[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (456)

Search Parameters:
Keywords = USGS

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
27 pages, 16054 KiB  
Article
Mitigating Disparate Elevation Differences between Adjacent Topobathymetric Data Models Using Binary Code
by William M. Cushing and Dean J. Tyler
Remote Sens. 2024, 16(18), 3418; https://doi.org/10.3390/rs16183418 - 14 Sep 2024
Viewed by 177
Abstract
Integrating coastal topographic and bathymetric data for creating regional seamless topobathymetric digital elevation models of the land/water interface presents a complex challenge due to the spatial and temporal gaps in data acquisitions. The Coastal National Elevation Database (CoNED) Applications Project develops topographic (land [...] Read more.
Integrating coastal topographic and bathymetric data for creating regional seamless topobathymetric digital elevation models of the land/water interface presents a complex challenge due to the spatial and temporal gaps in data acquisitions. The Coastal National Elevation Database (CoNED) Applications Project develops topographic (land elevation) and bathymetric (water depth) regional scale digital elevation models by integrating multiple sourced disparate topographic and bathymetric data models. These integrated regional models are broadly used in coastal and climate science applications, such as sediment transport, storm impact, and sea-level rise modeling. However, CoNED’s current integration method does not address the occurrence of measurable vertical discrepancies between adjacent near-shore topographic and bathymetric data sources, which often create artificial barriers and sinks along their intersections. To tackle this issue, the CoNED project has developed an additional step in its integration process that collectively assesses the input data to define how to transition between these disparate datasets. This new step defines two zones: a micro blending zone for near-shore transitions and a macro blending zone for the transition between high-resolution (3 m or less) to moderate-resolution (between 3 m and 10 m) bathymetric datasets. These zones and input data sources are reduced to a multidimensional array of zeros and ones. This array is compiled into a 16-bit integer representing a vertical assessment for each pixel. This assessed value provides the means for dynamic pixel-level blending between disparate datasets by leveraging the 16-bit binary notation. Sample site RMSE assessments demonstrate improved accuracy, with values decreasing from 0.203–0.241 using the previous method to 0.126–0.147 using the new method. This paper introduces CoNED’s unique approach of using binary code to improve the integration of coastal topobathymetric data. Full article
Show Figures

Figure 1

Figure 1
<p>Spatial extent of Coastal National Elevation Database’s (CoNED’s) regional topobathymetric model (TBDEM) products for the conterminous United States (CONUS) averaging 40,000 km<sup>2</sup>. Each data series is color-coded, representing its publication year, with 12 published between 2016 and 2023.</p>
Full article ">Figure 2
<p>The top half highlights an example of an artificial sink that occurs when high-resolution topobathymetric data are merged with coarser interpolated sonar records. The missing topobathymetric data are typically a result of poor water clarity, resulting in invalid lidar returns. The red line is a cross section of the sink with its elevation profile to the top right half. The bottom half highlights an example of an artificial barrier that occurs when older topographic elevation exposes areas of shoreline recession. These also typically occur when there are voids in the topobathymetric data, exposing the older topographic data that has receded. The green line is a cross section of the barrier artifact, and to the far right is its elevation profile—satellite imagery credit to Google and Airbus.</p>
Full article ">Figure 3
<p>(<b>A</b>) is the topobathymetric elevation model method (TEMM) integration workflow [<a href="#B3-remotesensing-16-03418" class="html-bibr">3</a>]. (<b>B</b>) highlights where the new micro/macro blending is added in the TEMM integration workflow.</p>
Full article ">Figure 4
<p>A map of the St. Augustine, Florida, coast illustrating the locations of the macro blending zone (MiBZ) [black] and macro blending zone (MaBZ) [white] transition zones. The red inset map shows the MiBZ and includes hydrologic breaklines (yellow lines) defined by the high-resolution topographic digital elevation models (DEMs). The green inset map is at the same scale as the MiBZ inset but highlights the typical location and a wider MaBZ.</p>
Full article ">Figure 5
<p>In the illustration, five datasets are grouped in the topobathymetric category (CAT02) assigned priorities one through five. The stacking order is in descending order with priority five on the bottom and one at the top. The solid gray color indicates land, the speckled black is bathymetric elevations, and the white indicates no data. The final composite DEM is made of primarily the 2021 priority one dataset, but where no data exist, the lower priority data are used to fill the no-data spaces in succession. Priority five dataset is not applied in the composite because the higher priority datasets cover that layer with data. The composite still has some no-data space in the lower right corner because none of the input sources had valid elevation data seen in panel B.</p>
Full article ">Figure 6
<p>Binary pair created from a category composite digital elevation model (DEM). Panel (<b>A</b>) illustrates pixel values in a composite DEM, where the blank squares (pixels) represent no data. Panel (<b>B</b>) illustrates the binary classification for the data/no-data binary layer, where a 1 represents pixels with valid elevation and a 0 represents pixels with no data. Panel (<b>C</b>) illustrates the binary classification of elevations below or above mean sea level (MSL), where a 1 represents a pixel at or below MSL and a 0 represents both pixels above MSL and no-data pixels. Together panels (<b>B</b>,<b>C</b>) represent a binary Pair.</p>
Full article ">Figure 7
<p>Three steps in converting 16 binary layers into a single band 16-bit integer Bit-pack geospatial data layer. Step 1 is stacking the 16 binary layers based on priority into a single 16-band data array. Step 2 is compiling the 16-band data array into a 2D 16-bit integer array that represents each pixel’s z-axis binary code. Step 3 is converting the 2D array into a geospatial readable raster format with appropriate geospatial positioning.</p>
Full article ">Figure 8
<p>Illustration of the removal of a receded shoreline. Panel (<b>A</b>) shows the ghost shoreline artifact. Panel (<b>B</b>) confirms the recession of the shoreline. Panel (<b>C</b>) is the resulting Bit-pack dataset indicating interpolation in the green area. Panel (<b>D</b>) is the result of an algorithm applying the information from Bit-pack results to create an improved representation of near-shore bathymetry. The red line indicates the topographic best available shoreline, and the black line represents the prior shoreline based on an earlier surface data acquisition. This example is east of the Cape Canaveral Launch Complex 46.</p>
Full article ">Figure 9
<p>Illustration of a 16-bit binary code. The smaller text on top indicates the position of the binary code from left to right. The larger 1s and 0s are the binary switches that compose a 16-bit binary integer. To the right of the equal sign is the integer value that the sequence of 1s and 0s represents. The arrows point to the description of each position or pair of positions. As of publishing, the “Open” label under positions 9, 8, and 1, 0 are not being used for analysis but are available for future use.</p>
Full article ">Figure 10
<p>The gridded squares represent a multidimensional spatial illustration of a binary stack. This stack of binary data layers represents the 16 layers to build the Bit-pack data layer noted at the top of the diagram in color. The description of each layer is to the right of the individual layer or layer pair. To the right of the binary stack is the position number that the layer is in the binary code. The two vertical lines transecting the binary stack highlight the two vertical stacks of pixels that, when compiled, represent the value 48,184.</p>
Full article ">Figure 11
<p>Illustration of the map algebra expressions used to generate the micro blending zone. The upper left grid represents the CAT01 composite digital elevation model. The upper right grid, labeled Null Data (eq1), is the result of the first expression that identifies no-data pixels as “1” and pixels with data as “0.” An expanded expression is applied to the no data to extend the 0 values out of three pixels, replacing their respective 1 value to create the Expanded Data (eq2) grid at the bottom left. The final expression sums the no-data and Expanded Data grids, then reassigns all the pixels not equal to 1 to 0. This results in the micro blending zone, where 1s indicate the micro zone and 0s are the pixels outside the zone. Each map algebra expression is defined at the bottom with the number corresponding to the equation grid (eq).</p>
Full article ">Figure 12
<p>Sample chart comparing inverse distance weighting (IDW) interpolated profile (black dashed line), source moderate-resolution (MR) profile (blue dotted line), progressive weighted interpolation (Δ<span class="html-italic">pw</span>) profile (solid green line), and the slope weighted interpolation (SWI) profile (red dashed line).</p>
Full article ">Figure 13
<p>This series of maps and profile graph shows the change in bathymetric values inside the macro blending zone (MaBZ) from an unblended digital elevation model (DEM), inverse distance weighting (IDW) interpolation, and weighted slope interpolation (WSI). The top left map is an unblended composite DEM, and the middle map is the same composite with an IDW interpolation applied in the zone. The right map is the same composite with the WSI applied in the zone. The elevation profile chart graphs each transect, and the line color corresponds to the respective map on which the transect is located.</p>
Full article ">Figure 14
<p>Panel (<b>A</b>) shows an unblended composite digital elevation model (DEM) as context to show how and where pixels are modified during this step. Panel (<b>B</b>) shows what category composite DEMs or blending methods are applied to create a micro/macro blended DEM. CAT01, CAT02, and CAT04 indicate the use of those respective composite DEMs, WSI indicates the use of the weighted slope interpolations, INMIN indicates the use of the input minimum value method, and INZERO indicates the use of the input zero truncated surface method. Panel (<b>C</b>) shows the results of implementing the interpolation methods indicated in panel B using the micro/macro method. Panel (<b>D</b>) is an aerial image to provide context. The polygons indicate the locations of the micro (red) and macro (orange) zones. The black-hatched polygons are areas where no interpolation is applied.</p>
Full article ">Figure 15
<p>This binary code for integer 47,356 illustrates how to identify data anomalies by analyzing the sequence of ones and zeros or switches in the binary code. This code sequence reveals that the CAT02 input data are likely errant values because these data deviate from the priority current topographic input, CAT01, as well as from the lower priority inputs.</p>
Full article ">Figure 16
<p>Example for Bit-pack value range aggregation. On the left is the original Bit-pack result for a spatial extent with unique values. The right is the results of joining the classification lookup table (LUT) with the Bit-pack dataset and aggregating to the elevation interpolation classification (EIC). The EIC has a potential of eight classes, but in this example spatial extent, only five classes are indicated.</p>
Full article ">Figure 17
<p>Visual representation of the six high-level steps of applying the micro/macro interpolation blending and applying it to the final topobathymetric elevation model (TBDEM). Panel (<b>A</b>) shows a composite of the five elevation categories, and panel (<b>B</b>) shows the mask used to remove the blending zones requiring interpolation. Panel (<b>C</b>) shows the results of removing those blending zones. Panel (<b>D</b>) shows the results of the inverse distance weighting (IDW) interpolation of the blending zones. Panel (<b>E</b>) indicates where the three blending methods (input minimum value [INMIN], input zero truncated surface [INZERO], and weighted slope interpolation [WSI]) will be applied; the green shade refers to valid input data. Panel (<b>F</b>) shows applied blending to zones in the final TBDEM product.</p>
Full article ">Figure 18
<p>Illustration of the quantitative micro blending zone root mean squared error (RMSE) analysis results. The three-color shaded relief maps represent the data sources used in the RMSE analysis. The top left map represents the topobathymetric control data source, the middle map represents the unblended topobathymetric elevation modeling method (TEMM) topobathymetric elevation model (TBDEM), and the map on the right represents the micro blending method. The white horizontal hatch feature in all the maps represents the area where the micro blending occurred and is the area used to derive the RMSE values. The line segments on the maps represent the elevation profile chart below the maps. The line color in each map corresponds to the profile on the chart. The white segments on the chart are the intersection of the elevation profiles and the micro blending zone analysis. The gray areas on the chart are the segments along the profile that reflect the source elevations with no interpolation. Note that the elevation range on the y-axis is 1.5 m, well within the error range of the typical submerged topobathymetric measurements. The white patches in the control map are areas where no valid lidar point could be acquired.</p>
Full article ">Figure 19
<p>Illustration of the quantitative macro blending zone root mean squared error (RMSE) analysis results. The three-color shaded relief maps represent the data sources used in the RMSE analysis. The top left map represents the topobathymetric control data source, the middle map represents the unblended topobathymetric elevation modeling method (TEMM) topobathymetric elevation model (TBDEM), and the map on the right represents the macro blending method. The white horizontal hatch feature in all the maps represents the area where the macro blending occurred and is the area used to derive the RMSE values. The line segments on the maps represent the elevation profile chart below the maps. The line color in each map corresponds to the profile on the chart. The white segment on the chart is the intersection of the elevation profiles and the macro blending zone analysis. The gray areas on the chart are the segments along the profile that reflect the source elevations with no interpolation. Note that the elevation range on the y-axis is 2 m, well within the error range of the typical submerged topobathymetric measurements.</p>
Full article ">Figure A1
<p>This is a general high-level workflow (WF) diagram of the topobathymetric elevation modeling method (TEMM) integration component (TIC) micro/macro blending method. Each step (process or output) is notated with a WF and number (WF-N) inside a black circle. These step notations are referenced in the text to visually illustrate where in the workflow each process and output occurs.</p>
Full article ">
33 pages, 11680 KiB  
Article
A Spatial–Seasonal Study on the Danube River in the Adjacent Danube Delta Area: Case Study—Monitored Heavy Metals
by Catalina Topa, Gabriel Murariu, Valentina Calmuc, Madalina Calmuc, Maxim Arseni, Cecila Serban, Carmen Chitescu and Lucian Georgescu
Water 2024, 16(17), 2490; https://doi.org/10.3390/w16172490 - 2 Sep 2024
Viewed by 567
Abstract
Monitoring and protecting flowing watercourses is a complex and challenging task that requires the collaboration and coordination of various stakeholders such as governments, industries, farmers, consumers and environmental groups. The study of the dynamics of the concentration of polluting factors and especially the [...] Read more.
Monitoring and protecting flowing watercourses is a complex and challenging task that requires the collaboration and coordination of various stakeholders such as governments, industries, farmers, consumers and environmental groups. The study of the dynamics of the concentration of polluting factors and especially the concentrations of heavy metals and highlighting a seasonal variation is a necessary element from this point of view. In this article, we present the results of our analyses carried out in two measurement campaigns executed in 10 monitoring points along the Danube River, between Braila city and Isaccea city in the pre-deltaic area, during the summer season and autumn season 2022. The importance of this area is given by the fact that the Danube Delta is part of the UNESCO heritage, and the monitoring of polluting factors is a necessity in the desire to protect this area. The data measured during the July and August 2022 campaign cover a wide range of chemical species: Phosphate, CCO, CBO5, NH4+, N-NO2, N-NO3, N-Total, P-PO4 3−, SO42−, Cl, phenols, as well as metals with a harmful effect: Al, As, Cd, Cr, Fe. The study includes an evaluation based on the statistical approach of the results to highlight the significant correlations and differences identified between the two data sets. Next, to highlight the obtained results, a numerical model was considered using HEC-RAS and ESRI ArcGIS applications in a two-dimensional unsteady flow model in order to obtain the non-homogenous concentrations’ distributions in the studied area. These two-dimensional models have been less studied in the specialized literature. In this way, interesting results could be obtained, and prediction methods regarding the dynamics of metal concentrations could be structured. The data obtained were used for the terrain model from the USGS service, and the flows of the Danube and its two tributaries were simulated using the data provided by the national services. In this work, we present the results obtained for the dynamics of the concentrations of the metals Al, As, Cd, Cr and Fe and the evaluation of the specific absorption coefficients for the explanation and correlation with the results of the measurements. Except for the numerical model presented, we would like to highlight the existence of some contributions of the main tributaries of the Danube in the study area. Such a systematic study has not been carried out due to conditions imposed by the border authorities. From this point of view, this study has an element of originality. The study is part of a more complex project in which the spatio-temporal distribution of the polluting factors in the water was evaluated, and the habitats in the study area were inventoried—especially those of community interest. In this way, we were able to expose the self-purification capacity of the Danube and highlight the existence of a concentration reduction gradient along the course of the river. The aspects related to the influence of the distribution of polluting factors on the state of health will be the subject of another article. Full article
Show Figures

