liquefaction susceptibility are related to fluvial landforms, the high to moderate liquefaction susceptibility ones in coastal and floodplain landforms, and low susceptibility in zones of marshes. The sediments classified in the highest group of liquefaction susceptibility cover 85.56 km2 of the study area (16.6%). Particular attention was drawn to critical infrastructure (Kavala International Airport “Alexander the Great”) constructed on the most prone to liquefaction areas. Full article
(This article belongs to the Special Issue Assessment of Earthquake-Induced Soil Liquefaction Hazard)
Show Figures

Figure 1

Figure 1
<p>Geophysical map highlighting the Nestos River zone. The study area is indicated with the red square. (1) Kavala–Xanthi–Komotini fault, (2) Simvolo fault, (3) Drama fault, (4) Maronia fault, (5) Orestias–Mikri Doxipara. Fault zones from [<a href="#B41-geosciences-12-00361" class="html-bibr">41</a>,<a href="#B42-geosciences-12-00361" class="html-bibr">42</a>].</p>
Full article ">Figure 2
<p>Geological map of the Nestos River delta provided by HSGME with (1) floodplain deposits, (2) channel deposits, (3) coastal deposits, (4) swamp deposits, (5) lagoons, (6) Nestos River, (7) Pleistocene deposits, (8) screes.</p>
Full article ">Figure 3
<p>Mapping of the geomorphological features using orthophoto maps from 1945 and KH-4 Corona frames.</p>
Full article ">Figure 4
<p>Geomorphological formations at the coastal zone of Nestos River delta. (<b>a</b>) orthophotomap from 1945, (<b>b</b>) mapping of old meanders, dunes and beach barrier in orthophoto map from 1945, (<b>c</b>) satellite image from 1960, (<b>d</b>) accretion of mouth river and extension of deltaic deposits and dunes.</p>
Full article ">Figure 5
<p>Recent evolution of meandering streams and progression of irrigation and drainage modifications according to (<b>a</b>) the orthophoto map from 1945 derived from Hellenic Cadastre, (<b>b</b>) KH-4 image from 1960 and (<b>c</b>) KH-4 image from 1968, (<b>d</b>) mapped geomorphological fluvial and deltaic formations.</p>
Full article ">Figure 6
<p>The new geomorphological map of the Nestos River floodplain resulting from the combination of orthophoto maps (1945) and Corona KH-4 images (1960, 1968). (1) former riverbed, (2) point-bar, (3) deltaic deposits, (4) current river channel, (5) anthropogenic levee, (6) coastal deposits (dunes), (7) coast, (8) levee, (9) oxbow, (10) floodplain deposits, (11) marsh deposits, (12) saltmarsh deposits, (13) lagoon, (14) beach barrier, (15) Pleistocene deposits.</p>
Full article ">Figure 7
<p>Liquefaction susceptibility map of the Nestos River delta derived from the application of the criteria proposed by Youd and Perkins [<a href="#B25-geosciences-12-00361" class="html-bibr">25</a>] to the geological maps provided by HSGME.</p>
Full article ">Figure 8
<p>Liquefaction susceptibility map of the Nestos River delta, resulting from the application of Youd and Perkins criteria [<a href="#B25-geosciences-12-00361" class="html-bibr">25</a>] to the new geomorphological map, produced by processing of satellite and aerial imagery.</p>
Full article ">Figure 9
<p>Comparison of liquefaction susceptibility maps focusing on the critical infrastructure of Kavala International Airport (KVA) (<b>a</b>) based on the geological map of HSGME, (<b>b</b>) based on the new geomorphological map produced by processing of satellite and aerial imagery.</p>
Full article ">
10 pages, 3545 KiB  
Article
The Investigation of Shallow Structures at the Meishan Fault Zone with Ambient Noise Tomography Using a Dense Array Data
by Wei-Jhe Wu, Chien-Min Su and Chau-Huei Chen
Appl. Sci. 2022, 12(12), 5847; https://doi.org/10.3390/app12125847 - 8 Jun 2022
Cited by 1 | Viewed by 1777
Abstract
Seismic monitoring relies on seismography. However, the high cost of seismic equipment has presented a challenge to increasing the density of seismic networks in previous decades. Due to the large station spacing and inferior coverage of stations, this situation has led to a [...] Read more.
Seismic monitoring relies on seismography. However, the high cost of seismic equipment has presented a challenge to increasing the density of seismic networks in previous decades. Due to the large station spacing and inferior coverage of stations, this situation has led to a loss of detail in many research results. Along with the improvement of technology, the problem of increasing the density of seismographic observations is no longer an impossible issue. This makes it feasible to deploy a dense seismic network for monitoring earthquakes. This study deployed a linear dense array across the Meishan Fault in west-southern Taiwan for the purpose of analyzing the shallow fault zone structure. While the 1906 Meishan earthquake occurred in a period when historic records were available, the surficial geology surveys of the Meishan Fault are challenging because farming and construction engineering have obscured the outcrop. Early surveys of the Meishan Fault were mainly seismic surveys. In recent decades, over thirty profiles have been completed. However, the reflection seismic records had poor signal-to-noise ratios because the Meishan Fault is buried under thick sediments. Thus, the shallow structure of the Meishan Fault is still not known in detail. This study applied double-beamforming tomography to a dense seismic array to obtain high-resolution images of the Meishan Fault zone. The result shows that there is a south-dipping interface near the fault trace as indicated by the Central Geological Survey of Taiwan. In addition, we observed velocity transitions of perturbation profiles that may be caused by a branch fault, the Chentsoliao Fault. This study demonstrates that the ambient noise double beamforming method is an effective tool for imaging the detailed shallow structure along with the dense seismic array. Full article
Show Figures

