Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (45)

Search Parameters:
Keywords = radar bathymetry

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 22139 KiB  
Article
Improving the Estimation of Lake Ice Thickness with High-Resolution Radar Altimetry Data
by Anna Mangilli, Claude R. Duguay, Justin Murfitt, Thomas Moreau, Samira Amraoui, Jaya Sree Mugunthan, Pierre Thibaut and Craig Donlon
Remote Sens. 2024, 16(14), 2510; https://doi.org/10.3390/rs16142510 - 9 Jul 2024
Viewed by 978
Abstract
Lake ice thickness (LIT) is a sensitive indicator of climate change, identified as a thematic variable of Lakes as an Essential Climate Variable (ECV) by the Global Climate Observing System (GCOS). Here, we present a novel and efficient analytically based retracking approach for [...] Read more.
Lake ice thickness (LIT) is a sensitive indicator of climate change, identified as a thematic variable of Lakes as an Essential Climate Variable (ECV) by the Global Climate Observing System (GCOS). Here, we present a novel and efficient analytically based retracking approach for estimating LIT from high-resolution Ku-band (13.6 GHz) synthetic-aperture radar (SAR) altimetry data. The retracker method is based on the analytical modeling of the SAR radar echoes over ice-covered lakes that show a characteristic double-peak feature attributed to the reflection of the Ku-band radar waves at the snow–ice and ice–water interfaces. The method is applied to Sentinel-6 Unfocused SAR (UFSAR) and Fully Focused SAR (FFSAR) data, with their corresponding tailored waveform model, referred to as the SAR_LIT and FFSAR_LIT retracker, respectively. We found that LIT retrievals from Sentinel-6 high-resolution SAR data at different posting rates are fully consistent with the LIT estimations obtained from thermodynamic lake ice model simulations and from low-resolution mode (LRM) Sentinel-6 and Jason-3 data over two ice seasons during the tandem phase of the two satellites, demonstrating the continuity between LRM and SAR LIT retrievals. By comparing the Sentinel-6 SAR LIT estimates to optical/radar images, we found that the Sentinel-6 LIT measurements are fully consistent with the evolution of the lake surface conditions, accurately capturing the seasonal transitions of ice formation and melt. The uncertainty in the LIT estimates obtained with Sentinel-6 UFSAR data at 20 Hz is in the order of 5 cm, meeting the GCOS requirements for LIT measurements. This uncertainty is significantly smaller, by a factor of 2 to 3 times, than the uncertainty obtained with LRM data. The FFSAR processing at 140 Hz provides even better LIT estimates, with 20% smaller uncertainties. The LIT retracker analysis performed on data at the higher posting rate (140 Hz) shows increased performance in comparison to the 20 Hz data, especially during the melt transition period, due to the increased statistics. The LIT analysis has been performed over two representative lakes, Great Slave Lake and Baker Lake (Canada), demonstrating that the results are robust and hold for lake targets that differ in terms of size, bathymetry, snow/ice properties, and seasonal evolution of LIT. The SAR LIT retrackers presented are promising tools for monitoring the inter-annual variability and trends in LIT from current and future SAR altimetry missions. Full article
(This article belongs to the Special Issue Remote Sensing of the Cryosphere (Second Edition))
Show Figures

Figure 1

Figure 1
<p>Illustration of the evolution of the bimodal lake ice thickness signature in Sentinel-6 UFSAR radargrams (<b>left column</b>) and the normalized waveforms (<b>right column</b>) at 20 Hz resolution at Great Slave Lake in December 2021 (<b>top</b>), February 2021 (<b>second row</b>), end of April 2021 (<b>third row</b>), and May 2021 (<b>bottom</b>). The black line in the plots of the right column corresponds to the mean waveform in the selected region of the lake.</p>
Full article ">Figure 2
<p>Illustration of the evolution of the bimodal lake ice thickness signature in Sentinel-6 FFSAR radargrams (<b>left column</b>) and the normalized waveforms (<b>right column</b>) at 140 Hz posting rate at Great Slave Lake in December 2021 (<b>top</b>), February 2021 (<b>second row</b>), end of April 2021 (<b>third row</b>), and May 2021 (<b>bottom</b>). The black line in the plots of the right column corresponds to the mean waveform in the selected region of the lake.</p>
Full article ">Figure 3
<p>The target lakes of the LIT analysis, Great Slave Lake and Baker Lake, Canada, are shown on the map (<b>bottom</b>) and with the satellite ground tracks superimposed on the lakes (<b>upper left</b> and <b>right</b>, respectively).</p>
Full article ">Figure 4
<p>Examples of Sentinel-6 UFSAR waveform with, in blue, the <tt>SAR_LIT</tt> fit (<b>left column</b>) and LIT histograms, with the corresponding Gaussian fits (<b>right column</b>), in the RoI of Great Slave Lake at the end of December 2020 (<b>top row</b>), in February 2021 (<b>second row</b>), in April 2021 (<b>third row</b>), and mid-May 2021 (<b>bottom row</b>).</p>
Full article ">Figure 5
<p>Examples of Sentinel-6 FFSAR waveforms with, in blue, the <tt>FFSAR_LIT</tt> fit (<b>left column</b>) and LIT histograms, with the corresponding Gaussian fits (<b>right column</b>), in the RoI of Great Slave Lake at the end of December 2020 (<b>top row</b>), in February 2021 (<b>second row</b>), in April 2021 (<b>third row</b>), and mid-May 2021 (<b>bottom row</b>).</p>
Full article ">Figure 6
<p>Example of the spatial evolution of the LIT estimates at Great Slave Lake (<b>left column</b>) and Baker Lake (<b>right column</b>) in February 2021. The top row plots show the results for the Sentinel-6 UFSAR data at 20 Hz, while the bottom row plots for the Sentinel-6 FFSAR data at 140 Hz. The gray lines in the bottom panels of the figures show the evolution of the reduced <math display="inline"><semantics> <msup> <mi>χ</mi> <mn>2</mn> </msup> </semantics></math> goodness of fit metric.</p>
Full article ">Figure 7
<p>Comparison of the LIT estimates obtained with Sentinel-6 high-resolution SAR data for one ice season at Great Slave Lake (<b>left</b>) and Baker Lake (<b>right</b>). The curves refer to UFSAR at 20 Hz (red), UFSAR at 140 Hz (purple), and FFSAR at 140 Hz (cyan).</p>
Full article ">Figure 8
<p>Evolution of LIT estimates at Great Slave Lake obtained with Sentinel-6 UFSAR at 20 Hz data (red), Sentinel-6 LRM data (green) and Jason-3 data (blue) for the 2020–2021 and 2021–2022 ice seasons (upper panel). The shaded regions of the corresponding colors refer to the LIT error envelopes at 1<math display="inline"><semantics> <mi>σ</mi> </semantics></math> for each case. The orange shaded area shows the evolution of LIT obtained from CLIMo thermodynamic simulations with different on-ice snow scenarios (see text in <a href="#sec4dot3-remotesensing-16-02510" class="html-sec">Section 4.3</a> for details). The middle panel shows the evolution of the mean 2 m air temperature (black) with the minimum and maximum values (gray shading) extracted from ERA5 data. The bottom panel shows the evolution of the 1<math display="inline"><semantics> <mi>σ</mi> </semantics></math> LIT uncertainties for the three datasets.</p>
Full article ">Figure 9
<p>Evolution of the LIT estimates at Baker Lake obtained with different datasets (see the caption of <a href="#remotesensing-16-02510-f008" class="html-fig">Figure 8</a> for details).</p>
Full article ">Figure 10
<p>Sentinel–6 20 Hz UFSAR LIT estimates superimposed on radar/optical images taken on the same dates for Great Slave Lake (<b>left</b>) and Baker Lake (<b>right</b>) on the lake area shown in the red boxes in the top row panels.</p>
Full article ">
17 pages, 8577 KiB  
Article
A Feasibility Study of Nearshore Bathymetry Estimation via Short-Range K-Band MIMO Radar
by Giovanni Ludeno, Matteo Antuono, Francesco Soldovieri and Gianluca Gennarelli
Remote Sens. 2024, 16(2), 261; https://doi.org/10.3390/rs16020261 - 9 Jan 2024
Cited by 1 | Viewed by 2336
Abstract
This paper provides an assessment of a 24 GHz multiple-input multiple-output radar as a remote sensing tool to retrieve bathymetric maps in coastal areas. The reconstruction procedure considered here exploits the dispersion relation and has been previously employed to elaborate the data acquired [...] Read more.
This paper provides an assessment of a 24 GHz multiple-input multiple-output radar as a remote sensing tool to retrieve bathymetric maps in coastal areas. The reconstruction procedure considered here exploits the dispersion relation and has been previously employed to elaborate the data acquired via X-band marine radar. The estimation capabilities of the sensor are investigated firstly on synthetic radar data. With this aim, case studies referring to sea waves interacting with a constant and a spatially varying bathymetry are both considered. Finally, the reconstruction procedure is tested by processing real data recorded at Bagnoli Bay, Naples, South Italy. The preliminary results shown here confirm the potential of the radar sensor as a tool for sea wave monitoring. Full article
Show Figures