Figure 1

Figure 1
<p>The map of monitoring points configuration and a broader regional map.</p>
Full article ">Figure 2
<p>USGS database terrain map for hydrodynamic model.</p>
Full article ">Figure 3
<p>The HEC-RAS hydrological terrain map with transversal cross sections for numerical model.</p>
Full article ">Figure 4
<p>Scatter plot representation for heavy metals and phenols in the two measurement campaigns in the Danube River study area: (<b>a</b>) scatterplot representation for Al concentration seasonally measured in the 10 monitoring points; (<b>b</b>) scatter plot representation for As concentration seasonally measured in the 10 monitoring points; (<b>c</b>) scatter plot representation for the Cd concentration seasonally measured in the 10 monitoring points; (<b>d</b>) scatter plot representation for the Cr concentration seasonally measured in the 10 monitoring points.</p>
Full article ">Figure 5
<p>Scatter plot representation for the two measurement campaigns in the Danube River study area: (<b>a</b>) scatter plot representation for phosphonate concentration seasonally measured in the 10 monitoring points; (<b>b</b>) scatter plot representation for CCO concentration seasonally measured in the 10 monitoring points; (<b>c</b>) scatter plot representation for CBO<sub>5</sub> concentration seasonally measured in the 10 monitoring points; (<b>d</b>) scatter plot representation for NH<sub>4</sub> concentration seasonally measured in the 10 monitoring points.</p>
Full article ">Figure 6
<p>Scatter plot representation for the two measurement campaigns in the Danube River study area: (<b>a</b>) scatter plot representation for the N-NO<sub>2</sub> concentration seasonally measured in the 10 monitoring points; (<b>b</b>) scatter plot representation for the N-NO<sub>2</sub> concentration seasonally measured in the 10 monitoring points; (<b>c</b>) scatter plot representation for the N-total concentration seasonally measured in the 10 monitoring points; (<b>d</b>) scatter plot representation for the P-PO<sub>4</sub> concentration seasonally measured in the 10 monitoring points.</p>
Full article ">Figure 7
<p>Scatter plot representation for the two measurement campaigns in the Danube River study area: (<b>a</b>) scatter plot representation for SO<sub>4</sub> concentration seasonally measured in the 10 monitoring points; (<b>b</b>) scatter plot representation for Cl concentration seasonally measured in the 10 monitoring points; (<b>c</b>) scatter plot representation for phenols concentration seasonally measured in the 10 monitoring points; (<b>d</b>) scatter plot representation for Fe concentration seasonally measured in the 10 monitoring points.</p>
Full article ">Figure 8
<p>PCA representation of the two measurement campaigns in the Danube River study area: (<b>a</b>) PCA representation with main factors 1 and 2 for the July campaign; (<b>b</b>) PCA representation with main factors 1 and 2 for the October campaign; (<b>c</b>) PCA representation with main factors 1 and 3 for the July campaign; (<b>d</b>) PCA representation with main factors 1 and 3 for the October campaign; (<b>e</b>) PCA representation with main factors 1 and 4 for the July campaign; (<b>f</b>) PCA representation with main factors 1 and 4 for the October campaign.</p>
Full article ">Figure 9
<p>(<b>a</b>) Tree Diagram for monitored Variables—Weighted-pair group centroid (median)—Euclidean distances—July monitoring campaign; (<b>b</b>) Tree Diagram for monitored Variables—Weighted-pair group centroid (median)—Euclidean distances—October monitoring campaign.</p>
Full article ">Figure 10
<p>PCA representation of the two measurement campaigns in the Danube River study area: (<b>a</b>) PCA representation with main factors 1 and 2 for the July campaign; (<b>b</b>) PCA representation with main factors 1 and 3 for the July campaign; (<b>c</b>) PCA representation with main factors 1 and 2 for the October campaign; (<b>d</b>) PCA representation with main factors 1 and 3 for the October campaign.</p>
Full article ">Figure 11
<p>The graph representation of the monthly evolution of the Danube flow in the monitored area—for the non-stationary study.</p>
Full article ">Figure 12
<p>The representation of distribution maps: (<b>a</b>) representation of distribution maps for Fe concentration—July campaign; (<b>b</b>) representation of distribution maps for Fe concentration—October campaign.</p>
Full article ">Figure 13
<p>The representation of distribution maps: (<b>a</b>) representation of distribution maps for Cr concentration—July campaign; (<b>b</b>) representation of distribution maps for Cr concentration—October campaign.</p>
Full article ">Figure 14
<p>The representation of distribution maps: (<b>a</b>) representation of distribution maps for Cd concentration—July campaign; (<b>b</b>) representation of distribution maps for Cd concentration—October campaign.</p>
Full article ">Figure 15
<p>The representation of distribution maps: (<b>a</b>) representation of distribution maps for As concentration—July campaign; (<b>b</b>) representation of distribution maps for As concentration—October campaign.</p>
Full article ">Figure 16
<p>The representation of distribution maps: (<b>a</b>) representation of distribution maps for Al concentration—July campaign; (<b>b</b>) representation of distribution maps for Al concentration—October campaign.</p>
Full article ">
15 pages, 5798 KiB  
Article
Recognition of Artificial Gases Formed during Drill-Bit Metamorphism Using Advanced Mud Gas
by Janaina Andrade de Lima Leon, Henrique Luiz de Barros Penteado, Geoffrey S. Ellis, Alexei Milkov and João Graciano Mendonça Filho
Energies 2024, 17(17), 4383; https://doi.org/10.3390/en17174383 - 2 Sep 2024
Viewed by 596
Abstract
Drill-bit metamorphism (DBM) is the process of thermal degradation of drilling fluid at the interface of the bit and rock due to the overheating of the bit. The heat generated by the drill when drilling into a rock formation promotes the generation of [...] Read more.
Drill-bit metamorphism (DBM) is the process of thermal degradation of drilling fluid at the interface of the bit and rock due to the overheating of the bit. The heat generated by the drill when drilling into a rock formation promotes the generation of artificial hydrocarbon and non-hydrocarbon gas, changing the composition of the gas. The objective of this work is to recognize and evaluate artificial gases originating from DBM in wells targeting oil accumulations in pre-salt carbonates in the Santos Basin, Brazil. For the evaluation, chromatographic data from advanced mud gas equipment, drilling parameters, drill type, and lithology were used. The molar concentrations of gases and gas ratios (especially ethene/ethene+ethane and dryness) were analyzed, which identified the occurrence of DBM. DBM is most severe when wells penetrate igneous and carbonate rocks with diamond-impregnated drill bits. The rate of penetration, weight on bit, and rotation per minute were evaluated together with gas data but did not present good correlations to assist in identifying DBM. The depth intervals over which artificial gases formed during DBM are recognized should not be used to infer pay zones or predict the composition and properties of reservoir fluids because the gas composition is completely changed. Full article
(This article belongs to the Topic Advances in Oil and Gas Wellbore Integrity)
Show Figures