Figure 1

Figure 1
<p>A topographic map showing the locations of the Meishan Fault, seismic stations, and the National Chung Cheng University (CCU) region. The blue box represents the CCU region, and the red and black lines indicated the fault traces of the Meishan Fault obtained by the CGS and Omori [<a href="#B3-applsci-12-05847" class="html-bibr">3</a>], respectively. The array of the sixty smartsolo geophones is represented by white triangles. The inset map shows the study area (Taiwan).</p>
Full article ">Figure 2
<p>Vertical to vertical component cross correlation from the source station to the receiver stations for each section. The first section is north of the Meishan Fault (MS01−MS17), and the next section (MS18−MS29) is inside CCU. The third (MS30−MS45) and fourth (MS45−MS60) sections are south of CCU.</p>
Full article ">Figure 3
<p>The maximum amplitude of the envelope of the stacked waveforms with respect to source beam slowness and receiver slowness for Stations MS22 and MS34, taken as an example. The color bar represents the amplitude of the stacked waveforms for each slowness, and the white cross indicates the location of the maximum amplitude.</p>
Full article ">Figure 4
<p>Histograms of repeated measurements of the slowness at Station MS22 at periods of 0.3 s, 0.5 s, and 0.7 s.</p>
Full article ">Figure 5
<p>The slowness measurements along the entire array at periods of 0.3 s, 0.5 s, and 0.7 s. The error bars show the uncertainties of the mean slowness.</p>
Full article ">Figure 6
<p>(<b>a</b>) The Rayleigh wave phase velocity profile and (<b>b</b>) its uncertainty profile as measured by the double beamforming method.</p>
Full article ">Figure 7
<p>(<b>a</b>) Cross-section of the shear velocities across the Meishan Fault. (<b>b</b>) Perturbations of the Vs profile, and perturbations relative to the average shear velocity at each depth.</p>
Full article ">
16 pages, 1674 KiB  
Communication
A Comparison of Stream Water and Shallow Groundwater Suspended Sediment Concentrations in a West Virginia Mixed-Use, Agro-Forested Watershed
by Kaylyn S. Gootman and Jason A. Hubbart
Land 2022, 11(4), 506; https://doi.org/10.3390/land11040506 - 31 Mar 2022
Cited by 1 | Viewed by 1487
Abstract
Suspended sediment is an important constituent of freshwater ecosystems that supports biogeochemical, geomorphological, and ecological processes. Current knowledge of suspended sediment is largely based on surface water studies; however, improved understanding of surface and in situ groundwater suspended sediment processes will improve pollutant [...] Read more.
Suspended sediment is an important constituent of freshwater ecosystems that supports biogeochemical, geomorphological, and ecological processes. Current knowledge of suspended sediment is largely based on surface water studies; however, improved understanding of surface and in situ groundwater suspended sediment processes will improve pollutant loading estimates and watershed remediation strategies. A study was conducted in a representative mixed-use, agro-forested catchment of the Chesapeake Bay Watershed of the northeast, USA, utilizing an experimental watershed study design, including eight nested sub-catchments. Stream water and shallow groundwater grab samples were collected monthly from January 2020 to December 2020 (n = 192). Water samples were analyzed for suspended sediment using gravimetric (mg/L) and laser particle diffraction (µm) analytical methods. Results showed that shallow groundwater contained significantly higher (p < 0.001) total suspended solid concentrations and smaller particle sizes, relative to stream water. Differences were attributed to variability between sites in terms of soil composition, land use/land cover, and surficial geology, and also the shallow groundwater sampling method used. Results hold important implications for pollutant transport estimates and biogeochemical modeling in agro-forested watersheds. Continued work is needed to improve shallow groundwater suspended sediment characterization (i.e., mass and particle sizes) and the utility of this information for strategies that are designed to meet water quality goals. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The location of West Virginia University Reymann Memorial Farm (RMF) within the West Virginia headwater region of the Chesapeake Bay Watershed (CBW) and (<b>b</b>) the land use/land cover for RMF, where red = mixed development; yellow = agriculture; green = upland forest; and blue = open water, and the locations of eight co-located, nested stilling wells and piezometers in Moore’s Run Watershed.</p>
Full article ">Figure 2
<p>Monthly (<span class="html-italic">n</span> = 12) stream water (SW) and shallow groundwater (SGW) (<b>A</b>) Total suspended solids concentration (TSS; mg/L); (<b>B</b>) mean particle size, MZ, (µm); (<b>C</b>) calculated surface area, CS, (m<sup>2</sup>/mL); and (<b>D</b>) skewness, Ski, (unitless) at each study site in Moore’s Run Watershed (<span class="html-italic">n</span> = 12, each source water type), located near Wardensville, West Virginia, USA, during the study period (January 2020–December 2020) and the study area average (Avg; <span class="html-italic">n</span> = 96, each source water type). Boxes define the interquartile range (IQR). Vertical lines show the range within 1.5 IQR. Midlines indicate the median. Open circles denote the mean. Filled-in diamonds represent data points.</p>
Full article ">Figure 3
<p>Monthly (<span class="html-italic">n</span> = 12) stream water (SW; blue) and shallow groundwater (SGW; red) particle size fractions (%), including sand, silt, and clay, from each study site (<span class="html-italic">n</span> = 8, both source water types) during the study period (January 2020–December 2020) in Moore’s Run Watershed, located near Wardensville, WV, USA. Average soil core particle size fractions by depth (i.e., dark brown = 0–5 cm, light brown = 25–30 cm, orange = 45–50 cm) from Gootman et al. [<a href="#B54-land-11-00506" class="html-bibr">54</a>] are also included for a comparison between water and soil particle size fractions.</p>
Full article ">Figure 3 Cont.
<p>Monthly (<span class="html-italic">n</span> = 12) stream water (SW; blue) and shallow groundwater (SGW; red) particle size fractions (%), including sand, silt, and clay, from each study site (<span class="html-italic">n</span> = 8, both source water types) during the study period (January 2020–December 2020) in Moore’s Run Watershed, located near Wardensville, WV, USA. Average soil core particle size fractions by depth (i.e., dark brown = 0–5 cm, light brown = 25–30 cm, orange = 45–50 cm) from Gootman et al. [<a href="#B54-land-11-00506" class="html-bibr">54</a>] are also included for a comparison between water and soil particle size fractions.</p>
Full article ">Figure 4
<p>Monthly (<span class="html-italic">n</span> = 12) relationships between stream water (SW) and shallow groundwater (SGW) particle size distribution (PSD) ratios (unitless) and particle diameter (µm) at each study site (<span class="html-italic">n</span> = 8) during the study period (January 2020–December 2020) in Moore’s Run Watershed, located near Wardensville, West Virginia, USA. Size classes are noted with solid vertical lines. Open symbols represent the monthly (<span class="html-italic">n</span> = 12) total suspended solids concentration (TSS; mg/L) versus the average particle diameter (Mz; µm).</p>
Full article ">
19 pages, 3888 KiB  
Article
Estimating Historically Cleared and Forested Land in Massachusetts, USA, Using Airborne LiDAR and Archival Records
by Katharine M. Johnson, William B. Ouimet, Samantha Dow and Cheyenne Haverfield
Remote Sens. 2021, 13(21), 4318; https://doi.org/10.3390/rs13214318 - 27 Oct 2021
Cited by 10 | Viewed by 4217
Abstract
In the northeastern United States, widespread deforestation occurred during the 17–19th centuries as a result of Euro-American agricultural activity. In the late 19th and early 20th centuries, much of this agricultural landscape was reforested as the region experienced industrialization and farmland became abandoned. [...] Read more.
In the northeastern United States, widespread deforestation occurred during the 17–19th centuries as a result of Euro-American agricultural activity. In the late 19th and early 20th centuries, much of this agricultural landscape was reforested as the region experienced industrialization and farmland became abandoned. Many previous studies have addressed these landscape changes, but the primary method for estimating the amount and distribution of cleared and forested land during this time period has been using archival records. This study estimates areas of cleared and forested land using historical land use features extracted from airborne LiDAR data and compares these estimates to those from 19th century archival maps and agricultural census records for several towns in Massachusetts, a state in the northeastern United States. Results expand on previous studies in adjacent areas, and demonstrate that features representative of historical deforestation identified in LiDAR data can be reliably used as a proxy to estimate the spatial extents and area of cleared and forested land in Massachusetts and elsewhere in the northeastern United States. Results also demonstrate limitations to this methodology which can be mitigated through an understanding of the surficial geology of the region as well as sources of error in archival materials. Full article
(This article belongs to the Special Issue Remote Sensing of Past Human Land Use)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Change in improved land (km<sup>2</sup>) in the northeastern U.S. by county, 1850–1870, compiled from U.S. Census Nonpopulation Schedule for Agriculture data [<a href="#B29-remotesensing-13-04318" class="html-bibr">29</a>,<a href="#B30-remotesensing-13-04318" class="html-bibr">30</a>]. This census recorded the amount of improved and unimproved farmland at the town level beginning in 1850. Counties in gray did not exist in 1850 and therefore continuous data are not available. Map projection is NAD 1983 UTM Zone 18N. Abbreviations for states in the inset map are as follows: CT, Connecticut; MA, Massachusetts; ME, Maine; NH, New Hampshire; NY, New York; RI, Rhode Island; VT, Vermont.</p>
Full article ">Figure 2
<p>(<b>A</b>,<b>D</b>) depict examples of extant historical land use features as seen in LiDAR DEM derivatives ((<b>A</b>) shows stone walls in a hillshaded DEM while (<b>D</b>) shows relict charcoal hearths in a slope raster). (<b>B</b>,<b>E</b>) depict current land cover (dense forest) in high-resolution aerial photography for those same extents, and (<b>C</b>,<b>F</b>) show examples of those feature types as they appear in the field.</p>
Full article ">Figure 3
<p>Overview of study town locations in Massachusetts. Change in improved farmland (km<sup>2</sup>) from 1850 to 1870 is indicated for each town, compiled from U.S. Census Nonpopulation Schedule for Agriculture data [<a href="#B25-remotesensing-13-04318" class="html-bibr">25</a>,<a href="#B26-remotesensing-13-04318" class="html-bibr">26</a>]. Map projection is NAD 1983 UTM Zone 18N.</p>
Full article ">Figure 4
<p>1830 forest cover in Massachusetts from the Harvard Forest dataset digitized from 1830 archival maps [<a href="#B57-remotesensing-13-04318" class="html-bibr">57</a>] (also see [<a href="#B34-remotesensing-13-04318" class="html-bibr">34</a>], Figure 6 on p. 1327). Towns in dark gray do not have any data. Map projection is NAD 1983 UTM Zone 18N.</p>
Full article ">Figure 5
<p>Example showing stone wall (<b>A</b>) and relict charcoal hearth (RCH) (<b>B</b>) density in West Stockbridge, Massachusetts, and reclassified areas where stone wall length exceeds 2 km/km<sup>2</sup> (<b>C</b>) and RCH density exceeds 10/km<sup>2</sup> (<b>D</b>). Map projection is NAD 1983 UTM Zone 18N.</p>
Full article ">Figure 6
<p>Scatterplots showing the relationship between improved (i.e., non-forested) area from archival records and LiDAR-derived area where stone wall density is greater than 2 km/km<sup>2</sup> in all study towns (black dots) for the years (<b>A</b>) 1830, (<b>B</b>) 1850, (<b>C</b>) 1860, and (<b>D</b>) 1870. Note that Monterey, Sandisfield, and Plainfield are not included in the 1860 plot (<b>C</b>), and Monterey and Sandisfield are not included in the 1870 plot (<b>D</b>). The boundaries of Monterey and Sandisfield changed after 1850, so their areas are not comparable, and Plainfield did not have agricultural census data available for 1860. The gray 1:1 line indicates the location where the estimated area from stone walls is equal to the value recorded in the agricultural census.</p>
Full article ">Figure 7
<p>LiDAR-derived area where stone wall length is &gt;2 km/km<sup>2</sup> plotted against areas of improved land in 1850 for Massachusetts (MA) and Connecticut (CT) towns combined. Data and plot for Connecticut towns are discussed in greater detail in [<a href="#B9-remotesensing-13-04318" class="html-bibr">9</a>,<a href="#B62-remotesensing-13-04318" class="html-bibr">62</a>]. The gray 1:1 line indicates the location where the estimated area from stone walls is equal to the value recorded in the agricultural census.</p>
Full article ">Figure 8
<p>Relict charcoal hearths (RCHs) tend to occur in areas of towns which were typically not well-suited for agriculture, as shown here in West Stockbridge, Massachusetts, where RCHs occur in areas of the town which were forested in 1830. Map projection is NAD 1983 UTM Zone 18N.</p>
Full article ">Figure 9
<p>Panels (<b>A</b>) through (<b>C</b>) show the same extent of the eastern portion of Clarksburg, Massachusetts: (<b>A</b>) is the archival 1830 map (Public Domain, [<a href="#B80-remotesensing-13-04318" class="html-bibr">80</a>]) showing forest covering mountainous areas, (<b>B</b>) shows that map georeferenced and overlaid with 30% transparency on a LiDAR DEM, and (<b>C</b>) shows (B) plus the current extent of digitized forest cover from the Harvard Forest dataset which was digitized from the same map in the early 2000s. Map projection is NAD 1983 UTM Zone 18N.</p>
Full article ">
23 pages, 17014 KiB  
Article
A High-Resolution, Random Forest Approach to Mapping Depth-to-Bedrock across Shallow Overburden and Post-Glacial Terrain
by Shane Furze, Antóin M. O’Sullivan, Serge Allard, Toon Pronk and R. Allen Curry
Remote Sens. 2021, 13(21), 4210; https://doi.org/10.3390/rs13214210 - 20 Oct 2021
Cited by 12 | Viewed by 3213
Abstract
Regolith, or unconsolidated materials overlying bedrock, exists as an active zone for many geological, geomorphological, hydrological and ecological processes. This zone and its processes are foundational to wide-ranging human needs and activities such as water supply, mineral exploration, forest harvesting, agriculture, and engineered [...] Read more.
Regolith, or unconsolidated materials overlying bedrock, exists as an active zone for many geological, geomorphological, hydrological and ecological processes. This zone and its processes are foundational to wide-ranging human needs and activities such as water supply, mineral exploration, forest harvesting, agriculture, and engineered structures. Regolith thickness, or depth-to-bedrock (DTB), is typically unavailable or restricted to finer scale assessments because of the technical and cost limitations of traditional drilling, seismic, and ground-penetrating radar surveys. The objective of this study was to derive a high-resolution (10 m2) DTB model for the province of New Brunswick, Canada as a case study. This was accomplished by developing a DTB database from publicly available soil profiles, boreholes, drill holes, well logs, and outcrop transects (n = 203,238). A Random Forest model was produced by modeling the relationships between DTB measurements in the database to gridded datasets derived from both a LiDAR-derived digital elevation model and photo-interpreted surficial geology delineations. In developing the Random Forest model, DTB measurements were split 70:30 for model development and validation, respectively. The DTB model produced an R2 = 92.8%, MAE = 0.18 m, and RMSE = 0.61 m for the training, and an R2 = 80.3%, MAE = 0.18 m, and RMSE = 0.66 m for the validation data. This model provides an unprecedented resolution of DTB variance at a landscape scale. Additionally, the presented framework provides a fundamental understanding of regolith thickness across a post-glacial terrain, with potential application at the global scale. Full article
Show Figures