Figure 1

Figure 1
<p>RF front end of the radar system.</p>
Full article ">Figure 2
<p>Representation of the shadowing and tilt modulation effect in the vertical plane. The Tx/Rx antenna pair is located at quota H above the MSL. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> delimit the radar FoV.</p>
Full article ">Figure 3
<p>Data processing chain for bathymetry map estimation.</p>
Full article ">Figure 4
<p>Synthetic SWF<sub>1</sub> at time <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>64</mn> <mo> </mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>. (<b>a</b>) Elevation profile. (<b>b</b>) Radar image.</p>
Full article ">Figure 5
<p>(<b>a</b>) SNR map for SWF<sub>1</sub> with, superimposed upon it, the black solid line enclosing the area where SNR ≥ 3dB. (<b>b</b>–<b>d</b>) Reconstructed bathymetry field for SWF<sub>1</sub>, SWF<sub>2</sub>, and SWF<sub>3</sub>, respectively.</p>
Full article ">Figure 6
<p>Synthetic SWF<sub>PSB1</sub> at time <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>64</mn> <mo> </mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>. (<b>a</b>) Sea waves propagating over the linear bathymetry. (<b>b</b>) Radar image.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>c</b>) Bathymetry maps based on the SNR ≥ 3 dB criterion retrieved from SWF<sub>PSB1</sub>, SWF<sub>PSB2</sub>, and SWF<sub>PSB3</sub>, respectively.</p>
Full article ">Figure 8
<p>Comparison between estimated bathymetry (dots) and ground truth (red dashed line). The equation of the regression line (black dashed line) is provided in the legend. (<b>a</b>–<b>c</b>) Scatterplot for SWF<sub>PSB1</sub>, SWF<sub>PSB2</sub>, and SWF<sub>PSB3</sub>, respectively.</p>
Full article ">Figure 9
<p>Satellite image of the test site. The inset displays the radar’s positioning overlaid on a radar image. A1 and A2 show the location of two anemometric stations.</p>
Full article ">Figure 10
<p>(<b>a</b>) SRK-band radar system installation. (<b>b</b>) Cross-section of normalized 3D radar data spectrum (color scale [0,1]).</p>
Full article ">Figure 11
<p>(<b>a</b>) Bathymetry field estimated at Bagnoli Bay. The black solid line encloses the area where the SNR ≥ 3 dB. (<b>b</b>) Georeferenced SNR bathymetry field based on SNR ≥ 3dB criterion. The white line is the isobath at depth 2 m.</p>
Full article ">
32 pages, 12596 KiB  
Article
Multi-Timescale Characteristics of Southwestern Australia Nearshore Surface Current and Its Response to ENSO Revealed by High-Frequency Radar
by Hongfei Gu and Yadan Mao
Remote Sens. 2024, 16(1), 209; https://doi.org/10.3390/rs16010209 - 4 Jan 2024
Viewed by 1596
Abstract
The surface currents in coastal areas are closely related to the ecological environment and human activities, and are influenced by both local and remote factors of different timescales, resulting in complex genesis and multi-timescale characteristics. In this research, 9-year-long, hourly high-frequency radar (HFR) [...] Read more.
The surface currents in coastal areas are closely related to the ecological environment and human activities, and are influenced by both local and remote factors of different timescales, resulting in complex genesis and multi-timescale characteristics. In this research, 9-year-long, hourly high-frequency radar (HFR) surface current observations are utilized together with satellite remote sensing reanalysis products and mooring data, and based on the Empirical Orthogonal Function (EOF) and correlation analysis, we revealed the multi-timescale characteristics of the surface currents in Fremantle Sea (32°S), Southwestern Australia, and explored the corresponding driving factors as well as the impact of El Niño-Southern Oscillation (ENSO) on the nearshore currents. Results show that the currents on the slope are dominated by the southward Leeuwin Current (LC), and the currents within the shelf are dominated by winds, which are subject to obvious diurnal and seasonal variations. The strong bathymetry variation there, from a wide shelf in the north to a narrow shelf in this study region, also plays an important role, resulting in the frequent occurrence of nearshore eddies. In addition, the near-zonal winds south of 30°S in winter contribute to the interannual variability of the Leeuwin Current at Fremantle, especially in 2011, when the onshore shelf circulation is particularly strong because of the climatic factors, together with the wind-driven offshore circulation, which results in significant and long-lasting eddies. The southward Leeuwin Current along Southwestern Australia shows a strong response to interannual climatic variability. During La Niña years, the equatorial thermal anomalies generate the westward anomalies in winds and equatorial currents, which in turn strengthen the Leeuwin Current and trigger the cross-shelf current as well as downwelling within the shelf at Fremantle, whereas during El Niño years, the climate anomalies and the response of coastal currents are opposite. This paper provides insights into the multi-timescale nature of coastal surface currents and the relative importance of different driving mechanisms. It also demonstrates the potential of HFR to reveal the response of nearshore currents to climate anomalies when combined with other multivariate data. Meanwhile, the methodology adopted in this research is applicable to other coastal regions with long-term available HFR observations. Full article
Show Figures