Figure 1

Figure 1
<p>Regional map of the Santos Basin, showing the location of wells that were drilled using advanced mud gas analysis.</p>
Full article ">Figure 2
<p>Generalized process of advanced mud gas extraction and analysis (modified from Ablard et al., 2012 [<a href="#B3-energies-17-04383" class="html-bibr">3</a>]). The schematic illustrates well drilling and mud circulation, positioning of mud extraction probes at the IN and OUT along the mudflow line, and subsequent analysis of the gas inside the mudlogging unit by gas chromatograph and mass spectrometer.</p>
Full article ">Figure 3
<p>Mud gas logs for well E3 are divided into four parts including formation tops, lithology, and gas ratios (Well-E3 a panel), ratios C<sub>2</sub>/C<sub>1</sub>, dryness, and ethene/(ethane+ethene) (Well-E3 b panel), gas chromatography (Well-E3 c panel), and normalized alkanes (Well-E3 d panel). See <a href="#energies-17-04383-t002" class="html-table">Table 2</a> for mnemonics of lithological types. Interval with drill-bit metamorphism marked with green arrow. From 5500 until the end of the well, changes were observed in the gas curves, mainly in the igneous rock interval caused by drill-bit metamorphism. In Well-E3 b, we observed an increase in the C<sub>2</sub>/C<sub>1</sub> curve and a decrease in dryness causing the inversion of these two curves. In Well-E3 c, an increase in C<sub>2</sub> is also observed, overlapping C<sub>1</sub> from 5500 m to the end of the well, and in Well-E3 d, the relative percentage of ethane is greater than that of methane depending on the increase in ethylene.</p>
Full article ">Figure 4
<p>Mud gas logs for well D3 are divided into four parts including formation tops, lithology, and gas ratios (Well-D3 a panel), ratios C<sub>2</sub>/C<sub>1</sub>, dryness, and ethene/(ethane+ethene) (Well-D3 b panel), and gas chromatography (Well-D3 c panel) and normalized alkanes (Well-D3 d panel). See <a href="#energies-17-04383-t002" class="html-table">Table 2</a> for mnemonics of lithological types.</p>
Full article ">Figure 5
<p>The panel is separated into three different wells. Well-B2 (c, d, and b), Well-C2 (c, d, and b), and Well-D5 (c, d, and b) are represented on all the graphs that identify the drill-bit metamorphism in wells B2, C2, and D5. For the three wells, the chromatographic distribution graphs of alkanes (Well-B2 c, Well-C2 c, and Well-D5 c), the concentration of normalized alkanes from C<sub>1</sub> to C<sub>5</sub> (Well-B2 d, Well-C2 d, and Well-D5 d), and ratios (Well-B2 b, Well-C2 b, and Well-D5 b) were evaluated. Comparison between the gas chromatography of wells B2 (Well-B2 c—without drill-bit metamorphism until 5918 m and with drill-bit metamorphism when started the igneous rock), well C2 (Well-C2 c—with drill-bit metamorphism in the interval below 5700 m after changing from PDC to impregnated drill), and well D5 (Well-D5 c—with drill-bit metamorphism throughout the well drilled with the impregnated drill). Interval with drill-bit metamorphism marked with green arrow.</p>
Full article ">Figure 6
<p>Correlations between the drilling parameters and the gas ratios that were used for the identification of DBM considering the groups of wells with and without DBM, separated by lithology (ROP x C<sub>2</sub>/C<sub>1</sub>, WOB x ethene/ethene+ethane, and RPM x dryness).</p>
Full article ">
32 pages, 14893 KiB  
Article
Mapping of Clay Montmorillonite Abundance in Agricultural Fields Using Unmixing Methods at Centimeter Scale Hyperspectral Images
by Etienne Ducasse, Karine Adeline, Audrey Hohmann, Véronique Achard, Anne Bourguignon, Gilles Grandjean and Xavier Briottet
Remote Sens. 2024, 16(17), 3211; https://doi.org/10.3390/rs16173211 - 30 Aug 2024
Viewed by 524
Abstract
The composition of clay minerals in soils, and more particularly the presence of montmorillonite (as part of the smectite family), is a key factor in soil swell–shrinking as well as off–road vehicle mobility. Detecting these topsoil clay minerals and quantifying the montmorillonite abundance [...] Read more.
The composition of clay minerals in soils, and more particularly the presence of montmorillonite (as part of the smectite family), is a key factor in soil swell–shrinking as well as off–road vehicle mobility. Detecting these topsoil clay minerals and quantifying the montmorillonite abundance are a challenge since they are usually intimately mixed with other minerals, soil organic carbon and soil moisture content. Imaging spectroscopy coupled with unmixing methods can address these issues, but the quality of the estimation degrades the coarser the spatial resolution is due to pixel heterogeneity. With the advent of UAV-borne and proximal hyperspectral acquisitions, it is now possible to acquire images at a centimeter scale. Thus, the objective of this paper is to evaluate the accuracy and limitations of unmixing methods to retrieve montmorillonite abundance from very-high-resolution hyperspectral images (1.5 cm) acquired from a camera installed on top of a bucket truck over three different agricultural fields, in Loiret department, France. Two automatic endmember detection methods based on the assumption that materials are linearly mixed, namely the Simplex Identification via Split Augmented Lagrangian (SISAL) and the Minimum Volume Constrained Non-negative Matrix Factorization (MVC-NMF), were tested prior to unmixing. Then, two linear unmixing methods, the fully constrained least square method (FCLS) and the multiple endmember spectral mixture analysis (MESMA), and two nonlinear unmixing ones, the generalized bilinear method (GBM) and the multi-linear model (MLM), were performed on the images. In addition, several spectral preprocessings coupled with these unmixing methods were applied in order to improve the performances. Results showed that our selected automatic endmember detection methods were not suitable in this context. However, unmixing methods with endmembers taken from available spectral libraries performed successfully. The nonlinear method, MLM, without prior spectral preprocessing or with the application of the first Savitzky–Golay derivative, gave the best accuracies for montmorillonite abundance estimation using the USGS library (RMSE between 2.2–13.3% and 1.4–19.7%). Furthermore, a significant impact on the abundance estimations at this scale was in majority due to (i) the high variability of the soil composition, (ii) the soil roughness inducing large variations of the illumination conditions and multiple surface scatterings and (iii) multiple volume scatterings coming from the intimate mixture. Finally, these results offer a new opportunity for mapping expansive soils from imaging spectroscopy at very high spatial resolution. Full article
(This article belongs to the Special Issue Remote Sensing for Geology and Mapping)
Show Figures