Figure 1

Figure 1
<p>Visualization of the geographic location of New Brunswick, Canada (46.5653° N, 66.4619° W) outlining elevation change and physiographic regions (<b>A</b>) and an overview of bedrock lithology (<b>B</b>), overlain on World Ocean Basemap [<a href="#B30-remotesensing-13-04210" class="html-bibr">30</a>] in ArcGIS 10.3 software [<a href="#B31-remotesensing-13-04210" class="html-bibr">31</a>]. Bedrock geology was acquired from [<a href="#B28-remotesensing-13-04210" class="html-bibr">28</a>] (Scale 1:2,500,000).</p>
Full article ">Figure 2
<p>Overview of sample locations for each DTB source: boreholes (<b>A</b>); drillholes (<b>B</b>); pedons (<b>C</b>); site cards (<b>D</b>); well logs (<b>E</b>); and bedrock outcrops (<b>F</b>). Total sample size = 170,719 (see also <a href="#remotesensing-13-04210-t002" class="html-table">Table 2</a>).</p>
Full article ">Figure 3
<p>Histogram and continuous density distribution (red line) of bedrock depth for boreholes (<b>A</b>), drillholes (<b>B</b>), pedons (<b>C</b>), site cards (<b>D</b>), and well logs (<b>E</b>), and all data combined (<b>F</b>), including sample size, minimum, maximum, and mean DTB values for each source. Note: Depth intervals are (i) 0.25 m for pedons and site cards, and 5 m for the remaining sources, and (ii) the total sample size with minimum, maximum, and mean exclude the rock outcrop samples since all of these samples have a DTB of 0.</p>
Full article ">Figure 4
<p>The framework for developing the DTB model covariates. The multi-scale generation of covariates beginning with the 10 m DEM is shown as an example (blue frame) and the correlation analyses that followed to reduce the number of covariates (red frame).</p>
Full article ">Figure 5
<p>Scree plot used to select the number of PCs to capture a minimum of 90% variability (<span class="html-italic">n</span> = 11).</p>
Full article ">Figure 6
<p>The RF variable importance plot highlighting the explanatory strength of the 20 most significant covariates with importance measured as decreasing from 100% (most important) to 40%. Covariate abbreviations are explained in <a href="#remotesensing-13-04210-t003" class="html-table">Table 3</a>.</p>
Full article ">Figure 7
<p>Actual vs. fitted model results for the training and validation subsets ((<b>A</b>,<b>B</b>), respectively) including 95% confidence limits, and histograms displaying densities of residual errors for DTB model on training and validation subsets ((<b>C</b>,<b>D</b>), respectively).</p>
Full article ">Figure 8
<p>Predicted depth to bedrock (DTB) at 10 m<sup>2</sup> resolution for the Province of New Brunswick, Canada (<b>A</b>) with an example of the fine-scale resolution showing sediment accumulation across an upland–hillslope–valley bottom transition (<b>B</b>).</p>
Full article ">
22 pages, 15387 KiB  
Article
High Resolution Apparent Thermal Inertia Mapping on Mars
by Marta Ciazela, Jakub Ciazela and Bartosz Pieterek
Remote Sens. 2021, 13(18), 3692; https://doi.org/10.3390/rs13183692 - 15 Sep 2021
Cited by 8 | Viewed by 3368
Abstract
Thermal inertia, which represents the resistance to change in temperature of the upper few centimeters of the surface, provides information to help understand the surficial geology and recent processes that are potentially still active today. It cannot be directly measured on Mars and [...] Read more.
Thermal inertia, which represents the resistance to change in temperature of the upper few centimeters of the surface, provides information to help understand the surficial geology and recent processes that are potentially still active today. It cannot be directly measured on Mars and is therefore usually modelled. We present a new analytical method based on Apparent Thermal Inertia (ATI), a thermal inertia proxy. Calculating ATI requires readily available input data: temperature, incidence angle, visible dust opacity, and a digital elevation model. Because of the high spatial resolution, the method can be used on sloping terrains, which makes possible thermal mapping using THEMIS in nearly any area of Mars. Comparison with results obtained by other approaches using modeled data shows similarity in flat areas and illustrates the significant influence of slope and aspect on albedo and diurnal temperature differences. Full article
(This article belongs to the Special Issue Cartography of the Solar System: Remote Sensing beyond Earth)
Show Figures