Figure 1

Figure 1
<p>Western Australia, the ROT radar site located in Fremantle Sea, and the effective sampling proportion of radar. (<b>a</b>) The research area, North West Cape is the north-south boundary of Western Australia; the red (blue) arrows represent Leeuwin Current (Capes Current) [<a href="#B31-remotesensing-16-00209" class="html-bibr">31</a>]; the magenta (blue) dashed box indicates the range of driving factors data used to investigate the causes of diurnal (seasonal and interannual) characteristics of the surface currents; (<b>b</b>) The observing range of ROT radar and the spatial distribution of temporal effective sampling proportion of it; WATR10, WATR20, and WACA20 are 3 mooring sites; the white (black) dashed box indicates Perth Canyon (Rottnest Island); (<b>c</b>) The percentage ratio of radar observing hours per month.</p>
Full article ">Figure 2
<p>Frequency spectra of the hourly HFR-derived current component time series. (<b>A1</b>–<b>F1</b>) shows the original hourly time series; (<b>A2</b>–<b>F2</b>) shows the corresponding frequency spectra of (<b>A1</b>–<b>F1</b>), in which the black vertical dashed lines indicate the Freq value of −3.9425 (i.e., 365 days), and the red dashed boxes indicate significant spectra peaks and intervals; the ‘Freq’ and ‘Amp’ values represent the frequency and amplitude, respectively; the ‘h’ of the Freq value represents one hour; the minimum value of the horizontal axis is −4.2553 (i.e., 750 days).</p>
Full article ">Figure 3
<p>Diurnal, seasonal, and interannual statistical values of the ROT radar-observed surface currents. (<b>A1</b>–<b>D2</b>) show the mean values of currents during the 4 diurnal and seasonal periods, respectively; the arrows represent the current direction and the color bars represent the absolute value of current speed; (<b>A3</b>–<b>D3</b>) show the RMS values of current interannual STD; the scales of horizontal and vertical lines of the cross (‘+’) represent the variability of current U and V components, respectively; ‘GUI’ and ‘FRE’ represent Guildton and Fremantle, respectively; ‘MAM’, ‘JJA’, SON’, and ‘DJF’ represent the austral autumn (March to May), winter (June to August), spring (September to November), and summer (December to February), respectively.</p>
Full article ">Figure 4
<p>The mean flow vectors in three spatial units at three timescales. Each subplot is labeled with spatial units, as well as the maximum (red ‘Max’), minimum (blue ‘Min’), and average (black ‘Mean’) speed value of the current; Each column of subplots belongs to the same timescale.</p>
Full article ">Figure 5
<p>Histogram statistics of the surface current direction at three timescales in three spatial units. Each subplot is labeled with timescales, as well as the maximum value (red ‘Max’), mean value (blue ‘Mean’), and median value (black ‘Median’) of current speed; Each column of subplots belongs to the same spatial unit.</p>
Full article ">Figure 6
<p>Primary EOF modes of the HFR-derived current at three timescales. Each column of subplots belongs to the same timescale; (<b>A1</b>–<b>C3</b>) display the spatial modes (EOF) in the upper panels and the corresponding temporal coefficient (PC) in the lower panels, the titles for each subplot indicate the timescale and the modal sequence number; The contribution to the total variance of each mode is marked in red ‘Contr’ and the following number; The color on each grid indicates the relative (normalized) speed of the current EOF mode. The red lines in (<b>B1</b>,<b>C1</b>) represent the seasonal and interannual signals of Fremantle Sea Level (in units of mm); The data are not converted into anomalies before EOF analysis (thus the contribution coefficients of the modes behind EOF1 are relatively low) to primarily reflect the mean state but not variation of current.</p>
Full article ">Figure 7
<p>The spatially averaged U-component time series of the HFR-derived current and winds in 3 spatial units at three timescales and their correlation; Each column of subplots belongs to the same timescale; The correlation coefficient between each pair of timeseries is marked as ‘Corr’; (<b>A1</b>–<b>C3</b>) display the comparison results of the wind and current time series; (<b>D1</b>–<b>D3</b>) display the scatter plot analysis results for diurnal, seasonal and interannual scales, respectively.</p>
Full article ">Figure 8
<p>The spatially averaged V-component time series of the HFR-derived current and wind in 3 spatial units at three timescales and their correlation; Each column of subplots belongs to the same timescale; The correlation coefficient between each pair of timeseries is marked as ‘Corr’; (<b>A1</b>–<b>C3</b>) display the comparison results of the wind and current time series; (<b>D1</b>–<b>D3</b>) display the scatter plot analysis results for diurnal, seasonal and interannual scales, respectively.</p>
Full article ">Figure 9
<p>Typical wind (<b>A1</b>–<b>C2</b>), satellite-observed current (<b>D</b>), and SSH (<b>E</b>) EOF modes correlated with the primary HFR current EOF modes at three timescales; The legend of each PC subplot indicates the mode pair; The ‘Corr’ marks the corresponding correlation coefficient; The blue dot indicates Fremantle; For each subplot, the upper panel displays the spatial mode (EOF) and the lower panel displays the comparison of PCs indicated by the legend; In (<b>D</b>), the ‘SC’ indicates the satellite-observed current with a broader spatial range than HFR current.</p>
Full article ">Figure 10
<p>Diurnal variation of the spatially averaged HFR-derived current in three spatial units and four seasons. For each subplot, the spatial unit, as well as the maximum (red ‘Max’), minimum (blue ‘Min’), and average (black ‘Mean’) current speed are labeled; Each column of subplots belongs to the same season.</p>
Full article ">Figure 11
<p>Diurnal variation of winds in 4 seasons (from March 2010 to February 2019) in Southwestern Australia. For each subplot, the season, diurnal duration are labeled in the title; Each row of subplots belongs to the same season (e.g., (<b>A1</b>–<b>D1</b>) show the results for DJF i.e. austral summer. (<b>A2</b>–<b>D2</b>) show the results for MAM i.e. austral autumn; (<b>A3</b>–<b>D3</b>) show the results for JJA i.e. austral winter; (<b>A4</b>–<b>D4</b>) show the results for SON i.e. austral spring).</p>
Full article ">Figure 12
<p>Correlation of SOI with FSL, satellite-observed current, and SSH in Southwestern Australia. (<b>a</b>–<b>c</b>) shows the spatial distribution of the correlation coefficients; (<b>d</b>) shows the time series of SOI and FSL; (<b>e</b>,<b>f</b>) shows the cross-shore and alongshore currents in two spatial grids; The black pentagrams and rhombus in (<b>a</b>–<b>c</b>) indicate two sampling points; the black dashed boxes in each subplot indicate important features; and the white dashed box in (<b>c</b>) indicates Perth Canyon.</p>
Full article ">Figure 13
<p>Composite satellite-observed surface current anomalies before and after the peak months of El Niño and La Niña events in Southwestern Australia. The blue dot indicates Fremantle; The red (blue) contour indicates the 50 m (2000 m) isobath; The northern and southern magenta dashed boxes in (<b>d</b>) indicate the slope narrowing and Pearh Canyon, respectively; The white dashed boxes in (<b>a</b>,<b>e</b>,<b>g</b>,<b>k</b>) are used to compare the current anomalies in the same region at different durations; (<b>a</b>–<b>l</b>) display the current anomalies during the composite El Niño and La Niña year, respectively.</p>
Full article ">Figure 14
<p>The mean states (<b>a</b>–<b>d</b>) and anomalies (<b>e</b>–<b>h</b>) of winter and summer HFR-derived currents during 2 typical ENSO events. The blue dot indicates Fremantle, and the color bars for winter and summer currents have different speed value ranges (0–0.8 m/s and 0–0.4 m/s).</p>
Full article ">Figure 15
<p>Comparison of the histogram statistics of the current components of winter and summer HFR-derived current in the Fremantle inner and outer shelf during 2 typical ENSO events. The ‘n1’ (blue) represents the number of sampling hours in each period, and ‘n2’ (red) represents the number of intersecting sampling hours of the comparison periods (such as (<b>a</b>) to (<b>b</b>)), Only the results belonging to the n2 h are calculated to avoid statistical bias; In (<b>a</b>,<b>b</b>), the red and blue dashed boxes indicate the comparison of outer shelf current zonal components in two winters, which is the most significant finding in this figure; The black dashed boxes in (<b>c</b>–<b>h</b>) are also used to compare the current components in different durations.</p>
Full article ">Figure 16
<p>Differences in sea interior response during El Niño and La Niña as reflected by moorings and HFR together. Multilayer temperatures (<b>A1</b>,<b>A2</b>), top-bottom temperature difference (<b>B1</b>,<b>B2</b>), vertical velocity at the mooring site of WATR10 (<b>C1</b>,<b>C2</b>), the averaged outer shelf current U-component (<b>D1</b>,<b>D2</b>), and the averaged inner shelf current V-component (<b>E1</b>,<b>E2</b>) measured by HFR during the same periods; In (<b>A1</b>–<b>E1</b>), the red dashed box indicates a downwelling event, which were captured by both the mooring and HFR; In (<b>A2</b>), the blue box indicates a water temperature rising during the La Niña winter.</p>
Full article ">Figure 17
<p>Comparison of the top-bottom temperature difference (<b>A1</b>,<b>A2</b>) (WATR20) and the U (<b>B1</b>,<b>B2</b>), V (<b>C1</b>,<b>C2</b>), and W (<b>D1</b>,<b>D2</b>) components of the current (WACA20) observed by moorings in winter during 2 typical ENSO events. The meanings of ‘n1’ and ‘n2’ are the same as in <a href="#remotesensing-16-00209-f015" class="html-fig">Figure 15</a>; In (<b>A1</b>,<b>A2</b>) and (<b>D1</b>,<b>D2</b>), the red and blue boxes are used to compare the mooring-observed top-bottom temperature differences (current W-components) at two durations.</p>
Full article ">
21 pages, 14693 KiB  
Article
Automated High-Resolution Bathymetry from Sentinel-1 SAR Images in Deeper Nearshore Coastal Waters in Eastern Florida
by Sanduni D. Mudiyanselage, Ben Wilkinson and Amr Abd-Elrahman
Remote Sens. 2024, 16(1), 1; https://doi.org/10.3390/rs16010001 - 19 Dec 2023
Cited by 4 | Viewed by 2832
Abstract
Synthetic aperture radar (SAR) imagers are active microwave sensors that could overcome many challenges of passive optical bathymetry inversion, yet their capacity to yield accurate high-resolution bathymetric mapping is not studied sufficiently. In this study, we evaluate the feasibility of applying fast Fourier [...] Read more.
Synthetic aperture radar (SAR) imagers are active microwave sensors that could overcome many challenges of passive optical bathymetry inversion, yet their capacity to yield accurate high-resolution bathymetric mapping is not studied sufficiently. In this study, we evaluate the feasibility of applying fast Fourier transform (FFT) to SAR data in coastal nearshore bathymetry derivation in Florida’s coastal waters. The study aims to develop a robust SAR bathymetry inversion framework across extensive spatial scales to address the dearth of bathymetric data in deeper nearshore coastal regions. By leveraging the Sentinel-1 datasets as a rich source of training data, our method yields high-resolution and accurate depth extraction up to 80 m. A comprehensive workflow to determine both the wavelength and peak wave period is associated with the proposed automated model compilation. A novel contour geometry-based spectral analysis technique for wavelength retrieval is presented that enables an efficient and scalable SAR bathymetry model. Multi-date SAR images were used to assess the robustness of the proposed depth-retrieval model. An accuracy assessment against the GMRT data demonstrated the high efficacy of the proposed approach, achieving a coefficient of determination (R2) above 0.95, a root-mean-square error (RMSE) of 1.56–10.20 m, and relative errors of 3.56–11.08% in automatically extracting the underwater terrain at every 50 m interval. A sensitivity analysis was conducted to estimate the uncertainty associated with our method. Overall, this study highlights the potential of SAR technology to produce updated, cost-effective, and accurate bathymetry maps of high resolution and to fill bathymetric data gaps worldwide. The code and datasets are made publicly available. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study region in Eastern Florida coastal waters in the United States (together with the depth classes derived from the GMRT profile including the transect orientation from 1 to 40 (transect 1 is the southernmost and transect 40 is the northernmost).</p>
Full article ">Figure 2
<p>Orientation of Sentinel-1 tile A9E4 covering eastern Florida and the zoomed-in version of imaged wave patterns within the red-colored enclosed area.</p>
Full article ">Figure 3
<p>FFT intensity spectrum representations of six selected SAR subsets. The enclosed sections (red) are zoomed in and depicted in the bottom right of each image. These reveal the untidiness of the high-intensity clusters from which the highest-intensity location must be identified.</p>
Full article ">Figure 4
<p>Frequency domain representation (<b>a</b>), contour geometry highlighting the high-intensity blobs in yellow (<b>b</b>), and identified peak intensity locations using the centroid of these blobs marked with red markers (<b>c</b>) in an example FFT output.</p>
Full article ">Figure 5
<p>Workflow of the methodology adopted in the present study, including the steps involved in the determination of the dominant wavelength of the SAR subset and nautical chart-based peak wave periods.</p>
Full article ">Figure 6
<p>Contour plots of the FFT outputs (<b>left</b>) and centroids of the high-intensity blobs (<b>right</b>) of the FFT outputs shown in <a href="#remotesensing-16-00001-f003" class="html-fig">Figure 3</a>. The centroids are marked using a red cross in each case. These centroid locations are used as the representative locations of the highest intensities on FFT outputs.</p>
Full article ">Figure 7
<p>Calculated wavelength variation along selected four transects—12, 21, 29, and 30—for image <b>A</b> (<b>top</b>) and image <b>B</b> (<b>bottom</b>).</p>
Full article ">Figure 8
<p>Comparison between the predicted and observed bathymetry for transects 12, 21, 29 and 30 using image A.</p>
Full article ">Figure 9
<p>Comparison between the predicted and observed bathymetry for transects 12, 21, 29 and 30 using image B.</p>
Full article ">Figure 10
<p>SAR bathymetry (<b>a</b>), ground-truth (<b>b</b>), and prediction error (<b>c</b>) maps for images A and B along all 40 transects using the proposed method.</p>
Full article ">Figure 11
<p>Predicted–measured plots for 31,200 data points along all 40 transects put together for images A and B highlighting the color-coded depth classes. The black line represents the predicted = measured graph. The R<sup>2</sup> values considering all 31,200 data points covering the entire 0–80 m depth range are 0.96 (image <b>A</b>) and 0.95 (image <b>B</b>).</p>
Full article ">Figure 12
<p>Sensitivity analysis for <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="bold-italic">T</mi> <mi mathvariant="bold-italic">p</mi> </msub> <mo> </mo> </mrow> </semantics></math> at each water depth. The values listed and the colors refer to the depth change for each incremental <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="bold-italic">T</mi> <mi mathvariant="bold-italic">p</mi> </msub> </mrow> </semantics></math> (vertical axis) at different depths (horizontal axis).</p>
Full article ">Figure 13
<p>Sensitivity analysis for λ at each water depth. The values listed and the colors refer to the depth change for each incremental λ (vertical axis) at different depths (horizontal axis).</p>
Full article ">
34 pages, 51108 KiB  
Article
Seafloor and Ocean Crust Structure of the Kerguelen Plateau from Marine Geophysical and Satellite Altimetry Datasets
by Polina Lemenkova
Geomatics 2023, 3(3), 393-426; https://doi.org/10.3390/geomatics3030022 - 10 Aug 2023
Cited by 1 | Viewed by 2247
Abstract
The volcanic Kerguelen Islands are formed on one of the world’s largest submarine plateaus. Located in the remote segment of the southern Indian Ocean close to Antarctica, the Kerguelen Plateau is notable for a complex tectonic origin and geologic formation related to the [...] Read more.
The volcanic Kerguelen Islands are formed on one of the world’s largest submarine plateaus. Located in the remote segment of the southern Indian Ocean close to Antarctica, the Kerguelen Plateau is notable for a complex tectonic origin and geologic formation related to the Cretaceous history of the continents. This is reflected in the varying age of the oceanic crust adjacent to the plateau and the highly heterogeneous bathymetry of the Kerguelen Plateau, with seafloor structure differing for the southern and northern segments. Remote sensing data derived from marine gravity and satellite radar altimetry surveys serve as an important source of information for mapping complex seafloor features. This study incorporates geospatial information from NOAA, EMAG2, WDMAM, ETOPO1, and EGM96 datasets to refine the extent and distribution of the extracted seafloor features. The cartographic joint analysis of topography, magnetic anomalies, tectonic and gravity grids is based on the integrated mapping performed using the Generic Mapping Tools (GMT) programming suite. Mapping of the submerged features (Broken Ridge, Crozet Islands, seafloor fabric, orientation, and frequency of magnetic anomalies) enables analysis of their correspondence with free-air gravity and magnetic anomalies, geodynamic setting, and seabed structure in the southwest Indian Ocean. The results show that integrating the datasets using advanced cartographic scripting language improves identification and visualization of the seabed objects. The results include 11 new maps of the region covering the Kerguelen Plateau and southwest Indian Ocean. This study contributes to increasing the knowledge of the seafloor structure in the French Southern and Antarctic Lands. Full article
(This article belongs to the Special Issue Advances in Ocean Mapping and Nautical Cartography)
Show Figures