Figure 1

Figure 1
<p>Site locations from AGEOTHYP project depicted with colored squares on: (<b>a</b>) topographic map by IGN (National Institute of Geographic and Forest Information) overlaid with smectite abundance from XRD analyses and (<b>b</b>) BRGM swelling hazard map. Soil digital photos of the three selected sites: (<b>c</b>) “Le Buisson” located in Coinces, (<b>d</b>) “Les Laps” located in Gémigny and (<b>e</b>) “La Malandière” located in Mareau.</p>
Full article ">Figure 2
<p>Acquisition setup with the HySpex cameras, RGB composite image from HySpex VNIR camera on Gémigny, Coinces and Mareau sites, with the sampling grid composed of 15 subzones (named after “SUB”), samples collected for laboratory soil characterization in subzones are delimited by red squares (<b>right</b>).</p>
Full article ">Figure 3
<p>NDVI and CAI values for the Mareau hyperspectral image. In red: the thresholds chosen for each index in order to characterize four classes.</p>
Full article ">Figure 4
<p>Grain size and SOC for each site (<b>left</b>), texture triangle for all samples (<b>right</b>).</p>
Full article ">Figure 5
<p>Processing scheme to estimate montmorillonite abundance.</p>
Full article ">Figure 6
<p>Endmembers from laboratory spectral libraries: (<b>a</b>) montmorillonite, (<b>b</b>) kaolinite, (<b>c</b>) illite, (<b>d</b>) quartz and (<b>e</b>) calcite.</p>
Full article ">Figure 7
<p>EM estimates over the Gémigny image. Comparison of the detected and Ducasse EM spectra and graphs of mixture simplex in the first two components space (PC 1 and PC 2) for (<b>a</b>) SISAL to detect 4 EM, (<b>b</b>) SISAL to detect 5 EM, (<b>c</b>) MVC-NMF to detect 4 EM and (<b>d</b>) MVC-NMF to detect 5 EM.</p>
Full article ">Figure 8
<p>Montmorillonite abundance estimations over all the subzones per site (gray boxplots with the median highlighted by a red line) compared to the XRD dataset (boxplots with a red square depicting the median). The inputs are the USGS library, the six preprocessings and REF followed by MLM.</p>
Full article ">Figure 9
<p>Montmorillonite abundance estimations over all the subzones per site (gray boxplots with the median highlighted by a red line) compared to the XRD dataset (boxplots with a red square depicting the median). The inputs are the Ducasse library, the six preprocessings and REF followed by MLM.</p>
Full article ">Figure 10
<p>Performances of Montmorillonite abundance estimations (wt%) obtained with (<b>a</b>) REF-MLM and (<b>b</b>) 1stSGD-MLM with the USGS library (red) and Ducasse spectral library (blue). Bars in the x axis correspond to the accuracy of XRD analysis, and bars in the y axis correspond to the standard deviation of estimated montmorillonite abundances.</p>
Full article ">Figure 11
<p>Results on Gémigny-SUB14: (<b>a</b>) RGB image (in black: masked areas), (<b>b</b>) hillshade map, (<b>c</b>) hillshade histogram (the red vertical line represents the median), (<b>d</b>) difference between the estimated montmorillonite abundance map obtained with REF-MLM and the XRD measured value (in white: masked areas), (<b>g</b>) the same for 1stSGD-MLM, (<b>e</b>) <span class="html-italic">p</span> value maps for REF-MLM (in white: masked areas), (<b>h</b>) the same for 1stSGD-MLM, (<b>f</b>) <span class="html-italic">p</span> value histogram for REF-MLM (the red vertical line represents the median) and (<b>i</b>) the same for 1stSGD-MLM.</p>
Full article ">Figure 12
<p>Results on Coinces-SUB2: (<b>a</b>) RGB image (in black: masked areas), (<b>b</b>) hillshade map, (<b>c</b>) hillshade histogram (the red vertical line represents the median), (<b>d</b>) difference between the estimated montmorillonite abundance map obtained with REF-MLM and the XRD measured value (in white: masked areas), (<b>g</b>) the same for 1stSGD-MLM, (<b>e</b>) <span class="html-italic">p</span> value maps for REF-MLM (in white: masked areas), (<b>h</b>) the same for 1stSGD-MLM, (<b>f</b>) <span class="html-italic">p</span> value histogram for REF-MLM (the red vertical line represents the median) and (<b>i</b>) the same for 1stSGD-MLM.</p>
Full article ">Figure 13
<p>Performances for Montmorillonite abundance estimation with REF-MLM for all subsites (gray boxplots with the median highlighted by a red line) plotted with the XRD dataset (boxplots with a red square depicting the median).</p>
Full article ">Figure 14
<p>Maps for Gémigny site (<b>a</b>) RGB composite image, (<b>b</b>) composite mask and (<b>c</b>) abundance map of montmorillonite obtained with the REF-MLM and USGS library.</p>
Full article ">Figure 15
<p>Maps for Coinces with wet area SUB10 site (<b>a</b>) RGB composite image, (<b>b</b>) composite mask and (<b>c</b>) abundance map of montmorillonite obtained with the REF-MLM and USGS library.</p>
Full article ">Figure 16
<p>Maps for Mareau site with wet area SUB15 (<b>a</b>) RGB composite image, (<b>b</b>) composite mask and (<b>c</b>) abundance map of montmorillonite obtained with the REF-MLM and USGS library.</p>
Full article ">Figure 17
<p>Comparison between mineral abundance estimations with REF-MLM and USGS library and the XRD dataset for each site: (<b>a</b>) Coinces, (<b>b</b>) Gémigny, (<b>c</b>) Mareau.</p>
Full article ">
12 pages, 252 KiB  
Article
Somatic Structure and Ultrasound Parameters of the Calcaneus Bone in Combat Sports Athletes in Relation to Vitamin D3 Levels
by Janusz Brudecki, Łukasz Rydzik, Wojciech Wąsacz, Pavel Ruzbarsky, Wojciech Czarny, Marlena Warowna and Tadeusz Ambroży
J. Clin. Med. 2024, 13(16), 4960; https://doi.org/10.3390/jcm13164960 - 22 Aug 2024
Viewed by 365
Abstract
Background/Objectives: Physical activity is widely recognized for its beneficial effects on bone density during adolescence, which could lead to enhanced bone density in later life, thus acting as a health-promoting activity with long-lasting implications. However, not all studies are conclusive regarding the [...] Read more.
Background/Objectives: Physical activity is widely recognized for its beneficial effects on bone density during adolescence, which could lead to enhanced bone density in later life, thus acting as a health-promoting activity with long-lasting implications. However, not all studies are conclusive regarding the type, intensity, duration, and frequency of the most effective physical activities. This study focuses on combat sports athletes and examines the relationship between their somatic build and heel bone parameters using ultrasound (USG) and their vitamin D3 levels. Methods: The study included 40 male athletes specializing in various combat sports. The measurements of body height, body mass, skinfold thickness, and bone widths at multiple sites were performed to estimate the somatic build. The USG parameters of the heel bone and the blood levels of vitamin D3 were also recorded. Statistical significance was determined using one-way ANOVA, with differences among sports disciplines also examined. Results: The study found significant differences in the body composition and USG bone parameters among athletes from different combat sports (p ≤ 0.05). The calcaneus stiffness index (SI) and speed of sound (SOS) were significantly higher in athletes with normal vitamin D3 levels compared to those with below-normal levels (p = 0.0015 and p = 0.001, respectively). These findings suggest that vitamin D3 may influence bone stiffness and density. Conclusions: The study underscores the importance of maintaining adequate vitamin D3 levels to support bone mineralization in athletes, particularly those training indoors with limited exposure to sunlight. It also highlights the potential of using USG as a non-invasive method to assess bone health, aiding in the optimization of training programs to prevent injuries and improve performance. Full article
(This article belongs to the Section Sports Medicine)
26 pages, 12993 KiB  
Article
Comparative Analysis of Satellite-Based Precipitation Data across the CONUS and Hawaii: Identifying Optimal Satellite Performance
by Saurav Bhattarai and Rocky Talchabhadel
Remote Sens. 2024, 16(16), 3058; https://doi.org/10.3390/rs16163058 - 20 Aug 2024
Viewed by 539
Abstract
Accurate precipitation estimates are crucial for various hydrological and environmental applications. This study presents a comprehensive evaluation of three widely used satellite-based precipitation datasets (SPDs)—PERSIANN, CHIRPS, and MERRA—and a monthly reanalysis dataset—TERRA—that include data from across the contiguous United States (CONUS) and Hawaii, [...] Read more.
Accurate precipitation estimates are crucial for various hydrological and environmental applications. This study presents a comprehensive evaluation of three widely used satellite-based precipitation datasets (SPDs)—PERSIANN, CHIRPS, and MERRA—and a monthly reanalysis dataset—TERRA—that include data from across the contiguous United States (CONUS) and Hawaii, at daily, monthly, and yearly timescales. We present the performance of these SPDs using ground-based observations maintained by the USGS (United States Geological Survey). We employ evaluation metrics, such as the coefficient of determination (R2), root mean square error (RMSE), mean square error (MSE), and mean absolute error (MAE), to identify optimal SPDs. Our findings reveal that MERRA outperforms PERSIANN and CHIRPS on a daily scale, while CHIRPS is the best-performing dataset on a monthly scale. However, all datasets show limitations in accurately estimating absolute amount of precipitation totals. The spatial analysis highlights regional variations in the datasets’ performance, with MERRA consistently performing well across most regions, while CHIRPS and PERSIANN show strengths in specific areas and months. We also observe a consistent seasonal pattern in the performance of all datasets. This study contributes to the growing body of knowledge on satellite precipitation estimates and their applications, guiding the selection of suitable datasets based on the required temporal resolution and regional context. As such SPDs continue to evolve, ongoing evaluation and improvement efforts are crucial to enhance their reliability and support informed decision-making in various fields, including water resource management, agricultural planning, and climate studies. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Spatial distribution of the United States Geological Survey (USGS) precipitation gauges across the study area.</p>
Full article ">Figure 2
<p>Overall methodology used in this study.</p>
Full article ">Figure 3
<p>Heatmap showing statistical summary of satellite-based precipitation data performance metrics across the United States for daily aggregated timeseries. Each panel represents a different metric: (<b>A</b>) coefficient of determination (R<sup>2</sup>), (<b>B</b>) root mean square error (RMSE), (<b>C</b>) mean square error (MSE), and (<b>D</b>) mean absolute error (MAE). The <span class="html-italic">y</span>-axis lists the SPDs (MERRA, CHIRPS, PERSIANN), and the <span class="html-italic">x</span>-axis shows the mean, median, max, and min of each statistic.</p>
Full article ">Figure 4
<p>Spatial representation of the best-performing satellite precipitation dataset at ground station locations across the United States, based on (<b>A</b>) coefficient of regression (R<sup>2</sup>), (<b>B</b>) root mean square error (RMSE), (<b>C</b>) mean square error (MSE), and (<b>D</b>) mean average error (MAE) at aggregated daily timeseries. In the maps, blue indicates MERRA, violet/pink indicates PERSIANN, and red indicates CHIRPS.</p>
Full article ">Figure 5
<p>Heatmap showing statistical summary of satellite-based precipitation data performance metrics across the United States for monthly aggregated timeseries. Each panel represents a different metric: (<b>A</b>) coefficient of determination (R<sup>2</sup>), (<b>B</b>) root mean square error (RMSE), (<b>C</b>) mean square error (MSE), and (<b>D</b>) mean absolute error (MAE). The <span class="html-italic">y</span>-axis lists the SPDs (MERRA, CHIRPS, PERSIANN), and the <span class="html-italic">x</span>-axis shows the mean, median, max, and min of each statistic.</p>
Full article ">Figure 6
<p>Spatial representation of the best-performing satellite precipitation dataset at ground station locations across the United States, based on (<b>A</b>) coefficient of regression (R<sup>2</sup>), (<b>B</b>) root mean square error (RMSE), (<b>C</b>) mean square error (MSE), and (<b>D</b>) mean average error (MAE) at aggregated monthly timeseries. In the maps, blue indicates MERRA, violet/pink indicates PERSIANN, red indicates CHIRPS, and black indicates CRU (TERRA).</p>
Full article ">Figure 7
<p>Monthly performance analysis of satellite precipitation datasets (PERSIANN, CHIRPS, and MERRA) at the daily scale. The left column displays box plots illustrating the monthly variability in R<sup>2</sup> values for each satellite dataset (PERSIANN, CHIRPS, and MERRA, from top to bottom), revealing consistent seasonal patterns, with performance peaking in April and November/December and reaching a minimum in July. The right column features heatmaps depicting the spatial distribution of R<sup>2</sup> values across all stations for each month, corroborating the temporal trends observed in the box plots. MERRA demonstrates superior performance compared to PERSIANN and CHIRPS, with higher median R<sup>2</sup> values and a greater prevalence of higher R<sup>2</sup> values across stations. Blank spots in the heatmaps indicate missing or insufficient data for calculating R<sup>2</sup>. The color scheme progresses from violet (lower R<sup>2</sup>) to green to yellow (higher R<sup>2</sup>).</p>
Full article ">Figure 8
<p>Spatial representation of the best-performing satellite precipitation dataset (MERRA, CHIRPS, or PERSIANN) for each station and month. The color-coding scheme assigns blue to MERRA, red to CHIRPS, and violet to PERSIANN. The 3 × 4 matrix of subplots covers all months from January to December, revealing the spatial and temporal variability in the performance of the satellite datasets. MERRA (blue) consistently appears as the best-performing dataset across most stations and months, confirming its overall superiority.</p>
Full article ">Figure 9
<p>Best satellite data Source by site (extreme comparison) for daily aggregated timeseries based on the (<b>A</b>) coefficient of determination (R<sup>2</sup>), (<b>B</b>) root mean square error (RMSE), (<b>C</b>) mean square error (MSE), and (<b>D</b>) mean absolute error (MAE). The upper row shows the results for precipitation extremes below the 10th percentile, and the lower row shows the results for precipitation extremes above the 90th percentile. The colors indicate the best-performing datasets: blue for MERRA, red for CHIRPS, and violet for PERSIANN.</p>
Full article ">Figure A1
<p>Spatial distribution of the coefficient of determination (R<sup>2</sup>) values for CHIRPS, PERSIANN, and MERRA satellite precipitation datasets at daily timeframe across the United States.</p>
Full article ">Figure A2
<p>Spatial distribution of the root mean square error (RMSE) values for CHIRPS, PERSIANN, and MERRA satellite precipitation datasets at daily timeframe across the United States.</p>
Full article ">Figure A3
<p>Spatial distribution of the mean square error (MSE) values for CHIRPS, PERSIANN, and MERRA satellite precipitation datasets at daily timeframe across the United States.</p>
Full article ">Figure A4
<p>Spatial distribution of the mean absolute error (MAE) values for CHIRPS, PERSIANN, and MERRA satellite precipitation datasets at daily timeframe across the United States.</p>
Full article ">Figure A5
<p>Spatial distribution of the coefficient of determination (R<sup>2</sup>) values for CHIRPS, PERSIANN, and MERRA satellite precipitation datasets at monthly timeframe across the United States.</p>
Full article ">Figure A6
<p>Spatial distribution of the root mean square error (RMSE) values for CHIRPS, PERSIANN, and MERRA satellite precipitation datasets at monthly timeframe across the United States.</p>
Full article ">Figure A7
<p>Spatial distribution of the mean square error (MSE) values for CHIRPS, PERSIANN, TERRA(CRU) and MERRA satellite precipitation datasets at monthly timeframe across the United States.</p>
Full article ">Figure A8
<p>Spatial distribution of the mean absolute error (MAE) values for CHIRPS, PERSIANN, TERRA(CRU), and MERRA satellite precipitation datasets at monthly timeframe across the United States.</p>
Full article ">Figure A9
<p>Spatial distribution of the coefficient of determination (R<sup>2</sup>) values for CHIRPS, PERSIANN, TERRA(CRU) and MERRA satellite precipitation datasets at yearly timeframe across the United States.</p>
Full article ">Figure A10
<p>Spatial distribution of the root mean square error (RMSE) values for CHIRPS, PERSIANN, TERRA(CRU)and MERRA satellite precipitation datasets at yearly timeframe across the United States.</p>
Full article ">Figure A11
<p>Spatial distribution of the mean square error (MSE) values for CHIRPS, PERSIANN, TERRA(CRU) and MERRA satellite precipitation datasets at yearly timeframe across the United States.</p>
Full article ">Figure A12
<p>Spatial distribution of the mean absolute error (MAE) values for CHIRPS, PERSIANN, TERRA(CRU), and MERRA satellite precipitation datasets at yearly timeframe across the United States.</p>
Full article ">Figure A13
<p>Heatmap showing statistical summary of satellite precipitation data performance metrics across the United States for yearly aggregated timeseries.</p>
Full article ">Figure A14
<p>Spatial representation of the best-performing satellite precipitation dataset at ground station locations across the United States, based on (<b>A</b>) coefficient of regression(R<sup>2</sup>), (<b>B</b>) root mean square error (RMSE), (<b>C</b>) mean square error (MSE), (<b>D</b>) mean average error (MAE) at aggregated annual timeseries.</p>
Full article ">
22 pages, 23123 KiB  
Article
Geological Study Based on Multispectral and Hyperspectral Remote Sensing: A Case Study of the Mahuaping Beryllium–Tungsten Deposit Area in Shangri-La
by Yunfei Hu, Zhifang Zhao, Xinle Zhang, Lunxin Feng, Yang Qin, Liu Ouyang and Ziqi Huang
Sustainability 2024, 16(15), 6387; https://doi.org/10.3390/su16156387 - 25 Jul 2024
Viewed by 702
Abstract
This study applied Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) multispectral data and ZY1-02D hyperspectral data to map the structural distribution and hydrothermal alteration in the polymetallic ore district in southern Shangri-La City, Yunnan Province, China. The study area hosts several polymetallic [...] Read more.
This study applied Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) multispectral data and ZY1-02D hyperspectral data to map the structural distribution and hydrothermal alteration in the polymetallic ore district in southern Shangri-La City, Yunnan Province, China. The study area hosts several polymetallic deposits, including the Mahuaping tungsten–beryllium deposit, which has significant mineral exploration potential. The deposit type is mainly magmatic–hydrothermal, with average grades of 0.41% WO3 and 0.22% BeO, and substantial reserves, prominently controlled by faults. Based on this, this study employed ASTER data for the visual interpretation of structures through false-color composites combined with DEM data. Additionally, ASTER and ZY1-02D data were processed using the principal component analysis and spectral angle mapper methods to extract anomalies related to tungsten mineralization such as carbonate alteration, sericitization, chloritization, and hematization of the hydrothermal origin. The results indicated that the structural trends in the study area predominantly align in north–south and northeast directions, with alteration anomalies concentrated in the central and fold areas. Our analysis of typical deposits revealed their close association with north–south faults and east–west joints, as well as the enrichment level of alteration anomalies, identifying five high-potential target areas for mineral exploration. Further evaluation involved field validation through the spectral scanning of samples, field verification, and a comparison with known lithology. These assessments confirmed that the spectral curves matched those in the USGS database, the structural interpretations aligned with the field observations (84% accuracy from 25 sampling points, with 21 matching extracted alteration types), and the alteration results corresponded well with the lithological units, indicating high accuracy in alteration extraction. Finally, a comparative discussion highlighted that the results derived from ZY1-02D data were more applicable to the local area. The outcomes of this study can support subsequent mineral exploration efforts, enhancing the sustainability of important mineral resources. Full article
(This article belongs to the Special Issue Sustainability in Mineral Potential Mapping of Key Mineral Resources)
Show Figures