Figure 1

Figure 1
<p>Mosaic of MRO/CTX images (P01_001351_1717_XI_08S084W, P01_001417_1718_XI_08S085W) used to evaluate CTX capability at determining albedo. Grid line spacing is 8 pixels per degree (ppd), similar to the TES albedo map available at <a href="https://www.mars.asu.edu/data/tes_albedo/" target="_blank">https://www.mars.asu.edu/data/tes_albedo/</a> (accessed on 14 September 2021). Average CTX and TES albedo are calculated for the TES pixels that are entirely located within the area covered by the CTX images.</p>
Full article ">Figure 2
<p>(<b>a</b>) CTX vs. TES albedo for the 8 ppd grid pixels displayed in <a href="#remotesensing-13-03692-f001" class="html-fig">Figure 1</a>; (<b>b</b>) comparison between 1-A (which is directly proportional to ATI, Equation (3)) for the CTX and TES data. CTX albedo is on average 8% higher than TES albedo (<b>a</b>), which implies CTX-calculated ATI would be 1.5% than TES-based ATI (<b>b</b>).</p>
Full article ">Figure 3
<p>Elevation map of the Valles Marineris region derived from the MEX HRSC Blended DEM Global 200m_v2 (<b>A</b>) with the location of the testing area on the MOLA shaded relief basemap adapted from [<a href="#B41-remotesensing-13-03692" class="html-bibr">41</a>] (<b>B</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>) CTX visible light image mosaic of the test area (P01_001417_1718_XI_08S085W: t = 15.56 h, Ls = 135.5°, and P01_001351_1717_XI_08S084W: t = 15.55 h, Ls = 133.0°); (<b>b</b>) albedo map derived from CTX data, showing the albedo values corrected against topographic slope and aspect; (<b>c</b>) classes generated by unsupervised classification using the isodata algorithm [<a href="#B45-remotesensing-13-03692" class="html-bibr">45</a>] in IDRISI Selva [<a href="#B46-remotesensing-13-03692" class="html-bibr">46</a>]; the white zones in (<b>b</b>,<b>c</b>), which cover 6.2% of the image, are excluded from calculations due to incidence angle &gt;79° or full shade; (<b>d</b>) Mars Express High-Resolution Stereo Camera (HRSC) digital terrain model (DTM) of the study area (in meters above sea level); (<b>e</b>) slope inclination map (in degrees) of the study area. Average albedo for each class is listed in <a href="#remotesensing-13-03692-t001" class="html-table">Table 1</a>. Coordinates of the top-left corner of the displayed area: 85°15′58.267″W, 7°8′57.982″S.</p>
Full article ">Figure 5
<p>(<b>a</b>) Daytime LST map; (<b>b</b>) nighttime map; (<b>c</b>) difference between daytime and nighttime LST; (<b>d</b>) difference between daytime and nighttime LST after correction following Equation (8); (<b>e</b>) difference between daytime and nighttime LST after correction following Equation (8) shifted to the real minimum and maximum of temperature (see second paragraph in <a href="#sec3dot2-remotesensing-13-03692" class="html-sec">Section 3.2</a>). Images are IDs are provided in <a href="#remotesensing-13-03692-t002" class="html-table">Table 2</a>. The true temperature ranges are 229.5–280.6 K (<b>a</b>), 174.4–211.3 K (<b>b</b>), 28.4–82.0 K (<b>c</b>), 32.8–105.7 K (<b>d</b>), and 46.2–155.4 K (<b>e</b>). The white zones in (<b>d</b>,<b>e</b>) are excluded from calculations due to incidence angle &gt; 79° or full shade.</p>
Full article ">Figure 6
<p>Seasonal variability of surface temperature during a martian year at 8° S at 00:00 LST, 06:00 LST, 12:00 LST and 18:00 LST using MARSTHERM. TauD is dust opacity, Agnd is ground albedo, I is thermal inertia, P is surface pressure, Lat is latitude, IceT is the semi-infinite amount of CO<sub>2</sub> ice with respect to albedo and emissivity in the MARSTHERM physical modelling scheme, NT is ground time step per day, with 10 min steps, NFQ is the number of ground time steps per atmospheric time step, SLANG is surface slope inclination, SLAZI is surface slope aspect.</p>
Full article ">Figure 7
<p>(<b>a</b>) ATIc map (J m<sup>−2</sup> K<sup>−1</sup>s<sup>−1/2</sup><b>)</b>; (<b>b</b>) thermal inertia (TI) map based on the thermal inertia mosaic of [<a href="#B2-remotesensing-13-03692" class="html-bibr">2</a>] released in 2014 (J m<sup>−2</sup> K<sup>−1</sup>s<sup>−1/2</sup>). The thermal inertia scale in (<b>a</b>,<b>b</b>) has been unified for easier comparison; the true ranges are 157–729 for (<b>a</b>) and 88–736 for (<b>b</b>). The white zones in (<b>a</b>,<b>b</b>) are excluded from calculations due to incidence angle &gt; 79° or full shade, which reduces the energy input on the slope. The thick black lines in (<b>a</b>,<b>b</b>) indicate boundaries between individual THEMIS images. The red-hatched zones in (<b>a</b>) receive high amount of reflected radiation (<span class="html-italic">I<sub>R</sub></span>) (see <a href="#sec5dot2-remotesensing-13-03692" class="html-sec">Section 5.2</a> Model Limitations). Areas 1–5 are discussed in <a href="#sec4dot1-remotesensing-13-03692" class="html-sec">Section 4.1</a>.</p>
Full article ">Figure 8
<p>MRO/CTX visible light image mosaic of the study area and the dune fields tested for <span class="html-italic">ATI<sub>c</sub></span> evaluation.</p>
Full article ">Figure 9
<p>Thermal inertia for dunes according to <span class="html-italic">ATI<sub>c</sub></span> calculation and other methods. Whiskers represent the total range of values for the dune fields in <a href="#remotesensing-13-03692-f008" class="html-fig">Figure 8</a>, and the blue boxes span from the arithmetic mean to ±1 standard deviation. The value of the red line (251) is calculated from the theoretical minimum dune sand grain size of 215 µm in Martian conditions [<a href="#B42-remotesensing-13-03692" class="html-bibr">42</a>] and empirical relationship between thermal conductivity and grain size derived by [<a href="#B43-remotesensing-13-03692" class="html-bibr">43</a>] assuming a specific heat of 850 J·kg<sup>−1</sup>/K<sup>−1</sup> and a bulk density of 1650.0 kg·m<sup>−3</sup> [<a href="#B52-remotesensing-13-03692" class="html-bibr">52</a>]. Similarly, the modelled TI values are derived assuming typical 470–600 µm grain size provided by [<a href="#B42-remotesensing-13-03692" class="html-bibr">42</a>].</p>
Full article ">Figure 10
<p>ATI<sub>c</sub> and TI [<a href="#B2-remotesensing-13-03692" class="html-bibr">2</a>] histogram distributions (extracted from <a href="#remotesensing-13-03692-f007" class="html-fig">Figure 7</a>a,b). The number of histograms peaks differ from albedo in <a href="#remotesensing-13-03692-f004" class="html-fig">Figure 4</a>c as the thermal inertia dependent both on albedo and temperature changes.</p>
Full article ">Figure 11
<p>Regional profile across a flat part of the study area. Thermal inertia (TI) results of Fergason et al. [<a href="#B2-remotesensing-13-03692" class="html-bibr">2</a>] are compared to our ATI results and geomorphological interpretation of the surface. Note the CTX image boundaries in both methods and their effects on the obtained results.</p>
Full article ">Figure 12
<p>High-resolution profiles across various slopes in the study area. Thermal inertia (TI) results of Fergason et al. [<a href="#B2-remotesensing-13-03692" class="html-bibr">2</a>] are compared to our ATI results and geomorphological interpretation of the surface.</p>
Full article ">Figure 13
<p>Percentage (<span class="html-italic">R</span>) of the reflected radiation (<span class="html-italic">I<sub>R</sub></span>) with respect to the total radiation calculated with Equation (14) as a function of the observed slope (<span class="html-italic">s</span>), opposite slope (<span class="html-italic">s*</span>), and albedo (<span class="html-italic">A</span>). The blue field indicates the expected error range (1.2 to 5.3%, dashed lines) for the valleys in the study area, where typical slopes are 20–30° and albedo is 0.10–0.21.</p>
Full article ">
23 pages, 2945 KiB  
Article
Bio-Inspired Hybridization of Artificial Neural Networks: An Application for Mapping the Spatial Distribution of Soil Texture Fractions
by Ruhollah Taghizadeh-Mehrjardi, Mostafa Emadi, Ali Cherati, Brandon Heung, Amir Mosavi and Thomas Scholten
Remote Sens. 2021, 13(5), 1025; https://doi.org/10.3390/rs13051025 - 8 Mar 2021
Cited by 42 | Viewed by 4354
Abstract
Soil texture and particle size fractions (PSFs) are a critical characteristic of soil that influences most physical, chemical, and biological properties of soil; furthermore, reliable spatial predictions of PSFs are crucial for agro-ecological modeling. Here, series of hybridized artificial neural network (ANN) models [...] Read more.
Soil texture and particle size fractions (PSFs) are a critical characteristic of soil that influences most physical, chemical, and biological properties of soil; furthermore, reliable spatial predictions of PSFs are crucial for agro-ecological modeling. Here, series of hybridized artificial neural network (ANN) models with bio-inspired metaheuristic optimization algorithms such as a genetic algorithm (GA-ANN), particle swarm optimization (PSO-ANN), bat (BAT-ANN), and monarch butterfly optimization (MBO-ANN) algorithms, were built for predicting PSFs for the Mazandaran Province of northern Iran. In total, 1595 composite surficial soil samples were collected, and 64 environmental covariates derived from terrain, climatic, remotely sensed, and categorical datasets were used as predictors. Models were tested using a repeated 10-fold nested cross-validation approach. The results indicate that the hybridized ANN methods were far superior to the reference approach using ANN with a backpropagation training algorithm (BP-ANN). Furthermore, the MBO-ANN approach was consistently determined to be the best approach and yielded the lowest error and uncertainty. The MBO-ANN model improved the predictions in terms of RMSE by 20% for clay, 10% for silt, and 24% for sand when compared to BP-ANN. The physiographical units, soil types, geology maps, rainfall, and temperature were the most important predictors of PSFs, followed by the terrain and remotely sensed data. This study demonstrates the effectiveness of bio-inspired algorithms for improving ANN models. The outputs of this study will support and inform sustainable soil management practices, agro-ecological modeling, and hydrological modeling for the Mazandaran Province of Iran. Full article
(This article belongs to the Special Issue Remote Sensing for Soil Mapping and Monitoring)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Localization of the Mazandaran province in Iran (<b>a</b>); and spatial distribution of soil sampling locations within the Mazandaran Province (<b>b</b>).</p>
Full article ">Figure 2
<p>Flowchart diagram of the ANN training using metaheuristic optimization algorithms. α and β represent the weights and biases, respectively.</p>
Full article ">Figure 3
<p>Distribution of soil texture classes for all samples based on the USDA soil texture triangle. (Cl: clay; SiCl: silty clay; SiClLo: silty clay loam; SaCl: sandy clay; SaClLo: sandy clay loam; ClLo: clay loam; Si: silt; SiLo: silt loam; Lo: loam; Sa: sand; LoSa: loamy sand; SaLo: sandy loam).</p>
Full article ">Figure 4
<p>The relative influence of environmental covariates for predicting clay, silt, and sand contents using the MBO-ANN algorithm. (Refer to <a href="#remotesensing-13-01025-t0A1" class="html-table">Table A1</a> for covariate codes).</p>
Full article ">Figure 5
<p>Continues maps produced using the MBO-ANN model for the Mazandaran Province of: surficial clay content (<b>a</b>); surficial sand content (<b>b</b>); and surficial silt content (<b>c</b>); and textual classes (<b>d</b>).</p>
Full article ">
14 pages, 4557 KiB  
Article
Periglacial Lake Origin Influences the Likelihood of Lake Drainage in Northern Alaska
by Mark Jason Lara and Melissa Lynn Chipman
Remote Sens. 2021, 13(5), 852; https://doi.org/10.3390/rs13050852 - 25 Feb 2021
Cited by 7 | Viewed by 3291
Abstract
Nearly 25% of all lakes on earth are located at high latitudes. These lakes are formed by a combination of thermokarst, glacial, and geological processes. Evidence suggests that the origin of periglacial lake formation may be an important factor controlling the likelihood of [...] Read more.
Nearly 25% of all lakes on earth are located at high latitudes. These lakes are formed by a combination of thermokarst, glacial, and geological processes. Evidence suggests that the origin of periglacial lake formation may be an important factor controlling the likelihood of lakes to drain. However, geospatial data regarding the spatial distribution of these dominant Arctic and subarctic lakes are limited or do not exist. Here, we use lake-specific morphological properties using the Arctic Digital Elevation Model (DEM) and Landsat imagery to develop a Thermokarst lake Settlement Index (TSI), which was used in combination with available geospatial datasets of glacier history and yedoma permafrost extent to classify Arctic and subarctic lakes into Thermokarst (non-yedoma), Yedoma, Glacial, and Maar lakes, respectively. This lake origin dataset was used to evaluate the influence of lake origin on drainage between 1985 and 2019 in northern Alaska. The lake origin map and lake drainage datasets were synthesized using five-year seamless Landsat ETM+ and OLI image composites. Nearly 35,000 lakes and their properties were characterized from Landsat mosaics using an object-based image analysis. Results indicate that the pattern of lake drainage varied by lake origin, and the proportion of lakes that completely drained (i.e., >60% area loss) between 1985 and 2019 in Thermokarst (non-yedoma), Yedoma, Glacial, and Maar lakes were 12.1, 9.5, 8.7, and 0.0%, respectively. The lakes most vulnerable to draining were small thermokarst (non-yedoma) lakes (12.7%) and large yedoma lakes (12.5%), while the most resilient were large and medium-sized glacial lakes (4.9 and 4.1%) and Maar lakes (0.0%). This analysis provides a simple remote sensing approach to estimate the spatial distribution of dominant lake origins across variable physiography and surficial geology, useful for discriminating between vulnerable versus resilient Arctic and subarctic lakes that are likely to change in warmer and wetter climates. Full article
(This article belongs to the Special Issue Dynamic Disturbance Processes in Permafrost Regions)
Show Figures