Figure 1

Figure 1
<p>Map of the Kerguelen Plateau region. Mapping: GMT. Map source: author.</p>
Full article ">Figure 2
<p>Age of the ocean crust in SW Indian Ocean and the Kerguelen Plateau region and east Antarctic. Mapping software: GMT v. 6-1-1. Map source: author.</p>
Full article ">Figure 3
<p>Asymmetry in crustal accretion on conjugate ridge flanks of the ocean crust over the Kerguelen Plateau region, east Antarctic and south-west Indian Ocean. Map source: author.</p>
Full article ">Figure 4
<p>Spreading half rates of the ocean crust over the Kerguelen Plateau region, east Antarctic and south-west Indian Ocean with added GSFML data. Mapping tool: GMT. Map source: author.</p>
Full article ">Figure 5
<p>Age uncertainty in lithosphere crust (m.y) over the Kerguelen Plateau region, east Antarctic and south-west Indian Ocean. Cartography: GMT. Map source: author.</p>
Full article ">Figure 6
<p>Sediment thickness (m) in East Antarctica, South-West Indian Ocean and the Kerguelen Plateau region. Cartography: GMT. Lambert azimuthal equal-area projection. Map source: author.</p>
Full article ">Figure 7
<p>Marine gravity field over the Kerguelen Plateau region, south-west Indian Ocean. Mapping: GMT. Map source: author.</p>
Full article ">Figure 8
<p>Vertical gravity gradient over Kerguelen Plateau. Mapping: GMT. Map source: author.</p>
Full article ">Figure 9
<p>Patterns of the marine and terrestrial airborne magnetic anomaly over the Kerguelen Plateau on a three-minute resolution grid of WDMAM, south-west Indian Ocean. Map source: author.</p>
Full article ">Figure 10
<p>Marine and Earth airborne magnetic anomaly grid based on EMAG-2 over the Kerguelen Plateau region. Black areas signify “no data” in the original grid. Map source: author.</p>
Full article ">Figure 11
<p>Geoid model based on EGM-96 over the Kerguelen Plateau region, east Antarctic and south-west Indian Ocean. Mapping: GMT. Map source: author.</p>
Full article ">
13 pages, 6090 KiB  
Article
Nearshore Observations and Modeling: Synergy for Coastal Flooding Prediction
by Matteo Postacchini, Lorenzo Melito and Giovanni Ludeno
J. Mar. Sci. Eng. 2023, 11(8), 1504; https://doi.org/10.3390/jmse11081504 - 28 Jul 2023
Cited by 6 | Viewed by 1326
Abstract
Coastal inundation has recently started to require significant attention worldwide. The increasing frequency and intensity of extreme events (sea storms, tsunami waves) are highly stressing coastal environments by endangering a large number of residential areas, ecosystems, and tourist facilities, and also leading to [...] Read more.
Coastal inundation has recently started to require significant attention worldwide. The increasing frequency and intensity of extreme events (sea storms, tsunami waves) are highly stressing coastal environments by endangering a large number of residential areas, ecosystems, and tourist facilities, and also leading to potential environmental risks. Predicting such events and the generated coastal flooding is thus of paramount importance and can be accomplished by exploiting the potential of different tools. An example is the combination of remote sensors, like marine radars, with numerical models. Specifically, while instruments like X-band radars are able to precisely reconstruct both wave field and bathymetry up to some kilometers off the coast, wave-resolving Boussinesq-type models can reproduce the wave propagation in the nearshore area and the consequent coastal flooding. Hence, starting from baseline simulations of wave propagation and the conversion of water elevation results into radar images, the present work illustrates the reconstruction of coastal data (wave field and seabed depth) using a specifically suited data processing method, named the “Local Method”, and the use of such coastal data to run numerical simulations of coastal inundation in different scenarios. Such scenarios were built using two different European beaches, i.e., Senigallia (Italy) and Oostende (Belgium), and three different directional spreading values to evaluate the performances in cases of either long- or short-crested waves. Both baseline and inundation simulations were run using the FUNWAVE-TVD solver. The overall validation of the methodology, in terms of maximum inundation, shows its good performance, especially in cases of short-crested wind waves. Furthermore, the application on Oostende Beach demonstrates that the present methodology might work using only open-access tools, providing an easy investigation of coastal inundation and potential low-cost integration into early warning systems. Full article
(This article belongs to the Topic Aquatic Environment Research for Sustainable Development)
Show Figures