Figure 1

Figure 1
<p>Regional geology and tectonics of the study area: (<b>a</b>) Shangri-La region tectonics [<a href="#B29-sustainability-16-06387" class="html-bibr">29</a>]. (<b>b</b>) Mahuaping region tectonics [<a href="#B30-sustainability-16-06387" class="html-bibr">30</a>].</p>
Full article ">Figure 2
<p>Geological sketch map of Mahuaping area. (<b>a</b>) Location of the northwestern part of Yunnan Province, China; (<b>b</b>) Location of the study area; (<b>c</b>) Geological sketch map of Mahuaping area.</p>
Full article ">Figure 3
<p>Technical roadmap.</p>
Full article ">Figure 4
<p>Preprocessed remote sensing images: (<b>a</b>) ASTER Image 321. (<b>b</b>) ZY1-02D Image 29-19-10.</p>
Full article ">Figure 5
<p>Structural interpretation map of linear and circular structures in the study area.</p>
Full article ">Figure 6
<p>Spectral curves: (<b>a</b>) measured spectral curve; (<b>b</b>) measured spectral curve resampled to ASTER; (<b>c</b>) measured spectral curve resampled to ZY1-02D; (<b>d</b>) resampled spectral curve from 500 nm to 900 nm; (<b>e</b>) resampled spectral curve from 2100 nm to 2400 nm.</p>
Full article ">Figure 7
<p>Extraction results of mineral alteration anomalies in Mahuaping mining area: (<b>a</b>) sericite alteration; (<b>b</b>) carbonate alteration; (<b>c</b>) chlorite alteration; (<b>d</b>) iron staining.</p>
Full article ">Figure 8
<p>Extraction of alteration anomalies using spectral angle method. (<b>a</b>) Sericite; (<b>b</b>) calcite; (<b>c</b>) chlorite; (<b>d</b>) limonite.</p>
Full article ">Figure 9
<p>Distribution map of prospective areas for mineralization prediction.</p>
Full article ">Figure 10
<p>(<b>a</b>) Measured spectral curves of alteration characteristic minerals; (<b>b</b>) USGS spectral curves of alteration characteristic minerals; (<b>c</b>) measured spectral curves in the 2100~2300 nm range; (<b>d</b>) USGS spectral curves in the 2100~2300 nm range; (<b>e</b>) measured spectral curves in the 2250~2450 nm range; (<b>f</b>) USGS spectral curves in the 2250~2450 nm range.</p>
Full article ">Figure 11
<p>Field survey route map.</p>
Full article ">Figure 12
<p>Field collection sample photos: (<b>a</b>) Sample 3—sericite feldspar schist; (<b>b</b>) Sample 6—carbonate-altered schist; (<b>c</b>) Sample 18—lazurite, scheelite; (<b>d</b>) scheelite fluorescence reaction; (<b>e</b>) Sample 27—marble; (<b>f</b>) Sample 37—sericite chlorite schist.</p>
Full article ">Figure 13
<p>Microscopic identification photos of field samples in the study area (<b>a</b>) Sample 16—Gn (galena) 2.5× (reflected light); (<b>b</b>) Sample 17—Hm (hematite), Py (pyrite) 2.5× (reflected light); (<b>c</b>) Sample 18—Wf (scheelite) 2.5× (reflected light); (<b>d</b>) Sample 21—Ber (beryl) 2.5× (single polarization); (<b>e</b>) Sample 28—Qtz (quartz) 2.5× (cross-polarized light); (<b>f</b>) Sample 28—Cal (calcite) 2.5× (single polarization); (<b>g</b>) Sample 33—Bit (biotite) 2.5× (single polarization); (<b>h</b>) Sample 38—Ser (sericite) 5× (cross-polarized light).</p>
Full article ">Figure 14
<p>Comparison of alteration results and stratigraphy.</p>
Full article ">
13 pages, 4244 KiB  
Article
Correction of Thin Cirrus Absorption Effects in Landsat 8 Thermal Infrared Sensor Images Using the Operational Land Imager Cirrus Band on the Same Satellite Platform
by Bo-Cai Gao, Rong-Rong Li, Yun Yang and Martha Anderson
Sensors 2024, 24(14), 4697; https://doi.org/10.3390/s24144697 - 19 Jul 2024
Viewed by 481
Abstract
Data from the Operational Land Imager (OLI) and the Thermal Infrared Sensor (TIRS) instruments onboard the Landsat 8 and Landsat 9 satellite platforms are subject to contamination by cloud cover, with cirrus contributions being the most difficult to detect and mask. To help [...] Read more.
Data from the Operational Land Imager (OLI) and the Thermal Infrared Sensor (TIRS) instruments onboard the Landsat 8 and Landsat 9 satellite platforms are subject to contamination by cloud cover, with cirrus contributions being the most difficult to detect and mask. To help address this issue, a cirrus detection channel (Band 9) centered within the 1.375-μm water vapor absorption region was implemented on OLI, with a spatial resolution of 30 m. However, this band has not yet been fully utilized in the Collection 2 Landsat 8/9 Level 2 surface temperature data products that are publicly released by U.S. Geological Survey (USGS). The temperature products are generated with a single-channel algorithm. During the surface temperature retrievals, the effects of absorption of infrared radiation originating from the warmer earth’s surfaces by ice clouds, typically located in the upper portion of the troposphere and re-emitting at much lower temperatures (approximately 220 K), are not taken into consideration. Through an analysis of sample Level 1 TOA and Level 2 surface data products, we have found that thin cirrus cloud features present in the Level 1 1.375-μm band images are directly propagated down to the Level 2 surface data products. The surface temperature errors resulting from thin cirrus contamination can be 10 K or larger. Previously, we reported an empirical and effective technique for removing thin cirrus scattering effects in OLI images, making use of the correlations between the 1.375-μm band image and images of any other OLI bands located in the 0.4–2.5 μm solar spectral region. In this article, we describe a variation of this technique that can be applied to the thermal bands, using the correlations between the Level 1 1.375-μm band image and the 11-μm BT image for the effective removal of thin cirrus absorption effects. Our results from three data sets acquired over spatially uniform water surfaces and over non-uniform land/water boundary areas suggest that if the cirrus-removed TOA 11-μm band BT images are used for the retrieval of the Level 2 surface temperature (ST) data products, the errors resulting from thin cirrus contaminations in the products can be reduced to about 1 K for spatially diffused cirrus scenes. Full article
(This article belongs to the Section Remote Sensors)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) A Landsat 8 OLI RGB image (Path: 014; Row: 033) acquired over eastern coastal area of Maryland State, USA, on 1 July 2018, (<b>B</b>) the corresponding Band 9 (cirrus band) image, (<b>C</b>) the surface temperature image, and (<b>D</b>) the false-color-coded surface temperature image.</p>
Full article ">Figure 2
<p>(<b>A</b>) A Landsat 8 OLI RGB image (Path: 001; Row: 085) acquired on 5 March 2014, over Pacific Ocean west of the coastal area of Chile, (<b>B</b>) the corresponding Band 9 (cirrus band) image, (<b>C</b>) the Level 2 OLI RGB surface reflectance (SR) image, and (<b>D</b>) the false-color-coded water surface temperature image with an attached color bar.</p>
Full article ">Figure 3
<p>Vertical temperature profiles for the Tropical and U.S. Standard (1976) model atmospheres.</p>
Full article ">Figure 4
<p>A diagram illustrating that the IR radiances originating from Earth’s surface are absorbed, scattered, and re-emitted at lower temperatures by cirrus clouds at high altitudes (approximately 10 km) and that the downward solar radiation at 1.375-μm is scattered backup by cirrus clouds and the transmitted 1.375-μm radiance through cirrus is absorbed by atmospheric water vapor beneath cirrus clouds.</p>
Full article ">Figure 5
<p>(<b>A</b>) A Landsat 8 OLI RGB image (Path: 001; Row: 085) acquired on 21 March 2014, over Pacific Ocean west of the coastal area of Chile, (<b>B</b>) the corresponding Band 9 (cirrus) image, (<b>C</b>) the scatter plot of 1.375-μm versus 0.86-μm band images, (<b>D</b>) the cirrus-corrected RGB image, (<b>E</b>) the black/white 11-μm BT image, (<b>F</b>) the false-colored version of 11-μm BT image, (<b>G</b>) the scatter plot of 1.375-μm band TOA reflectance image versus the 11-μm band TOA BT image, and (<b>H</b>) the 11-μm band BT image after the removal of cirrus absorption effects.</p>
Full article ">Figure 6
<p>(<b>A</b>) A portion of Landsat 8 OLI RGB image (Path: 014; Row: 034) acquired on 17 April 2014, over eastern coastal area of Maryland, USA, (<b>B</b>) the corresponding Band 9 (cirrus band) image, (<b>C</b>) the false-colored version of 11-μm band BT image, and (<b>D</b>) the 11-μm band BT image after the removal of cirrus absorption effects.</p>
Full article ">Figure 7
<p>(<b>A</b>) A portion of a Landsat 8 OLI RGB image (Path: 190; Row: 019) acquired on 11 August 2015 over the Baltic Sea, (<b>B</b>) the Band 9 (cirrus band) image, (<b>C</b>) the false-colored version of 11-μm band BT image, (<b>D</b>) the 11-μm band BT image after the removal of cirrus absorption effects, (<b>E</b>) the horizontal BT profiles along the red-colored line in (<b>C</b>) (bottom curve), and the cirrus-corrected BT profile along the same line in (<b>D</b>) (top curve).</p>
Full article ">
25 pages, 5510 KiB  
Article
Exploring PCSWMM for Large Mixed Land Use Watershed by Establishing Monitoring Sites to Evaluate Stream Water Quality
by Mohd Sohib Ansari, Suresh Sharma, Felicia P. Armstrong, Mark Delisio and Sahar Ehsani
Hydrology 2024, 11(7), 104; https://doi.org/10.3390/hydrology11070104 - 15 Jul 2024
Viewed by 793
Abstract
Extensive hydrologic and water quality modeling within a watershed benefits from long-term flow and nutrient data sets for appropriate model calibration and validation. However, due to a lack of local water quality data, simpler water quality modeling techniques are generally adopted. In this [...] Read more.
Extensive hydrologic and water quality modeling within a watershed benefits from long-term flow and nutrient data sets for appropriate model calibration and validation. However, due to a lack of local water quality data, simpler water quality modeling techniques are generally adopted. In this study, the monitoring sites were established at two different locations to collect hydraulic data for the hydraulic calibration and validation of the model. In addition, water quality samples were collected at eight monitoring sites and analyzed in the lab for various parameters for calibration. This includes total suspended solids (TSS), soluble phosphorus, five-day biochemical oxygen demand (BOD5), and dissolved oxygen (DO). The Personal Computer Storm Water Management Model (PCSWMM) 7.6 software was used to simulate all the pollutant loads using event mean concentrations (EMCs). The performance of the model for streamflow calibration at the two USGS gauging stations was satisfactory, with Nash–Sutcliffe Efficiency (NSE) values ranging from 0.51 to 0.54 and coefficients of determination (R2) ranging from 0.71 to 0.72. The model was also validated with the help of historical flow data with NSE values ranging from 0.5 to 0.79, and R2 values ranging from 0.6 to 0.95. The hydraulic calibration also showed acceptable results with reasonable NSE and R2 values. The water quality data recorded at the monitoring stations were then compared with the simulated water quality modeling results. The model reasonably simulated the water quality, which was evaluated through visual inspection using a scatter plot. Our analysis showed that the upstream tributaries, particularly from agricultural areas, were contributing more pollutants than the downstream tributaries. Overall, this study demonstrates that the PCSWMM, which was typically used for modeling urban watersheds, could also be used for modeling larger mixed land use watersheds with reasonable accuracy. Full article
(This article belongs to the Section Hydrological and Hydrodynamic Processes and Modelling)
Show Figures