Figure 1

Figure 1
<p>Upland tussock-shrub tundra region of Alaska. The bolded line bounds our region of interest, which includes (i) nearly 35,000 lakes greater than 1 ha in size, (ii) the known yedoma deposits, (iii) the Wisconsinan glacial extent during the last glacial maximum, and (iv) seven ecoregions: Brooks Foothills, Brooks Range, Davidson Mountains, Kobuk Ridges and Valleys, Kotzebue Sound Lowlands, and the Seward Peninsula.</p>
Full article ">Figure 2
<p>Schematic representation of the data processing workflow used to derive the Thermokarst lake Settlement Index (TSI) and map the distribution of lake origin across our study domain. Processing within Google Earth Engine (GEE), eCognition, R, and ArcGIS are indicated by background colors white, blue, grey, and green, respectively.</p>
Full article ">Figure 3
<p>Morphological lake metrics summarized across all lakes in our northern Alaska study domain. TSI = Thermokarst lake Settlement Index.</p>
Full article ">Figure 4
<p>Examples of the Thermokarst lake Settlement Index (TSI) projected on Landsat scenes (<b>top row</b>) within the Kobuk River Delta (<b>left column</b>), southwestern Noatak National Preserve (<b>middle column</b>), and central Seward Peninsula (<b>right column</b>). The TSI is overlaid on a greyscale hillshade surface computed with the Arctic Digital Elevation Model (DEM) (<b>bottom row</b>).</p>
Full article ">Figure 5
<p>The probability of thermokarst and non-thermokarst lakes using the Thermokarst lake Settlement Index (TSI). Violin plots display the differing distribution of TSI values for 3335 thermokarst and 1854 non-thermokarst lakes, where the median TSIs were 0.467 and 2.657, respectively (<b>A</b>). Logistic regression identified a TSI of 1.27 as the 50% probability threshold to classify thermokarst and non-thermokarst lakes (<b>B</b>).</p>
Full article ">Figure 6
<p>Lake origin map created for the northern tussock-shrub tundra region of Alaska. Red and orange panels show the diversity of lake origins within the northern Seward Peninsula and the eastern Noatak National Preserve, respectively.</p>
Full article ">Figure A1
<p>Lake-specific 500 m buffers used with the Arctic DEM to compute lake settlement.</p>
Full article ">Figure A2
<p>Lake sub-regions that correspond with relatively homogenous thermokarst (non-yedoma) and non-thermokarst lakes (including yedoma and glacial lakes).</p>
Full article ">
18 pages, 52251 KiB  
Article
Geology and Aquifer Sensitivity of Quaternary Glacial Deposits Overlying a Portion of the Mahomet Buried Bedrock Valley Aquifer System
by Andrew Watson, Eric W. Peterson, Dave Malone and Lisa Tranel
Hydrology 2020, 7(4), 69; https://doi.org/10.3390/hydrology7040069 - 23 Sep 2020
Cited by 2 | Viewed by 3160
Abstract
To characterize the distribution of Holocene and Late Quaternary deposits and to assess the contamination potential of the Mahomet Aquifer, surficial geologic and aquifer sensitivity maps of the Gibson City East 7.5-Minute Quadrangle were created. Geologic data, extent, and thickness of the geologic [...] Read more.
To characterize the distribution of Holocene and Late Quaternary deposits and to assess the contamination potential of the Mahomet Aquifer, surficial geologic and aquifer sensitivity maps of the Gibson City East 7.5-Minute Quadrangle were created. Geologic data, extent, and thickness of the geologic materials were coupled with LiDAR topographic data and analyzed using ESRI’s ArcGIS 10.6.1. Aquifer sensitivity to contamination was calculated based on the depth to the first aquifer unit, aquifer thickness, and the lithology of the aquifer materials. The surficial geologic mapping identified five lithostratigraphic units: the Cahokia Formation, the Equality Formation, the Henry Formation, and the Yorkville and Batestown Members of the Lemont Formation. The southeast to northwest trending Illiana Morainic System is the most prominent feature in the study area and delineates the maximum extent of the glaciers during the Livingston Phase of glaciation. Postglacial deposits of the Cahokia Formation, alluvium, interfinger, and overlie with glacial outwash of the Henry Formation along channels and drainage ways downslope of the moraine. The areas of least sensitivity are located over the Illiana Morainic System, whereas the greatest potential to contamination occurs where the thickest deposits of the Henry Formation and Cahokia Formation lie at or just below the land surface. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the Mahomet Aquifer (gray area) in east-central Illinois, USA. Mapping area is highlighted by red box. Figure modified from [<a href="#B45-hydrology-07-00069" class="html-bibr">45</a>].</p>
Full article ">Figure 2
<p>Generalized lithostratigraphy of Quaternary-age sediments in east-central Illinois. All units that rest unconformably above the bedrock strata are Quaternary in age. Modified from [<a href="#B39-hydrology-07-00069" class="html-bibr">39</a>].</p>
Full article ">Figure 3
<p>Surficial geologic map of the Gibson City East Quadrangle. A higher resolution pdf has been provided as a supplementary file.</p>
Full article ">Figure 4
<p>LiDAR hillshaded DEM depicting the following geomorphologic features within the Gibson City East Quadrangle: morainic system outlined in blue (and yellow where the boundary is shared with alluvium), flood plains/river channels outlined in yellow, outwash plain outlined in brown, glacial lake plains outlined in purple, and esker/kame-type features in red. Red numbers indicate the section lines. Dashed red line outlines the area illustrated in <a href="#hydrology-07-00069-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 5
<p>LiDAR hillshaded DEM outlined in <a href="#hydrology-07-00069-f004" class="html-fig">Figure 4</a> resembling the following geomorphologic features within the Gibson City East Quadrangle: glacial lake plains outlined in purple, outwash plain outlined in brown, and esker-type ridges/kame-like features outlined in red. The red numbers indicate the section number.</p>
Full article ">Figure 6
<p>Aquifer sensitivity map of the Gibson City East Quadrangle. A higher resolution pdf has been provided as a supplementary file.</p>
Full article ">
16 pages, 822 KiB  
Article
A Case-Study Application of the Experimental Watershed Study Design to Advance Adaptive Management of Contemporary Watersheds
by Jason A. Hubbart, Elliott Kellner and Sean J. Zeiger
Water 2019, 11(11), 2355; https://doi.org/10.3390/w11112355 - 9 Nov 2019
Cited by 26 | Viewed by 5000
Abstract
Land managers are often inadequately informed to make management decisions in contemporary watersheds, in which sources of impairment are simultaneously shifting due to the combined influences of land use change, rapid ongoing human population growth, and changing environmental conditions. There is, thus, a [...] Read more.
Land managers are often inadequately informed to make management decisions in contemporary watersheds, in which sources of impairment are simultaneously shifting due to the combined influences of land use change, rapid ongoing human population growth, and changing environmental conditions. There is, thus, a great need for effective collaborative adaptive management (CAM; or derivatives) efforts utilizing an accepted methodological approach that provides data needed to properly identify and address past, present, and future sources of impairment. The experimental watershed study design holds great promise for meeting such needs and facilitating an effective collaborative and adaptive management process. To advance understanding of natural and anthropogenic influences on sources of impairment, and to demonstrate the approach in a contemporary watershed, a nested-scale experimental watershed study design was implemented in a representative, contemporary, mixed-use watershed located in Midwestern USA. Results identify challenges associated with CAM, and how the experimental watershed approach can help to objectively elucidate causal factors, target critical source areas, and provide the science-based information needed to make informed management decisions. Results show urban/suburban development and agriculture are primary drivers of alterations to watershed hydrology, streamflow regimes, transport of multiple water quality constituents, and stream physical habitat. However, several natural processes and watershed characteristics, such as surficial geology and stream system evolution, are likely compounding observed water quality impairment and aquatic habitat degradation. Given the varied and complicated set of factors contributing to such issues in the study watershed and other contemporary watersheds, watershed restoration is likely subject to physical limitations and should be conceptualized in the context of achievable goals/objectives. Overall, results demonstrate the immense, globally transferrable value of the experimental watershed approach and coupled CAM process to address contemporary water resource management challenges. Full article
Show Figures