Figure 1

Figure 1
<p>Sketch of the applied methodology.</p>
Full article ">Figure 2
<p>(<b>a</b>) Location of the two coastal sites (adapted from Google Earth). (<b>b</b>,<b>c</b>) Bathymetries for the two sites. Contours at 0.5 m intervals are traced with thin black lines. The 0 m contour is given with a thick black line. A different vertical scale is used.</p>
Full article ">Figure 3
<p>Numerical propagation of short waves over the Senigallia bathymetry (lower color map) in the baseline (large-scale) FUNWAVE simulations, for different wave spreading parameters: <span class="html-italic">σ</span> = 10 (middle color map) and <span class="html-italic">σ</span> = 30 (upper color map).</p>
Full article ">Figure 4
<p>Synthetic sea-wave image obtained using FUNWAVE-TVD (<b>left</b>) and simulated radar image (<b>right</b>) for the Oostende case with a directional spreading <math display="inline"><semantics><mrow><mi>σ</mi><mo>=</mo><mn>30</mn></mrow></semantics></math>.</p>
Full article ">Figure 5
<p>Comparison between (<b>left</b>) ground-truth and (<b>right</b>) reconstructed bathymetry for Oostende case, with spreading <math display="inline"><semantics><mrow><mi>σ</mi><mo>=</mo><mn>30</mn></mrow></semantics></math>.</p>
Full article ">Figure 6
<p>Comparison between (<b>left</b>) ground-truth and (<b>right</b>) reconstructed spatial map of significant wave height for Oostende case, with spreading <math display="inline"><semantics><mrow><mi>σ</mi><mo>=</mo><mn>30</mn></mrow></semantics></math>.</p>
Full article ">Figure 7
<p>Large-scale (thick lines) and small-scale (dotted lines) inundations modeled at (<b>a</b>) Senigallia and (<b>b</b>) Oostende for different wave spreading parameters (<math display="inline"><semantics><mrow><mi>σ</mi><mo>=</mo><mn>2</mn></mrow></semantics></math>: yellow; <math display="inline"><semantics><mrow><mi>σ</mi><mo>=</mo><mn>10</mn></mrow></semantics></math>: red; <math display="inline"><semantics><mrow><mi>σ</mi><mo>=</mo><mn>30</mn></mrow></semantics></math>: blue). Each colored line represents the maximum runup locations reached during each simulation.</p>
Full article ">Figure 8
<p>Comparison of cross-shore coordinates of small-scale vs. large-scale inundation for (<b>a</b>–<b>c</b>) Senigallia and (<b>d</b>–<b>f</b>) Oostende, as a function of the wave spreading parameter. Classical error statistics (correlation coefficient <span class="html-italic">p</span>, RMSE, and coefficient of determination <span class="html-italic">R</span><sup>2</sup>) are given for each case.</p>
Full article ">
19 pages, 3829 KiB  
Article
Computational Oil-Slick Hub for Offshore Petroleum Studies
by Nelson F. F. Ebecken, Fernando Pellon de Miranda, Luiz Landau, Carlos Beisl, Patrícia M. Silva, Gerson Cunha, Maria Célia Santos Lopes, Lucas Moreira Dias and Gustavo de Araújo Carvalho
J. Mar. Sci. Eng. 2023, 11(8), 1497; https://doi.org/10.3390/jmse11081497 - 27 Jul 2023
Cited by 1 | Viewed by 1286
Abstract
The paper introduces the Oil-Slick Hub (OSH), a computational platform to facilitate the data visualization of a large database of petroleum signatures observed on the surface of the ocean with synthetic aperture radar (SAR) measurements. This Internet platform offers an information search and [...] Read more.
The paper introduces the Oil-Slick Hub (OSH), a computational platform to facilitate the data visualization of a large database of petroleum signatures observed on the surface of the ocean with synthetic aperture radar (SAR) measurements. This Internet platform offers an information search and retrieval system of a database resulting from >20 years of scientific projects that interpreted ~15 thousand offshore mineral oil “slicks”: natural oil “seeps” versus operational oil “spills”. Such a Digital Mega-Collection Database consists of satellite images and oil-slick polygons identified in the Gulf of Mexico (GMex) and the Brazilian Continental Margin (BCM). A series of attributes describing the interpreted slicks are also included, along with technical reports and scientific papers. Two experiments illustrate the use of the OSH to facilitate the selection of data subsets from the mega collection (GMex variables and BCM samples), in which artificial intelligence techniques—machine learning (ML)—classify slicks into seeps or spills. The GMex variable dataset was analyzed with simple linear discriminant analyses (LDAs), and a three-fold accuracy performance pattern was observed: (i) the least accurate subset (~65%) solely used acquisition aspects (e.g., acquisition beam mode, date, and time, satellite name, etc.); (ii) the best results (>90%) were achieved with the inclusion of location attributes (i.e., latitude, longitude, and bathymetry); and (iii) moderate performances (~70%) were reached using only morphological information (e.g., area, perimeter, perimeter to area ratio, etc.). The BCM sample dataset was analyzed with six traditional ML methods, namely naive Bayes (NB), K-nearest neighbors (KNN), decision trees (DT), random forests (RF), support vector machines (SVM), and artificial neural networks (ANN), and the most effective algorithms per sample subsets were: (i) RF (86.8%) for Campos, Santos, and Ceará Basins; (ii) NB (87.2%) for Campos with Santos Basins; (iii) SVM (86.9%) for Campos with Ceará Basins; and (iv) SVM (87.8%) for only Campos Basin. The OSH can assist in different concerns (general public, social, economic, political, ecological, and scientific) related to petroleum exploration and production activities, serving as an important aid in discovering new offshore exploratory frontiers, avoiding legal penalties on oil-seep events, supporting oceanic monitoring systems, and providing valuable information to environmental studies. Full article
(This article belongs to the Section Marine Environmental Science)
Show Figures