Figure 1

Figure 1
<p>Map of Mill Creek watershed, consisting of water quality sampling stations, HOBO loggers’ location, and USGS gauge stations for PCSWMM model development.</p>
Full article ">Figure 2
<p>PCSWMM model streamflow calibration (1999–2000) at 2 USGS gauge stations: (<b>a</b>) USGS gauge 03098513 (Outlet), and (<b>b</b>) USGS gauge 03098500.</p>
Full article ">Figure 3
<p>PCSWMM model streamflow validation (1950–1970) at USGS gauge 03098500 from (<b>a</b>–<b>i</b>) 2/11/1950–2/27/1950 to 12/10/1970–12/26/1970.</p>
Full article ">Figure 3 Cont.
<p>PCSWMM model streamflow validation (1950–1970) at USGS gauge 03098500 from (<b>a</b>–<b>i</b>) 2/11/1950–2/27/1950 to 12/10/1970–12/26/1970.</p>
Full article ">Figure 4
<p>PCSWMM model hydraulic model calibration: (<b>a</b>) station 1 at the outlet of the watershed, (<b>b</b>) station 2 located at the East Golf hike trail.</p>
Full article ">Figure 4 Cont.
<p>PCSWMM model hydraulic model calibration: (<b>a</b>) station 1 at the outlet of the watershed, (<b>b</b>) station 2 located at the East Golf hike trail.</p>
Full article ">Figure 5
<p>PCSWMM model streamflow depth validation: (<b>a</b>) station 1 at the outlet of the watershed, (<b>b</b>) station 2 located at the East Golf hike trail.</p>
Full article ">Figure 6
<p>EMC values from different literature sources: (<b>a</b>) BOD<sub>5</sub>, (<b>b</b>) soluble phosphorus, (<b>c</b>) TSS.</p>
Full article ">Figure 6 Cont.
<p>EMC values from different literature sources: (<b>a</b>) BOD<sub>5</sub>, (<b>b</b>) soluble phosphorus, (<b>c</b>) TSS.</p>
Full article ">Figure 7
<p>Water quality calibration at the monitoring station 14 for the period from 2022 to 2023: (<b>a</b>) TSS (<b>b</b>) BOD<sub>5</sub> (<b>c</b>) DO (<b>d</b>) soluble phosphorus.</p>
Full article ">Figure 7 Cont.
<p>Water quality calibration at the monitoring station 14 for the period from 2022 to 2023: (<b>a</b>) TSS (<b>b</b>) BOD<sub>5</sub> (<b>c</b>) DO (<b>d</b>) soluble phosphorus.</p>
Full article ">Figure 8
<p>Water quality validation at the monitoring station 14 for the period from 2017 to 2018: (<b>a</b>) TSS, (<b>b</b>) BOD<sub>5</sub>, (<b>c</b>) DO, (<b>d</b>) soluble phosphorus.</p>
Full article ">Figure 8 Cont.
<p>Water quality validation at the monitoring station 14 for the period from 2017 to 2018: (<b>a</b>) TSS, (<b>b</b>) BOD<sub>5</sub>, (<b>c</b>) DO, (<b>d</b>) soluble phosphorus.</p>
Full article ">Figure 9
<p>(<b>a</b>) Observed concentration of TSS in the stream for the period from 2022 to 2023, (<b>b</b>) simulated concentration of TSS by the model for the period from 2022 to 2023 when the precipitation depth was greater than 0.5 inches.</p>
Full article ">Figure 10
<p>(<b>a</b>) Observed concentration of DO in the stream for the period from 2022 to 2023, (<b>b</b>) simulated concentration of DO by the model for the period from 2022 to 2023 when the volume of precipitation was greater than 0.5 inches.</p>
Full article ">Figure 11
<p>(<b>a</b>) Observed concentration of BOD<sub>5</sub> in the stream for the period from 2022 to 2023, (<b>b</b>) simulated concentration of BOD<sub>5</sub> by the model for the period from 2022 to 2023 when the volume of precipitation was greater than 0.5 inches.</p>
Full article ">Figure 12
<p>(<b>a</b>) Observed concentration of soluble phosphorus in the stream for the period from 2022 to 2023, (<b>b</b>) simulated concentration of soluble phosphorus by the model for the period from 2022 to 2023 when the volume of precipitation was greater than 0.5 inches.</p>
Full article ">
21 pages, 5659 KiB  
Article
Estimating Brazilian Amazon Canopy Height Using Landsat Reflectance Products in a Random Forest Model with Lidar as Reference Data
by Pedro V. C. Oliveira, Hankui K. Zhang and Xiaoyang Zhang
Remote Sens. 2024, 16(14), 2571; https://doi.org/10.3390/rs16142571 - 13 Jul 2024
Viewed by 665
Abstract
Landsat data have been used to derive forest canopy structure, height, and volume using machine learning models, i.e., giving computers the ability to learn from data and make decisions and predictions without being explicitly programmed, with training data provided by ground measurement or [...] Read more.
Landsat data have been used to derive forest canopy structure, height, and volume using machine learning models, i.e., giving computers the ability to learn from data and make decisions and predictions without being explicitly programmed, with training data provided by ground measurement or airborne lidar. This study explored the potential use of Landsat reflectance and airborne lidar data as training data to estimate canopy heights in the Brazilian Amazon forest and examined the impacts of Landsat reflectance products at different process levels and sample spatial autocorrelation on random forest modeling. Specifically, this study assessed the accuracy of canopy height predictions from random forest regression models impacted by three different Landsat 8 reflectance product inputs (i.e., USGS level 1 top of atmosphere reflectance, USGS level 2 surface reflectance, and NASA nadir bidirectional reflectance distribution function (BRDF) adjusted reflectance (NBAR)), sample sizes, training/test split strategies, and geographic coordinates. In the establishment of random forest regression models, the dependent variable (i.e., the response variable) was the dominant canopy heights at a 90 m resolution derived from airborne lidar data, while the independent variables (i.e., the predictor variables) were the temporal metrics extracted from each Landsat reflectance product. The results indicated that the choice of Landsat reflectance products had an impact on model accuracy, with NBAR data yielding more trustful results than the other products despite having higher RMSE values. Training and test split strategy also affected the derived model accuracy metrics, with the random sample split (randomly distributed training and test samples) showing inflated accuracy compared to the spatial split (training and test samples spatially set apart). Such inflation was induced by the spatial autocorrelation that existed between training and test data in the random split. The inclusion of geographic coordinates as independent variables improved model accuracy in the random split strategy but not in the spatial split, where training and test samples had different geographic coordinate ranges. The study highlighted the importance of data processing levels and the training and test split methods in random forest modeling of canopy height. Full article
(This article belongs to the Special Issue Lidar for Forest Parameters Retrieval)
Show Figures