Figure 1

Figure 1
<p>Locations of gauge sites (where #4 includes the USGS gauging station) and corresponding drainage area to each gauge (bold line) in the Hinkson Creek Watershed (HCW), in Central Missouri, USA. A model urban nested-scale experimental watershed study.</p>
Full article ">
27 pages, 4605 KiB  
Article
Drone-Borne Hyperspectral and Magnetic Data Integration: Otanmäki Fe-Ti-V Deposit in Finland
by Robert Jackisch, Yuleika Madriz, Robert Zimmermann, Markku Pirttijärvi, Ari Saartenoja, Björn H. Heincke, Heikki Salmirinne, Jukka-Pekka Kujasalo, Louis Andreani and Richard Gloaguen
Remote Sens. 2019, 11(18), 2084; https://doi.org/10.3390/rs11182084 - 5 Sep 2019
Cited by 55 | Viewed by 10390
Abstract
The technical evolution of unmanned aerial systems (UAS) for mineral exploration advances rapidly. Recent sensor developments and improved UAS performance open new fields for research and applications in geological and geophysical exploration among others. In this study, we introduce an integrated acquisition and [...] Read more.
The technical evolution of unmanned aerial systems (UAS) for mineral exploration advances rapidly. Recent sensor developments and improved UAS performance open new fields for research and applications in geological and geophysical exploration among others. In this study, we introduce an integrated acquisition and processing strategy for drone-borne multi-sensor surveys combining optical remote sensing and magnetic data. We deploy both fixed-wing and multicopter UAS to characterize an outcrop of the Otanmäki Fe-Ti-V deposit in central Finland. The lithology consists mainly of gabbro intrusions hosting ore bodies of magnetite-ilmenite. Large areas of the outcrop are covered by lichen and low vegetation. We use two drone-borne multi- and hyperspectral cameras operating in the visible to near-infrared parts of the electromagnetic spectrum to identify dominant geological features and the extents of ore bodies via iron-indicating proxy minerals. We apply band ratios and unsupervised and supervised image classifications on the spectral data, from which we can map surficial iron-bearing zones. We use two setups with three-axis fluxgate magnetometers deployed both by a fixed-wing and a multi-copter UAS to measure the magnetic field at various flight altitudes (15 m, 40 m, 65 m). The total magnetic intensity (TMI) computed from the individual components is used for further interpretation of ore distribution. We compare to traditional magnetic ground-based survey data to evaluate the UAS-based results. The measured anomalies and spectral data are validated and assigned to the outcropping geology and ore mineralization by performing surface spectroscopy, portable X-ray fluorescence (pXRF), magnetic susceptibility, and traditional geologic mapping. Locations of mineral zones and magnetic anomalies correlate with the established geologic map. The integrated survey strategy allowed a straightforward mapping of ore occurrences. We highlight the efficiency, spatial resolution, and reliability of UAS surveys. Acquisition time of magnetic UAS surveying surpassed ground surveying by a factor of 20 with a comparable resolution. The proposed workflow possibly facilitates surveying, particularly in areas with complicated terrain and of limited accessibility, but highlights the remaining challenges in UAS mapping. Full article
(This article belongs to the Special Issue Trends in UAV Remote Sensing Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Hill-shaded total magnetic intensity (TMI) map of the Otanmäki area, central Finland, that is based on regional airborne surveys from GTK [<a href="#B31-remotesensing-11-02084" class="html-bibr">31</a>]. Main regional geologic structures and units are plotted on top. Our study area, the Metsämalmi outcrop, is marked as a white polygon.</p>
Full article ">Figure 2
<p>(<b>a</b>) Location of the study area at Otanmäki, central Finland. Rectangles outline the areas of the different magnetic surveys. (<b>b</b>) Photos (I–IV) from the surface of the Metsämalmi outcrop show magnetite-ilmenite ore lenses and host rock (gabbro). Hammer handle length for scale is ~1 m. (<b>c</b>) UAS-borne orthophoto shows sampling locations at the Metsämalmi outcrop. (<b>d</b>) Geologic map from Metsämalmi that is provided by company Otanmäki Mine Oy (modified). Larger parts of the outcropping area are mapped as high grade (Class I) ore that contains 60–70% of magnetite and ilmenite. Red lines indicate surface measurements (i.e., magnetic susceptibilities along scan lines) performed by Otanmäki Mine Oy.</p>
Full article ">Figure 3
<p>(<b>a</b>) Aibot UAS equipped with the HSI camera Rikola. (<b>b</b>) Tholeg UAS with a standalone MagDrone fluxgate magnetometer. (<b>c</b>) Radai’s Albatros fixed-wing UAS with a fluxgate sensor in the tail.</p>
Full article ">Figure 4
<p>Principal magnetic processing workflows for UAS surveys. The workflows differ slightly, and the ELM technique is not applied on the multicopter data for this study.</p>
Full article ">Figure 5
<p>Acquisition patterns of the different magnetic surveys across the mapped ore zones (light blue polygons) in the Metsämalmi outcrop (flight lines are shown as dark blue lines): (<b>a</b>) ground survey; (<b>b</b>) 15 m multicopter; (<b>c</b>) 40 m multicopter; (<b>d</b>) 65 m multicopter; (<b>e</b>) 40 m fixed-wing magnetics.</p>
Full article ">Figure 6
<p>Results of multispectral fixed-wing survey. (<b>a</b>) CIR (color infrared) plot of camera bands 3, 2, and 1. Enhanced map pronounces outcrop ridge area, where the surface was cleaned before the survey. (<b>b</b>) RGB bands after NDVI cut and MNF transformation with using 4, 3, and 1. (<b>c</b>) Band ratio of bands 3/4 with 735/790 nm.</p>
Full article ">Figure 7
<p>Results from hyperspectral data collected from the multicopter. (<b>a</b>) RGB plot (bands 17, 7, 2) of the HSI mosaic from the eastern part of the outcrop (the inset map shows the whole mosaic). The map enhances surface details and includes an area with pronounced occurrences of gabbroic host rock and iron ore lenses. (<b>b</b>) RGB plot of eigenimages 3, 2, 1 from the same area after a MNF transformation was applied to the hyperspectral mosaic. The eigenimages of the MNF outline surfacing shapes and textures. Red and green colors highlight outcropping iron stains and blue colors are associated with area of the host rock surface and remaining soil–vegetation mixture.</p>
Full article ">Figure 8
<p>(<b>a</b>) Result of SAM mapping using input spectra from the USGS spectral library [<a href="#B49-remotesensing-11-02084" class="html-bibr">49</a>]. (<b>b</b>) Result of unsupervised k-means clustering, where two classes represent iron proxies and gabbroic rock surfaces. Different iron minerals are combined in the iron class.</p>
Full article ">Figure 9
<p>Plots from exemplary spectra that are extracted from the UAS-borne hyperspectral datasets at pixels, where individual spectra are clearly associated with specific surface materials. 20 spectra are grouped per exemplary spectrum with its standard deviation plotted in a grey shaded envelope: (<b>a</b>) Iron oxide (σ = 2.1); (<b>b</b>) Iron sulphate (σ = 2.9); (<b>c</b>) Host rock in mixture with top-soil (σ = 1.1); (<b>d</b>) Lichen remnants (σ = 3.4).</p>
Full article ">Figure 10
<p>Results of the acquired handheld spectra after sorting them by unsupervised clustering. The clusters were attributed to magnetite, iron alteration, gabbroic rock, and lichen using k-means with six input classes, and refined together with field observations at the scanned spots. Calculated mean spectra are indicated as black lines and the minimum and maximum spectral values are described by the grey envelopes.</p>
Full article ">Figure 11
<p>TMI plots from all magnetic surveys with flight heights given in AGL, blue line defines the outcrop border. (<b>a</b>) Ground survey–dashed square outlines the reference area for <a href="#remotesensing-11-02084-t006" class="html-table">Table 6</a>. (<b>b</b>) Multicopter survey at 15 m flight height, consisting of two stitched flights, seen in inset map. (<b>c</b>) Multicopter survey at 40 m AGL. (<b>d</b>) Multicopter survey at 65 m AGL. (<b>e</b>) Fixed-wing survey at 40 m AGL.</p>
Full article ">Figure 12
<p>(<b>a</b>) The total magnetic intensity extracted from the ground magnetic survey and multi-copter survey made at a flight height of 15 m AGL along a profile (see I and II in (<b>b</b>)) across the outcrop area. To better compare the datasets, the ground data has been upward-continued to 15 m. (<b>b</b>) The location of the extraction profile shown as a white line on the TMI map of the ground survey.</p>
Full article ">Figure 13
<p>(<b>a</b>) Boxplot distribution of pXRF measurements of the major compounds, with black dots represent outliers. (<b>b</b>) Biplot of the first two principal components, with PCA scores as grey dots, after transformation of the four selected compounds (rays represent measured pXRF compounds), that describe the outcrop’s major elemental compositions.</p>
Full article ">Figure 14
<p>Comparison of ground-based measurements and multi-copter magnetic data. (<b>a</b>) Point plot of iron band ratio (760/888 nm) from handheld spectroscopy and pXRF values taken along one respectively two sampling lines during the field campaign. (<b>b</b>) Point plots of magnetic susceptibility collected by the Otanmäki Mine Oy were measured along five profiles with an inline sampling distance of ~3 cm and line spacing of ~25 m. (<b>c</b>) Comparison of iron indications from SAM HSI classification (increased pixel size for better visibility) and TMI map from the multicopter (15 m AGL and 2 m grid spacing). The blue polygons in all three figures sketch the mapped ore zones.</p>
Full article ">Figure 15
<p>Scatter plots showing the relationships of different scanning methods from the outcrop, point pairs on the same location, or from the same pixel. The Pearson correlation r and significance p are given for each pair. (<b>a</b>) Correlation plot for handheld magnetic susceptibility vs. pXRF iron-oxide. (<b>b</b>) Correlation for UAS Rikola iron band ratio (760/898 nm) vs. handheld iron band ratio (760/898 nm). (<b>c</b>) Correlation for UAS Rikola iron band ratio (760/898 nm) vs. fixed-wing eBee iron band ratio (735/790 nm). (<b>d</b>) Correlation of Rikola iron band ratio (760/898 nm) vs. ground TMI. The inset plot in each bottom-right corner distributes the kernel density per variable pair.</p>
Full article ">Figure 16
<p>(<b>a</b>) SfM-MVS DSM clipped to the outcrop surface. Plotted on top are automatically extracted lineament features. Rose plot as line direction histogram shows NW–SE trending of extracted lineaments (n = 12,311). (<b>b</b>) Combined UAS-borne MSI and HSI classification (semi-transparent colors) and iron ratio results, resampled to the 2 m grid size of the 15 m multicopter TMI grid. (<b>c</b>) Integrated results of spectral and magnetic UAS-survey, giving a probability for ore occurrences, where higher values indicate alignment of detected features (a.u. = arbitrary unit). Further directional context is given by the interpreted surface lineaments, based on the automatically extracted lineaments.</p>
Full article ">
19 pages, 8012 KiB  
Article
Multispectral Multibeam Echo Sounder Backscatter as a Tool for Improved Seafloor Characterization
by Craig J. Brown, Jonathan Beaudoin, Mike Brissette and Vicki Gazzola
Geosciences 2019, 9(3), 126; https://doi.org/10.3390/geosciences9030126 - 12 Mar 2019
Cited by 75 | Viewed by 13981
Abstract
The establishment of multibeam echosounders (MBES), as a mainstream tool in ocean mapping, has facilitated integrative approaches towards nautical charting, benthic habitat mapping, and seafloor geotechnical surveys. The combined acoustic response of the seabed and the subsurface can vary with MBES operating frequency. [...] Read more.
The establishment of multibeam echosounders (MBES), as a mainstream tool in ocean mapping, has facilitated integrative approaches towards nautical charting, benthic habitat mapping, and seafloor geotechnical surveys. The combined acoustic response of the seabed and the subsurface can vary with MBES operating frequency. At worst, this can make for difficulties in merging the results from different mapping systems or mapping campaigns. However, at best, having observations of the same seafloor at different acoustic wavelengths allows for increased discriminatory power in seabed classification and characterization efforts. Here, we present the results from trials of a multispectral multibeam system (R2Sonic 2026 MBES, manufactured by R2Sonic, LLC, Austin, TX, USA) in the Bedford Basin, Nova Scotia. In this system, the frequency can be modified on a ping-by-ping basis, which can provide multi-spectral acoustic measurements with a single pass of the survey platform. The surveys were conducted at three operating frequencies (100, 200, and 400 kHz), and the resulting backscatter mosaics revealed differences in parts of the survey area between the frequencies. Ground validation surveys using a combination of underwater video transects and benthic grab and core sampling confirmed that these differences were due to coarse, dredge spoil material underlying a surface cover of mud. These innovations offer tremendous potential for application in the area of seafloor geological and benthic habitat mapping. Full article
(This article belongs to the Special Issue Geological Seafloor Mapping)
Show Figures