Figure 1

Figure 1
<p>Location of the explored oil slicks in Campeche Bay (Mexican Gulf of Mexico—GMex): oil spills (green circles) and oil seeps (yellow circles).</p>
Full article ">Figure 2
<p>Fourteen investigated sedimentary basins on the Brazilian Continental Margin (BCM).</p>
Full article ">Figure 3
<p>Main Oil-Slick Hub (OSH) methodological steps.</p>
Full article ">Figure 4
<p>Oil-Slick Hub (OSH) main screen (top panel) and map with roads and street names (bottom panel). The layer’s menu is shown on the top right.</p>
Full article ">Figure 5
<p>Coastal bathymetry (top panel) and exploration block shapefiles (bottom panel).</p>
Full article ">Figure 6
<p>Data system menu to search, display, and filter for satellite and oil-slick data.</p>
Full article ">Figure 7
<p>Satellite-image shapefile frame (red square) and its associated information menu (top panel). Display of a synthetic aperture radar (SAR) satellite image (bottom panel).</p>
Full article ">Figure 8
<p>Example of oil-slick shapefiles and satellite frame (red square) displayed after a combined georeferenced satellite-image search with a bathymetry layer (2250 m).</p>
Full article ">Figure 9
<p>Classification results of the 1st experiment that used the Oil-Slick Hub (OSH) to select variable subsets of the Gulf of Mexico (GMex) database applied to linear discriminant analysis (LDA); see <a href="#sec5dot1-jmse-11-01497" class="html-sec">Section 5.1</a>. (<b>A</b>) Overall accuracy. (<b>B</b>) Sensitivity. (<b>C</b>) Specificity. (<b>D</b>) Positive-predictive value. (<b>E</b>) Negative-predictive value. Data transformations: non-transformed (blue), cube root (red), and log10 (green). The investigated nineteen variable subsets are numbered: GMex-1 to GMex.19. Variable subsets were chosen from the three available attribute types: (i) acquisition aspects (8 variables: satellite name, beam mode, date, month, season, acquisition time, and if the image was acquired at daylight or nighttime); (ii) morphological information (17 variables, e.g., area, perimeter, perimeter to area ratio, etc.); and (iii) location attributes (3 variables: latitude, longitude, and bathymetry). Variable subset groups: (i) only acquisition aspects (GMex.1); (ii) all three attribute types, only morphological information with location attributes, or only location (GMex.1 to GMex.9); and (iii) only morphological information (GMex.10 to GMex.19). Different data transformation: none with cube (blue), none with log (red), and cube and log (green).</p>
Full article ">
48 pages, 26374 KiB  
Review
On the Interpretation of Synthetic Aperture Radar Images of Oceanic Phenomena: Past and Present
by Kazuo Ouchi and Takero Yoshida
Remote Sens. 2023, 15(5), 1329; https://doi.org/10.3390/rs15051329 - 27 Feb 2023
Cited by 4 | Viewed by 3626
Abstract
In 1978, the SEASAT satellite was launched, carrying the first civilian synthetic aperture radar (SAR). The mission was the monitoring of ocean: application to land was also studied. Despite its short operational time of 105 days, SEASAT-SAR provided a wealth of information on [...] Read more.
In 1978, the SEASAT satellite was launched, carrying the first civilian synthetic aperture radar (SAR). The mission was the monitoring of ocean: application to land was also studied. Despite its short operational time of 105 days, SEASAT-SAR provided a wealth of information on land and sea, and initiated many spaceborne SAR programs using not only the image intensity data, but also new technologies of interferometric SAR (InSAR) and polarimetric SAR (PolSAR). In recent years, artificial intelligence (AI), such as deep learning, has also attracted much attention. In the present article, a review is given on the imaging processes and analyses of oceanic data using SAR, InSAR, PolSAR data and AI. The selected oceanic phenomena described here include ocean waves, internal waves, oil slicks, currents, bathymetry, ship detection and classification, wind, aquaculture, and sea ice. Full article
(This article belongs to the Special Issue SAR, Interferometry and Polarimetry Applications in Geoscience)
Show Figures