Figure 1

Figure 1
<p>Location of the 20 airborne lidar transects (red bars) within Acre state in the Brazilian Amazon biome and the 17,110 × 110 km HLS tiles intersecting these transects. Transects A, B, and C are evaluated in detail in the Results section.</p>
Full article ">Figure 2
<p>The main (pre-)processes to model dominant canopy height using L8L1, L8L2, and HLS L30 as independent variables. B04, B05, B06, and B07 are Landsat 8 red, NIR, SWIR 1, and SWIR 2 spectral bands.</p>
Full article ">Figure 3
<p>RF-predicted versus airborne lidar ground-truth DHs for Landsat 8 level 1 (L8L1), Landsat 8 level 2 (L8L2), and HLS L30 products. In the first and second rows, 80% of the randomly sampled observations from the 20 airborne lidar transects were used for training and 20% for testing the RF (80%-20%). In the third and fourth rows, the sampled observations from 16 transects were used for training and from 4 for testing the RF (16-4 spatial split). Latitude and longitude were not included (Lat/Lon: No) in the first and third rows and were included (Lat/Lon: Yes) in the second and fourth rows. All the RFs used a 50% sample size. The color gradient from dark blue to yellow shows increasing concentration of data points. Dotted line is 1:1 line.</p>
Full article ">Figure 4
<p>Sensitivity analysis of RF-dominant canopy height prediction root mean squared error (RMSE) sensitivity analysis to sample size for Landsat 8 level 1 (L8L1), Landsat 8 level 2 (L8L2), and HLS L30 products. In the left column, 80% of the randomly sampled observations from the 20 airborne lidar transects were used for training and 20% for testing the RF (80%-20%). In the right column, the sampled observations from 16 transects were used for training and 4 for testing the RF (16-4 spatial split). Latitude and longitude were not included (Lat/Lon: No) in the first row and were included (Lat/Lon: Yes) in the second row. Each dot is the median root mean squared error (RMSE), and error bars are 95% confidence interval.</p>
Full article ">Figure 5
<p>Sensitivity analysis of RF-dominant canopy height prediction to sample size for L8L1, L2L1, and HLS L30 products. In the left column, 80% of the sampled observations from the 20 airborne lidar transects were used for training and 20% for testing the RF (random split 80%-20%). In the right column, the sampled observations from 16 transects were used for training and 4 for testing the RF (spatial split 16-4). Latitude and longitude were not included (Lat/Lon: No) in the first row and were included (Lat/Lon: Yes) in the second row. Each dot is the median mean error (ME), and error bars are 95% confidence interval.</p>
Full article ">Figure 6
<p>Spatial autocorrelation on the average residuals of 300 RF models using L8L1, L8L2, and HLS L30 products at Transects A (top row), B (middle row), and C (bottom row) (see <a href="#remotesensing-16-02571-f001" class="html-fig">Figure 1</a>). The sample size of the RF models was 50%. Results are displayed in terms of training/test strategy and inclusion or not of latitude and longitude as independent variables. Latitude and longitude were not included (Lat/Lon: No) in the first and second rows and were included (Lat/Lon: Yes) in the third and fourth rows. Each lag indicated a distance corresponding to a 90 m pixel times the lag value (i.e., 90–1080 m range).</p>
Full article ">Figure 7
<p>Variable importance for RF models using L8L1, L8L2, and HLS L30. The error bars are the standard deviation of each variable in the 300 RF models. The sample size was 50%. Results were divided by split strategy (i.e., 80%-20% random or 16-4 spatial). When included as independent variables in the RF, latitude is represented by <span class="html-italic">y</span> and longitude by <span class="html-italic">x</span>.</p>
Full article ">Figure 8
<p>Spatial autocorrelation intrinsic to L8L1 temporal metrics product. In the left column are the Moran’s <span class="html-italic">I</span> correlograms showing the difference between the L8L2 and L8L1 (L2-L1) red, NIR, and SWIR bands. In the middle column is the solar zenith angle (SZA) and in the right column is the view zenith angle (VZA) red, NIR, and SWIR band correlograms. Here, only correlograms representing the 50th percentile (P50) of temporal metrics for Transect A are included. Each lag indicates a distance corresponding to a 90 m pixel times the lag value (i.e., 90–2700 m range).</p>
Full article ">Figure 9
<p>Spatial distribution of the residuals (i.e., predicted−ground truth) of the RF models within Transect A, Transect B, and Transect C. Transects have a 90 m pixel size.</p>
Full article ">
9 pages, 2884 KiB  
Comment
Comment on Yu et al. Land Surface Temperature Retrieval from Landsat 8 TIRS—Comparison between Radiative Transfer Equation-Based Method, Split Window Algorithm and Single Channel Method. Remote Sens. 2014, 6, 9829–9852
by Almustafa Abd Elkader Ayek and Bilel Zerouali
Remote Sens. 2024, 16(14), 2514; https://doi.org/10.3390/rs16142514 - 9 Jul 2024
Viewed by 691
Abstract
Accurate land surface temperature (LST) retrieval from satellite data is pivotal in environmental monitoring and scientific research. This study addresses the impact of variability in the effective wavelengths used for LST retrieval from the Thermal Infrared Sensor (TIRS) data of Landsat 8. We [...] Read more.
Accurate land surface temperature (LST) retrieval from satellite data is pivotal in environmental monitoring and scientific research. This study addresses the impact of variability in the effective wavelengths used for LST retrieval from the Thermal Infrared Sensor (TIRS) data of Landsat 8. We conduct a detailed analysis comparing the effective wavelengths reported by Yu et al. (2014) and those derived from data provided by the USGS. Our analysis reveals significant variability in the effective wavelengths for bands 10 and 11 of Landsat 8. By applying Planck’s Law and utilizing the K1 and K2 coefficients available in the metadata of Landsat 8 products, we derive the effective wavelengths for bands 10 and 11. We also rederive the effective wavelength by integrating the spectral response function of the TIRS1 sensor. Our findings indicate that the effective wavelength for band 10 is 10.814 μm, aligning with the USGS data, while the effective wavelength for band 11 is 12.013 μm. We discuss the implications of these corrected effective wavelengths on the accuracy of LST retrieval algorithms, particularly the single channel algorithm (SC) and the radiative transfer equation (RT) employed by Yu et al. The importance of using precise effective wavelengths in satellite-based temperature retrieval is emphasized, to ensure the reliability and consistency of results. This analysis underscores the critical role of accurate spectral calibration parameters in remote sensing studies and provides valuable insights in the field of land surface temperature retrieval from Landsat 8 TIRS data. Full article
Show Figures

Figure 1

Figure 1
<p>TIRS relative spectral response.</p>
Full article ">Figure 2
<p>Spectral response function for Landsat 8 product (B10).</p>
Full article ">Figure 3
<p>Standard deviation for Landsat 8 product (B10).</p>
Full article ">Figure 4
<p>Spectral response function for Landsat 8 product (B11).</p>
Full article ">Figure 5
<p>Standard deviation for Landsat 8 product (B11).</p>
Full article ">Figure 6
<p>The relative spectral response function for band 10 using cubic interpolation.</p>
Full article ">Figure 7
<p>The relative spectral response function for band 11 using cubic interpolation.</p>
Full article ">
19 pages, 5541 KiB  
Article
Application of Normalized Radar Backscatter and Hyperspectral Data to Augment Rangeland Vegetation Fractional Classification
by Matthew Rigge, Brett Bunde, Kory Postma, Simon Oliver and Norman Mueller
Remote Sens. 2024, 16(13), 2315; https://doi.org/10.3390/rs16132315 - 25 Jun 2024
Viewed by 1151
Abstract
Rangeland ecosystems in the western United States are vulnerable to climate change, fire, and anthropogenic disturbances, yet classification of rangeland areas remains difficult due to frequently sparse vegetation canopies that increase the influence of soils and senesced vegetation, the overall abundance of senesced [...] Read more.
Rangeland ecosystems in the western United States are vulnerable to climate change, fire, and anthropogenic disturbances, yet classification of rangeland areas remains difficult due to frequently sparse vegetation canopies that increase the influence of soils and senesced vegetation, the overall abundance of senesced vegetation, heterogeneity of life forms, and limited ground-based data. The Rangeland Condition Monitoring Assessment and Projection (RCMAP) project provides fractional vegetation cover maps across western North America using Landsat imagery and artificial intelligence from 1985 to 2023 at yearly time-steps. The objectives of this case study are to apply hyperspectral data from several new data streams, including Sentinel Synthetic Aperture Radar (SAR) and Earth Surface Mineral Dust Source Investigation (EMIT), to the RCMAP model. We run a series of five tests (Landsat-base model, base + SAR, base + EMIT, base + SAR + EMIT, and base + Landsat NEXT [LNEXT] synthesized from EMIT) over a difficult-to-classify region centered in southwest Montana, USA. Our testing results indicate a clear accuracy benefit of adding SAR and EMIT data to the RCMAP model, with a 7.5% and 29% relative increase in independent accuracy (R2), respectively. The ability of SAR data to observe vegetation height allows for more accurate classification of vegetation types, whereas EMIT’s continuous characterization of the spectral response boosts discriminatory power relative to multispectral data. Our spectral profile analysis reveals the enhanced classification power with EMIT is related to both the improved spectral resolution and representation of the entire domain as compared to legacy Landsat. One key finding is that legacy Landsat bands largely miss portions of the electromagnetic spectrum where separation among important rangeland targets exists, namely in the 900–1250 nm and 1500–1780 nm range. Synthesized LNEXT data include these gaps, but the reduced spectral resolution compared to EMIT results in an intermediate 18% increase in accuracy relative to the base run. Here, we show the promise of enhanced classification accuracy using EMIT data, and to a smaller extent, SAR. Full article
Show Figures

Figure 1

Figure 1
<p>Study area used to analyze impact of Synthetic Aperture Radar (SAR) and Earth Surface Mineral Dust Source Investigation (EMIT) hyperspectral data in rangeland vegetation classification. Base image is a 50th percentile composite of 2016 Landsat imagery. Inset map shows study location in the conterminous United States.</p>
Full article ">Figure 2
<p>Scatterplots of independent validation (n = 399) for selected component cover predictions by model run. Line of best fit and 1-to-1 line indicated by dashed blue and solid red line, respectively. For improved visualization, the x and y ranges vary by plot. SAR = Synthetic Aperture Radar; EMIT = Earth Surface Mineral Dust Source Investigation; LNEXT = Landsat NEXT.</p>
Full article ">Figure 3
<p>Component predictions by model for shrub (<b>top</b>) and herbaceous cover (<b>bottom</b>). White indicates either land cover or Earth Surface Mineral Dust Source Investigation (EMIT) masking. SAR = Synthetic Aperture Radar.</p>
Full article ">Figure 4
<p>Correlation (<span class="html-italic">r</span>) between Rangeland Condition Monitoring Assessment and Projection (RCMAP) high-resolution training data and Synthetic Aperture Radar (SAR) horizontal transmit and vertical receive (VH)/vertical transmit and vertical receive (VV) data at <span class="html-italic">n</span> = 60,000 pixels.</p>
Full article ">Figure 5
<p>Average spectral profiles of important rangeland targets (see Methods Section for classification details) based on high-resolution Rangeland Condition Monitoring Assessment and Projection (RCMAP) data. Line data reflect Earth Surface Mineral Dust Source Investigation (EMIT) profiles, and points of the same color represent Landsat.</p>
Full article ">Figure 6
<p>Separability of rangeland classes plotted in <a href="#remotesensing-16-02315-f005" class="html-fig">Figure 5</a> as measured by the standard deviation among spectral profiles across the classes at each Earth Surface Mineral Dust Source Investigation (EMIT) band. Width of Landsat band ranges (bands 2–7 used in Rangeland Condition Monitoring Assessment and Projection [RCMAP] analysis) are plotted in black.</p>
Full article ">Figure 7
<p>Average spectral profile from Earth Surface Mineral Dust Source Investigation (EMIT) by shrub 5% cover bins (color) in Rangeland Condition Monitoring Assessment and Projection (RCMAP) high-resolution training sites. Some bins are omitted for clarity. We removed pixels with &gt;0% tree cover and ≥20% herbaceous cover from this analysis. All species of shrub are included.</p>
Full article ">
17 pages, 8249 KiB  
Article
Development and Validation of an Ultrasonography-Based Machine Learning Model for Predicting Outcomes of Bruxism Treatments
by Kaan Orhan, Gokhan Yazici, Merve Önder, Cengiz Evli, Melek Volkan-Yazici, Mehmet Eray Kolsuz, Nilsun Bağış, Nihan Kafa and Fehmi Gönüldaş
Diagnostics 2024, 14(11), 1158; https://doi.org/10.3390/diagnostics14111158 - 31 May 2024
Viewed by 571
Abstract
Background and Objectives: We aimed to develop a predictive model for the outcome of bruxism treatments using ultrasonography (USG)-based machine learning (ML) techniques. This study is a quantitative research study (predictive modeling study) in which different treatment methods applied to bruxism patients are [...] Read more.
Background and Objectives: We aimed to develop a predictive model for the outcome of bruxism treatments using ultrasonography (USG)-based machine learning (ML) techniques. This study is a quantitative research study (predictive modeling study) in which different treatment methods applied to bruxism patients are evaluated through artificial intelligence. Materials and Methods: The study population comprised 102 participants with bruxism in three treatment groups: Manual therapy, Manual therapy and Kinesio Tape or Botulinum Toxin-A injection. USG imaging was performed on the masseter muscle to calculate muscle thickness, and pain thresholds were evaluated using an algometer. A radiomics platform was utilized to handle imaging and clinical data, as well as to perform a subsequent radiomics statistical analysis. Results: The area under the curve (AUC) values of all machine learning methods ranged from 0.772 to 0.986 for the training data and from 0.394 to 0.848 for the test data. The Support Vector Machine (SVM) led to excellent discrimination between bruxism and normal patients from USG images. Radiomics characteristics in pre-treatment ultrasound scans of patients, showing coarse and nonuniform muscles, were associated with a greater chance of less effective pain reduction outcomes. Conclusions: This study has introduced a machine learning model using SVM analysis on ultrasound (USG) images for bruxism patients, which can detect masseter muscle changes on USG. Support Vector Machine regression analysis showed the combined ML models can also predict the outcome of the pain reduction. Full article
(This article belongs to the Section Machine Learning and Artificial Intelligence in Diagnostics)
Show Figures