Figure 1

Figure 1
<p>Generalized approach for the production of benthic habitat maps, illustrating typical data sets used for this type of application. The multispectral backscatter layers (i.e., multi-frequency) afforded by the R2Sonic system offer potential advantages over conventional, monochromatic (i.e., single-frequency) systems for improved seafloor characterization.</p>
Full article ">Figure 2
<p>Overview map of the location of the 2016 and 2017 survey areas in the Bedford Basin, Nova Scotia, Canada.</p>
Full article ">Figure 3
<p>Data acquisition and post-processing of multispectral backscatter data.</p>
Full article ">Figure 4
<p>2016 multibeam echosounders (MBES) bathymetry and multispectral backscatter: 100 kHz, 200 kHz, and 400 kHz mosaics.</p>
Full article ">Figure 5
<p>2017 MBES bathymetry and multispectral backscatter: 100 kHz, 200 kHz, and 400 kHz mosaics.</p>
Full article ">Figure 6
<p>Classified video and photographic ground truthing data sets overlaid on a combined 2016 and 2017 100 kHz backscatter mosaic.</p>
Full article ">Figure 7
<p>Sediment grain size data from van veen grab samples presented as percentage gravel, sand, and mud, overlaid on the 2017 100 kHz backscatter mosaic.</p>
Full article ">Figure 8
<p>Example of benthic core samples from three adjacent sampling stations overlaid on the 2017 multispectral backscatter mosaics (100, 200, and 400k Hz). Differences in the backscatter intensities correspond to changes in core penetration depths, suggesting that the high intensity backscatter features visible in the lower frequency mosaics correspond to harder material (dredge spoil) beneath a surface covering of mud.</p>
Full article ">Figure 9
<p>Top: 2016 data set. Bottom: 2017 data set. Comparison of differences in backscatter intensities within the survey site between the multispectral mosaics. Backscatter intensities from the three multispectral mosaics are compared along the transect shown on the 100 kHz mosaic (left). Backscatter intensities are similar in the hard-substrate region at the mouth of the narrows (SE of the site), and differences are visible in deeper water that is associated with the dredge spoil deposits.</p>
Full article ">
35 pages, 58464 KiB  
Article
The Glacial Geomorphology of the Ice Cap Piedmont Lobe Landsystem of East Mýrdalsjökull, Iceland
by David J. A. Evans, Marek Ewertowski, Chris Orton and David J. Graham
Geosciences 2018, 8(6), 194; https://doi.org/10.3390/geosciences8060194 - 30 May 2018
Cited by 17 | Viewed by 6211
Abstract
A surficial geology and geomorphology map of the forelands of the Sandfellsjökull and Oldufellsjökull piedmont lobes of the east Mýrdalsjökull ice cap is used to characterise the historical and modern landscape imprint in a glacial landsystems context. This serves as a modern analogue [...] Read more.
A surficial geology and geomorphology map of the forelands of the Sandfellsjökull and Oldufellsjökull piedmont lobes of the east Mýrdalsjökull ice cap is used to characterise the historical and modern landscape imprint in a glacial landsystems context. This serves as a modern analogue for palaeoglaciological reconstructions of ice cap systems that operated outlet lobes of contrasting dynamics, but the subtle variability in process-form regimes is encoded in the geomorphology. The landsystems of the two piedmont lobes reflect significantly different process-form regimes, and hence contrasting historical glacier dynamics, despite the fact that they are nourished by the same ice cap. The Sandfellsjökull landsystem displays the diagnostic criteria for active temperate glacier operation, including arcuate assemblages of inset minor push moraines and associated flutings, kame terrace and ice-dammed lake deposits, linear sandar directed by overridden moraine arcs, and since 1945, features, such as ice-cored, pitted, and glacially pushed outwash fans that are linked to englacial esker networks representative of recession into an overdeepening. Moraine plan forms have also changed from weakly crenulated and discontinuous curvilinear ridges to sawtooth features and crevasse-squeeze ridges and till eskers in response to changing proglacial drainage conditions. The Oldufellsjökull landsystem displays subtle signatures of jökulhlaup-driven surges, including sparse and widely spaced moraine clusters that are separated by exceptionally long flutings. The subtlety of the surge imprint at Oldufellsjökull was recognised only by comparison with nearby Sandfellsjökull, suggesting that palaeo-surging has likely been under-estimated in the ancient landform record. Hence, the simple imprint of sparse and widely spaced moraine clusters that are separated by exceptionally long flutings should be included as possible surge-diagnostic criteria. Full article
(This article belongs to the Special Issue Glacial and Geomorphological Cartography)
Show Figures