Figure 1

Figure 1
<p>ERS-1 SAR image of the English Channel (scene center: 50.57°N, 1.17°W), showing several oceanic features. (Courtesy of ESA).</p>
Full article ">Figure 2
<p>SEASAT-SAR image of the North Sea in northern Scotland, where the island of Foula is on the left and the mainland on the right. The images of ocean waves can be seen propagating from left to right in the range direction, and those refracted around the islands in the enlarged images.</p>
Full article ">Figure 3
<p>Illustration of the imaging processes of range-traveling ocean waves by the tilt, weak hydrodynamic and foreshortening modulations.</p>
Full article ">Figure 4
<p>(<b>a</b>) Relative contribution to ocean wave image modulation in terms of the azimuth angle. (<b>b</b>) Illustration of the non-linear image modulation by velocity bunching.</p>
Full article ">Figure 5
<p>Measurement of directional wave spectrum using the method of weighted cross-spectra.</p>
Full article ">Figure 6
<p>RADARSAT-2 X-band SAR image of Chuuk Lagoon, Micronesia, showing the bright azimuth streaks caused by breaking waves along the coastal barrier.</p>
Full article ">Figure 7
<p><b>Top left</b>: Azimuth phases of a stationary point scatterer (CR: corner reflector) and the point scatterer moving in random motion. <b>Bottom left</b>: The corresponding point spread functions (PSF) of the stationary and moving CRs. <b>Right</b>: TerraSAR-X image showing stationary and moving corner reflectors.</p>
Full article ">Figure 8
<p>ENVISAT-ASAR and MODIS-Terra images of the South China Sea acquired on 31 July 2010 with 55 min time difference. Two tidally generated internal wave packets are separated by the distance proportional to the semi-diurnal tidal period [<a href="#B64-remotesensing-15-01329" class="html-bibr">64</a>].</p>
Full article ">Figure 9
<p>Imaging process of IWs by SAR. The varying surface roughness is caused by the circular motion of water particles as the internal wave propagates, resulting in the current converging and diverging areas. The current converging surface becomes rough compared with the diverging area. The image signatures depend on the depression (<b>left</b>) and elevation (<b>right</b>) of the pycnocline.</p>
Full article ">Figure 10
<p>Different image signatures of nonlinear internal waves [<a href="#B63-remotesensing-15-01329" class="html-bibr">63</a>]. <b>Top left</b> and <b>bottom right</b>: TerrasAR-X images over the Gulf of Main west of Cape Cod acquired, respectively, on the 23rd of June and 3rd of July, both in 2008. <b>Top right</b>: ERS-1 SAR image over the Mozambique Channel acquired on the 24th of September 2001. <b>Bottom left</b>: ENVISAT-ASAR image over the Andaman Sea acquired on the 18th of November 2006. The propagation direction is from right to left, except the ENVISAT-ASAR image, in which the interaction can also be seen between the two wave packets propagating predominantly from left to right.</p>
Full article ">Figure 11
<p>Estimated and theoretical IW phase velocities of selected points in <a href="#remotesensing-15-01329-f008" class="html-fig">Figure 8</a>. (<b>a</b>) Velocities estimated using the difference between the ENVISAT and MODIS images with a time difference of 55 min. (<b>b</b>) Velocities estimated using the IW packets with a time difference 12.42 h [<a href="#B64-remotesensing-15-01329" class="html-bibr">64</a>].</p>
Full article ">Figure 12
<p>An example for the analysis of SAR images of IWs.</p>
Full article ">Figure 13
<p>Upper: X-band SAR image of IWs and instrumented sites in Loch Linnhe, Scotland (5 × 2 km). Lower: Relative intensity modulation of multi-frequency AIRSAR images.</p>
Full article ">Figure 14
<p>Major oil spills by tankers from 1967 to 2021. (Courtesy of ITOPF 2021 [<a href="#B96-remotesensing-15-01329" class="html-bibr">96</a>]).</p>
Full article ">Figure 15
<p>ENVISAT-ASAR C-band image, showing the Hebei Spirit tanker oil spill off the west coast of Korea. The dark areas around the position of the collision (white circle) is the spilled oil. The image was acquired 4 days after the collision.</p>
Full article ">Figure 16
<p>From left to right: Natural oil seeps, oil slicks from off-shore oil-rigs, low wind and up-welling look-alikes, and illegal oil discharged by a ship identified by the automatic identification system (AIS) signals.</p>
Full article ">Figure 17
<p>General procedure for oil slick detection system.</p>
Full article ">Figure 18
<p>Deep learning consisting of input, multiple hidden neural network layers, and output layer.</p>
Full article ">Figure 19
<p>Prediction of oil movement, where <b>V</b><sub>oil</sub>, <b>V</b><sub>current</sub>, and <b>V</b><sub>wind</sub> are the velocity vectors of oil, current and wind, respectively. <span class="html-italic">Q</span> is the wind drift factor.</p>
Full article ">Figure 20
<p>(<b>A,B</b>) An example of the prediction of oil slick movement for the <span class="html-italic">Hebei Spirit</span> tanker oil spill off the west coast of Korea shown in <a href="#remotesensing-15-01329-f015" class="html-fig">Figure 15</a>. Oil particles are extracted from the ENVISAT-ASAR image and simulated and compared with the oil particles in the RADARSAT-1 SAR image. The top-right image shows the optimum wind drift factor in terms of wind speed [<a href="#B137-remotesensing-15-01329" class="html-bibr">137</a>].</p>
Full article ">Figure 21
<p>Principle of current velocity measurements by the Doppler shift. <b>Left</b>: the center of the azimuth return signal from a range moving scatterer is shifted. <b>Right</b>: The center of the spectrum over the sea surface is shifted from that of the stationary surface.</p>
Full article ">Figure 22
<p>From top left in clockwise direction: Geometry of ATI SAR, intensity image (8 km square) of the Gulf Stream boundary off Virginia Beach, USA, produced by the JPL-AIRSAR, InSAR phase image, and estimated current velocity. The survey vessel (arrowed) is used as a reference to compute the absolute interferometric phase and the current velocity [<a href="#B145-remotesensing-15-01329" class="html-bibr">145</a>,<a href="#B146-remotesensing-15-01329" class="html-bibr">146</a>].</p>
Full article ">Figure 23
<p>Illustrating the dependence of the backscatter radar cross section (RCS) through the surface roughness changes caused by varying surface currents.</p>
Full article ">Figure 24
<p>Flowchart for the bathymetry measurements by the RCS-based method (<b>left</b>) and ATI SAR-based approach (<b>right</b>).</p>
Full article ">Figure 25
<p><b>Left</b>: (<b>a</b>) The current velocity vector measured by the X-band ATI SAR to the north of the German island of Sylt (3.5 km square). (<b>b</b>) Depth map from echo soundings. (<b>c</b>) 78 selected reference depth points (white dots). (<b>d</b>) Depth map derived in combination of the reference depth and the ATI-derived current field [<a href="#B152-remotesensing-15-01329" class="html-bibr">152</a>]. <b>Right</b>: TerraSAR-X StripMap scene over the island of Jersey in the southern exit of the English Channel (<b>left</b>), and the estimated bathymetry from the dispersion relation of ocean waves is shown at the points over the bathymetry data by the European Marine Observation Data Network [<a href="#B162-remotesensing-15-01329" class="html-bibr">162</a>,<a href="#B163-remotesensing-15-01329" class="html-bibr">163</a>].</p>
Full article ">Figure 26
<p>Illustration of the microwave backscatter from the sea surface and a ship. While the backscatter from the sea surface is dominated by the single-bounce surface scattering, the backscattering from the ship is due to the double-bounce, volume and helix scattering as well as surface scattering.</p>
Full article ">Figure 27
<p>ALOS-PALSAR PLR images of Tokyo Bay, Japan, showing HH-, HV- and VV-polarizations from left to right. The image size is approximately 44 km and 28 km in the azimuth and range directions, respectively. The white rectangular box indicates the test site for detection of aquaculture.</p>
Full article ">Figure 28
<p>Illustrating the principle of CFAR for ship detection.</p>
Full article ">Figure 29
<p><b>Left</b>: Illustration of the principle of ship detection by the wavelet transform. <b>Right</b>: Detected ships by WT in the COSMO-SkyMed X-band HH-polarization SpotLight image of Tokyo Bay, Japan.</p>
Full article ">Figure 30
<p>Color composite representation of the ALOS-PALSAR PLR images shown in <a href="#remotesensing-15-01329-f027" class="html-fig">Figure 27</a>.</p>
Full article ">Figure 31
<p><b>Left</b>: TerraSAR-X HH-polarization SpotLight SAR image of Tokyo Bay, Japan, showing the detected and identified ships. <b>Right</b>: HH/VV-polarization images of the test area (upper) and the detection results (lower).</p>
Full article ">Figure 32
<p>ALOS-2 PALSAR HH-polarization image (4 December 2016) showing the ships detected by the ATM with <span class="html-italic">c</span> =2.5 and identified with AIS in the northern waters of Taiwan [<a href="#B196-remotesensing-15-01329" class="html-bibr">196</a>].</p>
Full article ">Figure 33
<p>Left: Flowchart for ship classification. Right: Flowchart for ship detection and classification by the GMV Spain.</p>
Full article ">Figure 34
<p>A part of the classification results by the feature-based template matching of the TerraSAR-X data shown in <a href="#remotesensing-15-01329-f031" class="html-fig">Figure 31</a>. The length (L) and width (W) in the brackets are the true values of AIS data [<a href="#B201-remotesensing-15-01329" class="html-bibr">201</a>].</p>
Full article ">Figure 35
<p>A form of CMOD GMF showing the NRCS for different wind azimuth angles of wind and wind speeds.</p>
Full article ">Figure 36
<p><b>Left</b>: ERS-1 SAR image off the northwest coast of the Shetland Islands. The solid lines are the wind direction with 180° directional ambiguity estimated from the image spectrum shown in the right. <b>Right</b>: Image spectrum computed by Fourier transform of a part of the left SAR image showing the wavelength (~300 m) of ocean waves in both the directions from the center. The ellipsoid shows the direction orthogonal to the wind streaks [<a href="#B210-remotesensing-15-01329" class="html-bibr">210</a>].</p>
Full article ">Figure 37
<p>RADASAT-SAR images of fish cargoes on the left, and fish traps on the right as bright features against dark background of the sea surface (4 February 2001) [<a href="#B220-remotesensing-15-01329" class="html-bibr">220</a>].</p>
Full article ">Figure 38
<p><b>Left</b>: ALOS-PALSARHH-polarization image of the laver cultivation area in Tokyo Bay, Japan. <b>Center</b>: Entropy image. <b>Right</b>: Photograph of the test site.</p>
Full article ">Figure 39
<p>Illustrating the radar backscatter on sea ice and open water [<a href="#B5-remotesensing-15-01329" class="html-bibr">5</a>].</p>
Full article ">Figure 40
<p>The total power (<b>left</b>) and composite color images (<b>right</b>) of the Beaufort Sea acquired by the NASA/JPL airborne AIRSAR (11/02/1988) [<a href="#B34-remotesensing-15-01329" class="html-bibr">34</a>].</p>
Full article ">
19 pages, 6773 KiB  
Article
Bathymetry Refinement over Seamount Regions from SAR Altimetric Gravity Data through a Kalman Fusion Method
by Yihao Wu, Junjie Wang, Yueqian Shen, Dongzhen Jia and Yu Li
Remote Sens. 2023, 15(5), 1288; https://doi.org/10.3390/rs15051288 - 26 Feb 2023
Cited by 1 | Viewed by 1875
Abstract
Seafloor topography over seamount areas is crucial for studying plate motions, seafloor seismicity, and seamount ecosystems. However, seamount bathymetry modeling is difficult due to the complex hydrodynamic environment, biodiversity, and scarcity of shipborne echo sounding data. The use of satellite altimeter-derived gravity data [...] Read more.
Seafloor topography over seamount areas is crucial for studying plate motions, seafloor seismicity, and seamount ecosystems. However, seamount bathymetry modeling is difficult due to the complex hydrodynamic environment, biodiversity, and scarcity of shipborne echo sounding data. The use of satellite altimeter-derived gravity data is a complementary way of bathymetry computation; in particular, the incorporation of synthetic aperture radar (SAR) altimeter data may be useful for seamount bathymetry modeling. Moreover, the widely used filtering method may have difficulty in combing different bathymetry data sets and may affect the quality of the computed bathymetry. To mitigate this issue, we introduce a Kalman fusion method for weighting and combining gravity-derived bathymetry data and the reference bathymetry model. Numerical experiments in the seamount regions over the Molloy Ridge show that the use of SAR-based altimetric gravity data improves the local bathymetry model, by a magnitude of 14.27 m, compared to the result without SAR data. In addition, the developed Kalman fusion method outperforms the traditionally used filtering method, and the bathymetry computed from the Kalman fusion method is improved by a magnitude of 9.34 m. Further comparison shows that our solution has improved quality compared to a recently released global bathymetry model, namely, GEBCO 2022 (GEBCO: General Bathymetric Chart of the Oceans), by a magnitude of 34.34 m. The bathymetry model in this study may be substituted for existing global bathymetry models for geophysical investigations over the target area. Full article
(This article belongs to the Special Issue Space-Geodetic Techniques II)
Show Figures