Figure 1

Figure 1
<p>Volume of interest (red area) outlined on the ultrasonographic image of the masseter muscle that was used for machine learning and radiomics analysis. Yellow markers indicate the edge of masseter muscle.</p>
Full article ">Figure 2
<p>USG at baseline and 4 weeks after BT-A therapy. Ultrasonographic images of the masseter muscle before (<b>A</b>) and after (<b>B</b>) botulinum toxin injection. As seen from measurements listed in the text boxes superimposed on the images, muscle volume decreased after treatment, which indicates that the treatment was successful. Yellow Lines indicates the measurement of the muscles from different region.</p>
Full article ">Figure 3
<p>Receiver operating characteristic curve analysis results of the ability of the six classifiers to differentiate between bruxism patients and normal subjects for the training set (<b>a</b>) and test set (<b>b</b>) datasets. The diagonal line represents random chance.</p>
Full article ">Figure 3 Cont.
<p>Receiver operating characteristic curve analysis results of the ability of the six classifiers to differentiate between bruxism patients and normal subjects for the training set (<b>a</b>) and test set (<b>b</b>) datasets. The diagonal line represents random chance.</p>
Full article ">
19 pages, 7091 KiB  
Article
Analysis of Vibration Characteristics of Bridge Structures under Seismic Excitation
by Ling’ai Li and Shengxiang Huang
Entropy 2024, 26(6), 465; https://doi.org/10.3390/e26060465 - 29 May 2024
Viewed by 500
Abstract
Bridges may undergo structural vibration responses when exposed to seismic waves. An analysis of structural vibration characteristics is essential for evaluating the safety and stability of a bridge. In this paper, a signal time-frequency feature extraction method (NTFT-ESVD) integrating standard time-frequency transformation, singular [...] Read more.
Bridges may undergo structural vibration responses when exposed to seismic waves. An analysis of structural vibration characteristics is essential for evaluating the safety and stability of a bridge. In this paper, a signal time-frequency feature extraction method (NTFT-ESVD) integrating standard time-frequency transformation, singular value decomposition, and information entropy is proposed to analyze the vibration characteristics of structures under seismic excitation. First, the experiment simulates the response signal of the structure when exposed to seismic waves. The results of the time-frequency analysis indicate a maximum relative error of only 1% in frequency detection, and the maximum relative errors in amplitude and time parameters are 5.9% and 6%, respectively. These simulation results demonstrate the reliability of the NTFT-ESVD method in extracting the time-frequency characteristics of the signal and its suitability for analyzing the seismic response of the structure. Then, a real seismic wave event of the Su-Tong Yangtze River Bridge during the Hengchun earthquake in Taiwan (2006) is analyzed. The results show that the seismic waves only have a short-term impact on the bridge, with the maximum amplitude of the vibration response no greater than 1 cm, and the maximum vibration frequency no greater than 0.2 Hz in the three-dimensional direction, indicating that the earthquake in Hengchun will not have any serious impact on the stability and security of the Su-Tong Yangtze River Bridge. Additionally, the reliability of determining the arrival time of seismic waves by extracting the time-frequency information from structural vibration response signals is validated by comparing it with results from seismic stations (SSE/WHN/QZN) at similar epicenter distances published by the USGS. The results of the case study show that the combination of dynamic GNSS monitoring technology and time-frequency analysis can be used to analyze the impact of seismic waves on the bridge, which is of great help to the manager in assessing structural seismic damage. Full article
(This article belongs to the Section Multidisciplinary Applications)
Show Figures

Figure 1

Figure 1
<p>The singular value distribution curve of the signal. (<b>a</b>,<b>c</b>,<b>e</b>): The pure noise signal, pure signal and noisy signal; (<b>b</b>,<b>d</b>,<b>f</b>): Singular value distribution curves of the signals corresponding to sub-figures (<b>a</b>,<b>c</b>,<b>e</b>) respectively.</p>
Full article ">Figure 2
<p>Singular value difference spectrum of the noisy signal.</p>
Full article ">Figure 3
<p>The analysis flowchart of the vibration response signal.</p>
Full article ">Figure 4
<p>The simulated signal with a signal-to-noise ratio of −2.</p>
Full article ">Figure 5
<p>The FFT spectrum of the simulated signal.</p>
Full article ">Figure 6
<p>The components’ extraction results of simulated signal by the ‘Inaction method’ (<b>a</b>) Extraction results of Signal 1; (<b>b</b>) Extraction results of Signal 2.</p>
Full article ">Figure 7
<p>The signal extracted by NTFT was processed using the SVD method. (<b>a</b>,<b>b</b>): Singular value distribution curves of the Signal 1 and Signal 2; (<b>c</b>,<b>d</b>): The ‘ambiguity points’ distribution of singular value difference spectrum; (<b>e</b>,<b>f</b>): Extraction results of Signal 1 and Signal 2 using NTFT-ESVD method.</p>
Full article ">Figure 8
<p>The results of the signal extraction using the NTFT method and the NTFT-ESVD method. (<b>a</b>) Extraction results of the simulation signal; (<b>b</b>) Deviation statistics of signal extraction results.</p>
Full article ">Figure 9
<p>The time-frequency spectrum of the simulated signal. (<b>a</b>,<b>c</b>): Signal 1 extracted by the NTFT method and the NTFT-ESVD method; (<b>b</b>,<b>d</b>): Signal 2 extracted by the NTFT method and the NTFT-ESVD method.</p>
Full article ">Figure 10
<p>Overview map showing the location of the Hengchun earthquake, the Su-Tong Yangtze River Bridge, and the seismic station.</p>
Full article ">Figure 11
<p>The GPS receivers were arranged at the top of the bridge tower.</p>
Full article ">Figure 12
<p>The vibration response signal in three dimensions. (<b>a</b>): The vibration response signal in the X direction; The red rectangular box shows the obviously abnormal vibration; (<b>b</b>): The vibration response signal in the Y direction; (<b>c</b>): The vibration response signal in the Z direction.</p>
Full article ">Figure 13
<p>The wind speed measured from 12:00 UTC to 13:00 UTC by the anemometer.</p>
Full article ">Figure 14
<p>The FFT spectrum of the vibration response signal in three-dimensional direction. (<b>a</b>): The FFT spectrum in the X direction; (<b>b</b>): The FFT spectrum in the Y direction; (<b>c</b>): The FFT spectrum in the Z direction.</p>
Full article ">Figure 15
<p>The extracted signal in three-dimensional direction. (<b>a</b>): Signal extraction results in the X direction; (<b>b</b>): Signal extraction results in the Y direction; (<b>c</b>): Signal extraction results in the Z direction.</p>
Full article ">Figure 16
<p>The NTFT spectrum of the extracted signal in three-dimensional direction. (<b>a</b>,<b>c</b>,<b>e</b>): Time-frequency spectrum of the original vibration response signals in the X, Y, and Z directions; (<b>b</b>,<b>d</b>,<b>f</b>): Time-frequency spectrum of the vibration response signals extracted in the X, Y, and Z directions.</p>
Full article ">
15 pages, 1684 KiB  
Article
Association between Urine Specific Gravity as a Measure of Hydration Status and Risk of Type 2 Diabetes: The Kailuan Prospective Cohort Study
by Yinqiao Dong, Shuohua Chen, Yaohui Yu, Wenjuan Li, Zhongqing Xu, Juan Du, Shan Huang, Shouling Wu and Yong Cai
Nutrients 2024, 16(11), 1643; https://doi.org/10.3390/nu16111643 - 27 May 2024
Viewed by 948
Abstract
Diabetes, especially type 2 diabetes (T2D), poses an unprecedented challenge to global public health. Hydration status also plays a fundamental role in human health, especially in people with T2D, which is often overlooked. This study aimed to explore the longitudinal associations between hydration [...] Read more.
Diabetes, especially type 2 diabetes (T2D), poses an unprecedented challenge to global public health. Hydration status also plays a fundamental role in human health, especially in people with T2D, which is often overlooked. This study aimed to explore the longitudinal associations between hydration status and the risk of T2D among the Chinese population. This study used data from the large community-based Kailuan cohort, which included adults who attended physical examinations from 2006 to 2007 and were followed until 2020. A total of 71,526 participants who eventually met the standards were divided into five hydration-status groups based on their levels of urine specific gravity (USG). Multivariable and time-dependent Cox proportional hazards models were employed to evaluate the associations of baseline and time-dependent hydration status with T2D incidence. Restricted cubic splines (RCS) analysis was used to examine the dose–response relationship between hydration status and the risk of T2D. Over a median 12.22-year follow-up time, 11,804 of the participants developed T2D. Compared with the optimal hydration-status group, participants with dehydration and severe dehydration had a significantly increased risk of diabetes, with adjusted hazard ratios (95% CI) of 1.30 (1.04–1.63) and 1.38 (1.10–1.74). Time-dependent analyses further confirmed the adverse effects of impending dehydration, dehydration, and severe dehydration on T2D incidence by 16%, 26%, and 33% compared with the reference group. Inadequate hydration is significantly associated with increased risks of T2D among Chinese adults. Our findings provided new epidemiological evidence and highlighted the potential role of adequate hydration status in the early prevention of T2D development. Full article
(This article belongs to the Special Issue Nutritional Epidemiology of Diabetes)
Show Figures

Figure 1

Figure 1
<p>Kaplan–Meier curves of cumulative incidence of T2D among the overall population. Note: hydration status categories were based on urine specific gravity (USG) as Group 1 (G1): 1.000 ≤ USG ≤ 1.010 g/mL; Group 2 (G2): 1.010 &lt; USG &lt; 1.015 g/mL; Group 3 (G3): 1.015 ≤ USG &lt; 1.020 g/mL; Group 4 (G4): 1.020 ≤ USG &lt; 1.030 g/mL; Group 5 (G5): USG ≥ 1.030 g/mL. G1 was used as the reference group. <span class="html-italic">p</span> &lt; 0.001 for differences among curves using the log-rank test.</p>
Full article ">Figure 2
<p>Restricted cubic spline analyses of the associations of continuous urinary specific gravity (USG) with risk of type 2 diabetes in (<b>A</b>) the overall population (<b>B</b>) female (<b>C</b>) male. Note: point estimates (solid line) and 95% confidence intervals (dashed lines) were based on Cox regression models of the restricted cubic spline with 3 knots at 10th, 50th, and 90th percentiles. All models were adjusted for sex, age, education, smoking status, alcohol drinking, physical activity, salt intake, history of hypertension, body mass index, high-sensitivity C-reactive protein, total cholesterol, triglyceride, eGFR, hemoglobin, serum uric acid, blood urea nitrogen, and plasma creatinine.</p>
Full article ">Figure 3
<p>Subgroup analysis of the association between hydration status and risk of T2D based on the multivariable Cox regression models. Note: incidence rate, per 1000 person years. Model 1: crude model; Model 2: adjusted for age, gender, BMI (categorical), education, smoking, drinking, physical activity, and intake of salt based on model 1; Model 3: further adjusted for history of hypertension, total cholesterol (TC), triglyceride (TG), C-reactive protein (CRP), serum uric acid (SUA), estimated glomerular filtration rate (eGFR), blood urea nitrogen (BUN), plasma creatinine (Cre), and hematocrit based on model 2.</p>
Full article ">
Back to TopTop