Figure 1

Figure 1
<p>Location map showing the Mýrdalsjökull ice cap and its outlet lobes and key volcanic features and the jökulhlaup flood paths that have impacted on the Sandfellsjökull and Oldufellsjökull forelands. Inset map shows the extent of the central volcanic zone (stippled pattern).</p>
Full article ">Figure 2
<p>Glacier oscillations for selected snouts of Mýrdalsjökull and adjacent snouts (compiled by the Icelandic Glaciological Society). Note that Oldufellsjökull has been monitored since 1960 but no records exist for Sandfellsjökull.</p>
Full article ">Figure 3
<p>Historical glacier snout margins identified on aerial photograph archives (post 1945) and using lichenometric dating (pre 1945) by Evans et al. [<a href="#B25-geosciences-08-00194" class="html-bibr">25</a>]. Dated margins are superimposed on the aerial photograph mosaic from 2007 together with the moraine layer from the geomorphological mapping reported in this paper: (<b>a</b>) Oldufellsjökull; and, (<b>b</b>) Sandfellsjökull.</p>
Full article ">Figure 3 Cont.
<p>Historical glacier snout margins identified on aerial photograph archives (post 1945) and using lichenometric dating (pre 1945) by Evans et al. [<a href="#B25-geosciences-08-00194" class="html-bibr">25</a>]. Dated margins are superimposed on the aerial photograph mosaic from 2007 together with the moraine layer from the geomorphological mapping reported in this paper: (<b>a</b>) Oldufellsjökull; and, (<b>b</b>) Sandfellsjökull.</p>
Full article ">Figure 4
<p>Aerial photograph extracts (Landmælingar Islands) of the same area of the Oldufellsjökull snout from the 1970–1990s period of glacier oscillation. Note the development and destruction of multiple push moraines as well as the emergence of a crevasse-fill ridge between 1984 and 1994 that mimics the localized arcuate crevasse patterns. The white dash line in 1978 demarcates the 1974 surge moraine.</p>
Full article ">Figure 5
<p>Ground photographs at the northern margin of Sandfellsjökull taken in 2007 (<b>a</b>) and 1994 (<b>b</b>), showing the significant reduction in the glacier surface and the 1990s push moraine complex in the middleground.</p>
Full article ">Figure 6
<p>Surficial geology and glacial geomorphology map of the East Mýrdalsjökull outlets Sandfellsjökull and Oldufellsjökull. A larger format version of this map is available in <a href="#app1-geosciences-08-00194" class="html-app">Supplementary Information</a> where it can be downloaded and printed at a recommended A0 paper size (<a href="#app1-geosciences-08-00194" class="html-app">Supplementary Materials Figure S1</a>).</p>
Full article ">Figure 7
<p>The outermost moraines on the glacier forelands: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) of the outermost moraines on the Sandfellsjökull foreland. Crenulate recessional push moraines are superimposed on glacially overridden (fluted) moraines; (<b>b</b>) The closely spaced and locally superimposed, post-1930 push moraines on the north Sandfellsjökull foreland (ice flow was from right to left); (<b>c</b>) The outermost moraines of the Oldufellsjökull foreland, showing the three main moraine ridges and their subsidiary minor, locally superimposed ridges; and, (<b>d</b>) A section through one of the outer Sandfellsjökull moraines, with annotations showing the major boundaries between sedimentary facies of poorly-sorted sand and gravel lenses overlain by sand and gravel rich diamictons.</p>
Full article ">Figure 7 Cont.
<p>The outermost moraines on the glacier forelands: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) of the outermost moraines on the Sandfellsjökull foreland. Crenulate recessional push moraines are superimposed on glacially overridden (fluted) moraines; (<b>b</b>) The closely spaced and locally superimposed, post-1930 push moraines on the north Sandfellsjökull foreland (ice flow was from right to left); (<b>c</b>) The outermost moraines of the Oldufellsjökull foreland, showing the three main moraine ridges and their subsidiary minor, locally superimposed ridges; and, (<b>d</b>) A section through one of the outer Sandfellsjökull moraines, with annotations showing the major boundaries between sedimentary facies of poorly-sorted sand and gravel lenses overlain by sand and gravel rich diamictons.</p>
Full article ">Figure 7 Cont.
<p>The outermost moraines on the glacier forelands: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) of the outermost moraines on the Sandfellsjökull foreland. Crenulate recessional push moraines are superimposed on glacially overridden (fluted) moraines; (<b>b</b>) The closely spaced and locally superimposed, post-1930 push moraines on the north Sandfellsjökull foreland (ice flow was from right to left); (<b>c</b>) The outermost moraines of the Oldufellsjökull foreland, showing the three main moraine ridges and their subsidiary minor, locally superimposed ridges; and, (<b>d</b>) A section through one of the outer Sandfellsjökull moraines, with annotations showing the major boundaries between sedimentary facies of poorly-sorted sand and gravel lenses overlain by sand and gravel rich diamictons.</p>
Full article ">Figure 8
<p>Aerial photograph extract (NERC ARSF 2007) showing the fluted surface with scattered boulders and occasional stoss boulders in the moraine-free area of the inner Sandfellsjökull foreland.</p>
Full article ">Figure 9
<p>Aerial photograph extract (NERC ARSF 2007) and ground views of a large former subglacial cavity infill lying downflow of a bedrock cliff on the inner foreland of Sandfellsjökull. Flutings, till eskers, and angular boulders adorn the surface of the cavity infill (see Evans et al., [<a href="#B26-geosciences-08-00194" class="html-bibr">26</a>]).</p>
Full article ">Figure 10
<p>Aerial photograph extract (NERC ARSF 2007) of an area within the fluted till surface on the inner foreland of Sandfellsjökull that includes abundant geometric ridge networks (crevasse squeeze ridges) and sinuous, ice flow-parallel till ridges (till eskers; see Evans et al., [<a href="#B8-geosciences-08-00194" class="html-bibr">8</a>]).</p>
Full article ">Figure 11
<p>Aerial photograph extract (NERC ARSF 2007) of the 1980s-1990s moraine belt on the Oldufellsjökull foreland (compare with <a href="#geosciences-08-00194-f004" class="html-fig">Figure 4</a> to see the evolution of the landforms since the 1990s). The supraglacial infill of a crevasse-controlled re-entrant, created between 1984 and 1994, is circled.</p>
Full article ">Figure 12
<p>Glacifluvial deposits in narrow ribbons of terraced outwash infilling channels between arcuate overridden moraine ridges and push moraines on the Sandfellsjökull foreland; (<b>a</b>) Aerial photograph extract (NERC ARSF 2007; former ice flow from the northwest); and, (<b>b</b>) View from the summit of Sandfell, showing the isolation of elongate “islands” of till and moraine (former ice flow from the left).</p>
Full article ">Figure 13
<p>Glacifluvial features on the Oldufellsjökull eastern foreland and their relationships with the 50 m high bedrock cliff: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) showing the cliff, the LIA maximum moraine, the 934 AD jökulhlaup deposits, and inset channels demarcating early ice margin recession and containing dry waterfalls (DW); (<b>b</b>) Subglacially engorged eskers (EE) at the bedrock cliff base; and, (<b>c</b>) Superimposed foreset bedding separated by flood gravels with boulders and recording the sequential infilling and draining of an ice-dammed lake by “inwash” deltas in the valley dammed by the north edge of Oldufellsjökull (Lotte 10 and Tara Evans 12 for scale).</p>
Full article ">Figure 13 Cont.
<p>Glacifluvial features on the Oldufellsjökull eastern foreland and their relationships with the 50 m high bedrock cliff: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) showing the cliff, the LIA maximum moraine, the 934 AD jökulhlaup deposits, and inset channels demarcating early ice margin recession and containing dry waterfalls (DW); (<b>b</b>) Subglacially engorged eskers (EE) at the bedrock cliff base; and, (<b>c</b>) Superimposed foreset bedding separated by flood gravels with boulders and recording the sequential infilling and draining of an ice-dammed lake by “inwash” deltas in the valley dammed by the north edge of Oldufellsjökull (Lotte 10 and Tara Evans 12 for scale).</p>
Full article ">Figure 13 Cont.
<p>Glacifluvial features on the Oldufellsjökull eastern foreland and their relationships with the 50 m high bedrock cliff: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) showing the cliff, the LIA maximum moraine, the 934 AD jökulhlaup deposits, and inset channels demarcating early ice margin recession and containing dry waterfalls (DW); (<b>b</b>) Subglacially engorged eskers (EE) at the bedrock cliff base; and, (<b>c</b>) Superimposed foreset bedding separated by flood gravels with boulders and recording the sequential infilling and draining of an ice-dammed lake by “inwash” deltas in the valley dammed by the north edge of Oldufellsjökull (Lotte 10 and Tara Evans 12 for scale).</p>
Full article ">Figure 14
<p>Glacifluvial landforms around the margins of Sandfellsjökull: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) of extensively pitted outwash and kame and kettle topography on the south foreland, and containing chains of elongate ponds; (<b>b</b>) Aerial photograph extract (NERC ARSF 2007) showing englacial eskers emerging on the downwasting glacier surface and associated ice-contact fans; and, (<b>c</b>) Glacially overridden, pre-LIA outwash on the south foreland overlain by a gravelly till veneer and push moraines.</p>
Full article ">Figure 14 Cont.
<p>Glacifluvial landforms around the margins of Sandfellsjökull: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) of extensively pitted outwash and kame and kettle topography on the south foreland, and containing chains of elongate ponds; (<b>b</b>) Aerial photograph extract (NERC ARSF 2007) showing englacial eskers emerging on the downwasting glacier surface and associated ice-contact fans; and, (<b>c</b>) Glacially overridden, pre-LIA outwash on the south foreland overlain by a gravelly till veneer and push moraines.</p>
Full article ">Figure 14 Cont.
<p>Glacifluvial landforms around the margins of Sandfellsjökull: (<b>a</b>) Aerial photograph extract (NERC ARSF 2007) of extensively pitted outwash and kame and kettle topography on the south foreland, and containing chains of elongate ponds; (<b>b</b>) Aerial photograph extract (NERC ARSF 2007) showing englacial eskers emerging on the downwasting glacier surface and associated ice-contact fans; and, (<b>c</b>) Glacially overridden, pre-LIA outwash on the south foreland overlain by a gravelly till veneer and push moraines.</p>
Full article ">Figure 15
<p>Glacifluvial landforms on the north Sandfellsjökull foreland: (<b>a</b>) Ground view of kame terraces, deltas and shorelines on the lower slopes of the fluted bedrock ridge. Visible at the far left is the col between the bedrock ridge and the western mountain slopes; (<b>b</b>) Aerial photograph extract (NERC ARSF 2007) of the kame terraces, deltas and shorelines, showing small areas of push moraine development; and, (<b>c</b>) Small lake (“inwash”) delta in a steep-sided mountain valley to the north of the foreland.</p>
Full article ">Figure 15 Cont.
<p>Glacifluvial landforms on the north Sandfellsjökull foreland: (<b>a</b>) Ground view of kame terraces, deltas and shorelines on the lower slopes of the fluted bedrock ridge. Visible at the far left is the col between the bedrock ridge and the western mountain slopes; (<b>b</b>) Aerial photograph extract (NERC ARSF 2007) of the kame terraces, deltas and shorelines, showing small areas of push moraine development; and, (<b>c</b>) Small lake (“inwash”) delta in a steep-sided mountain valley to the north of the foreland.</p>
Full article ">Figure 15 Cont.
<p>Glacifluvial landforms on the north Sandfellsjökull foreland: (<b>a</b>) Ground view of kame terraces, deltas and shorelines on the lower slopes of the fluted bedrock ridge. Visible at the far left is the col between the bedrock ridge and the western mountain slopes; (<b>b</b>) Aerial photograph extract (NERC ARSF 2007) of the kame terraces, deltas and shorelines, showing small areas of push moraine development; and, (<b>c</b>) Small lake (“inwash”) delta in a steep-sided mountain valley to the north of the foreland.</p>
Full article ">Figure 16
<p>Aerial photograph extracts from 1984 (Landmælingar Islands) and 2007 (NERC ARSF) showing the development and collapse of push moraines in ice-cored glacifluvial deposits.</p>
Full article ">Figure 17
<p>Exposure through deposits classified as glacilacustrine deposits modified by glacifluvial processes on the inner northern Sandfellsjökull foreland. Delta foreset bedded gravels and sands are truncated and overlain by glacifluvial outwash.</p>
Full article ">
Back to TopTop