Figure 1

Figure 1
<p>Flow chart of bathymetry modeling from gravity anomaly based on the Kalman filter fusion method.</p>
Full article ">Figure 2
<p>(<b>a</b>) Geographical overview of the Arctic Ocean where the area noted with red star is the study area. The area in the black box shows the test area. (<b>b</b>) Distribution of shipborne sounding data. (<b>c</b>) Vertical gradient of the gravity anomaly.</p>
Full article ">Figure 3
<p>(<b>a</b>) Band-pass-filtered reference bathymetry and (<b>b</b>) band-pass-filtered gravity data. The dashed lines are two profiles passing through the Atla Seamount. The Atla seamount noted with area 1 and the Eistla seamount noted with area 2.</p>
Full article ">Figure 4
<p>Difference between the filtered reference model and computed bathymetry from altimetric gravity data along profiles 2.95°E (<b>a</b>) and 79.35°N (<b>b</b>) passing through the Atla Seamount.</p>
Full article ">Figure 5
<p>Error of the reference model and predicted bathymetry over (<b>a</b>) the Atla Seamount area along longitude 2.95°E and (<b>b</b>) the Eistla Seamount area along longitude 1.95°E; estimation of Kalman gain coefficients over (<b>c</b>) the Atla Seamount area along longitude 2.95°E and (<b>d</b>) the Eistla Seamount area along longitude 1.95°E.</p>
Full article ">Figure 6
<p>Differences between shipborne sounding data and (<b>a</b>) Model_SIO31 computed without the Kalman fusion method; (<b>b</b>) Model_kf_SIO31 calculated using the Kalman fusion method, and (<b>c</b>) GEBCO_2022. The dashed line represents a shipborne sounding track passing through the Atla Seamount area.</p>
Full article ">Figure 7
<p>Discrepancies between shipborne sounding data and various bathymetry models along a shipborne sounding track over the Atla Seamount area.</p>
Full article ">Figure 8
<p>Misfits between the bathymetry model calculated from (<b>a</b>) DTU15GRA, (<b>b</b>) DTU17GRA, (<b>c</b>) DTU21GRA, (<b>d</b>) SIO V23.1, (<b>e</b>) SIO V27.1, (<b>f</b>) SIO V29.1, (<b>g</b>) SIO V30.1, (<b>h</b>) SIO V31.1, and shipborne sounding data over the Atla Seamount area. (<b>i</b>) demonstrates the misfits between GEBCO_2022 and shipborne sounding data. Note: only DTU21GRA, SIO V31.1, SIO V30.1, and SIO V29.1 were computed with the SAR altimetry data derived from Sentinel-3A/B over this area.</p>
Full article ">Figure 9
<p>Misfits between shipborne sounding data and different bathymetry models along a shipborne sounding track over the Atla Seamount area.</p>
Full article ">Figure 10
<p>Misfits between the bathymetry model calculated from (<b>a</b>) DTU15GRA, (<b>b</b>) DTU17GRA, (<b>c</b>) DTU21GRA, (<b>d</b>) SIO V23.1, (<b>e</b>) SIO V27.1, (<b>f</b>) SIO V29.1, (<b>g</b>) SIO V30.1, (<b>h</b>) SIO V31.1, and shipborne sounding data over the Eistla Seamount area. (<b>i</b>) demonstrates the misfits between the GEBCO_2022 model and shipborne sounding data.</p>
Full article ">Figure 11
<p>Misfits between shipborne sounding data and different bathymetry models along a shipborne sounding track over the Eistla Seamount area.</p>
Full article ">
18 pages, 5957 KiB  
Article
High-Precision Inversion of Shallow Bathymetry under Complex Hydrographic Conditions Using VGG19—A Case Study of the Taiwan Banks
by Jiaxin Cui, Xiaowen Luo, Ziyin Wu, Jieqiong Zhou, Hongyang Wan, Xiaolun Chen and Xiaoming Qin
Remote Sens. 2023, 15(5), 1257; https://doi.org/10.3390/rs15051257 - 24 Feb 2023
Cited by 1 | Viewed by 2003
Abstract
Shallow bathymetry is important for ocean exploration, and the development of high-precision bathymetry inversion methods, especially for shallow waters with poor quality, is a major research aim. Synthetic aperture radar (SAR) image data benefit from a wide coverage, high measurement density, rapidity, and [...] Read more.
Shallow bathymetry is important for ocean exploration, and the development of high-precision bathymetry inversion methods, especially for shallow waters with poor quality, is a major research aim. Synthetic aperture radar (SAR) image data benefit from a wide coverage, high measurement density, rapidity, and low consumption but are limited by low accuracy. Alternatively, multibeam data have low coverage and are difficult to obtain but have a high measurement accuracy. In this paper, taking advantage of the complementary properties, we use SAR image data as the content map and multibeam images as the migrated style map, applying the VGG19 neural network (optimizing the loss function formula) for bathymetric inversion. The model was universal and highly accurate for bathymetric inversion of shallow marine areas, such as turbid water in Taiwan. There was a strong correlation between bathymetric inversion data and measured data (R2 = 0.8822; RMSE = 1.86 m). The relative error was refined by 9.22% over those of previous studies. Values for different bathymetric regions were extremely correlated in the region of 20–40 m. The newly developed approach is highly accurate over 20 m in the open ocean, providing an efficient, precise shallow bathymetry inversion method for complex hydrographic conditions. Full article
(This article belongs to the Special Issue Precision Orbit Determination of Satellites)
Show Figures

Figure 1

Figure 1
<p>Architecture of the CNN-based VGG19 pretraining model.</p>
Full article ">Figure 2
<p>Overview of the model process.</p>
Full article ">Figure 3
<p>Schematic diagram of the Taiwan Banks location, where A is the panoramic position of SAR image data and B is the position of multibeam survey line data.</p>
Full article ">Figure 4
<p>Flowchart of this experiment.</p>
Full article ">Figure 5
<p>The 2012 bathymetric inversion map. a and b are the location of the interception area.</p>
Full article ">Figure 6
<p>Original content image of area a (<b>left</b>) and water depth inversion image (<b>right</b>).</p>
Full article ">Figure 7
<p>Original image of area b (<b>left</b>) and water depth inversion image (<b>right</b>).</p>
Full article ">Figure 8
<p>Overall loss function curve.</p>
Full article ">Figure 9
<p>Content map loss function curve.</p>
Full article ">Figure 10
<p>The 2018 bathymetric inversion map.</p>
Full article ">Figure 11
<p>Water depth correlation analysis for 2012.</p>
Full article ">Figure 12
<p>Water depth correlation analysis for 2018.</p>
Full article ">Figure 13
<p>Trends in R<sup>2</sup> across different water depth intervals.</p>
Full article ">Figure 14
<p>Error analysis chart.</p>
Full article ">
26 pages, 10842 KiB  
Article
Insights for Estimating and Predicting Reservoir Sedimentation Using the RUSLE-SDR Approach: A Case of Darbandikhan Lake Basin, Iraq–Iran
by Arsalan Ahmed Othman, Salahalddin S. Ali, Sarkawt G. Salar, Ahmed K. Obaid, Omeed Al-Kakey and Veraldo Liesenberg
Remote Sens. 2023, 15(3), 697; https://doi.org/10.3390/rs15030697 - 24 Jan 2023
Cited by 7 | Viewed by 2345
Abstract
Soil loss (SL) and its related sedimentation in mountainous areas affect the lifetime and functionality of dams. Darbandikhan Lake is one example of a dam lake in the Zagros region that was filled in late 1961. Since then, the lake has received a [...] Read more.
Soil loss (SL) and its related sedimentation in mountainous areas affect the lifetime and functionality of dams. Darbandikhan Lake is one example of a dam lake in the Zagros region that was filled in late 1961. Since then, the lake has received a considerable amount of sediments from the upstream area of the basin. Interestingly, a series of dams have been constructed (13 dams), leading to a change in the sedimentation rate arriving at the main reservoir. This motivated us to evaluate a different combination of equations to estimate the Revised Universal Soil Loss Equation (RUSLE), Sediment Delivery Ratio (SDR), and Reservoir Sedimentation (RSed). Sets of Digital Elevation Model (DEM) gathered by the Shuttle Radar Topography Mission (SRTM), Tropical Rainfall Measuring Mission (TRMM), Harmonized World Soil Database (HWSD), AQUA eMODIS NDVI V6 data, in situ surveys by echo-sounding bathymetry, and other ancillary data were employed for this purpose. In this research, to estimate the RSed, five models of the SDR and the two most sensitive factors affecting soil-loss estimation were tested (i.e., rainfall erosivity (R) and cover management factor (C)) to propose a proper RUSLE-SDR model suitable for RSed modeling in mountainous areas. Thereafter, the proper RSed using field measurement of the bathymetric survey in Darbandikhan Lake Basin (DLB) was validated. The results show that six of the ninety scenarios tested have errors <20%. The best scenario out of the ninety is Scenario #18, which has an error of <1%, and its RSed is 0.46458 km3·yr−1. Moreover, this study advises using the Modified Fournier index (MIF) equations to estimate the R factor. Avoiding the combination of the Index of Connectivity (IC) model for calculating SDR and land cover for calculating the C factor to obtain better estimates is highly recommended. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the Darbandikhan basin.</p>
Full article ">Figure 2
<p>Changing the area of the catchment area of the stream sediments for the Darbandikhan dam over time.</p>
Full article ">Figure 3
<p>Correlation between rainfall data collected at the Sulaymaniyah meteorological station and the corresponding cell of the TRMM data; (<b>A</b>) all months and (<b>B</b>) mean of the months of the year for the period between September 1998 and August 2019 [<a href="#B79-remotesensing-15-00697" class="html-bibr">79</a>].</p>
Full article ">Figure 4
<p>The distribution of the (<b>A</b>) R-factor (Equation (5)) and (<b>B</b>) K-factor maps within the DLB for 2 May 2008.</p>
Full article ">Figure 5
<p>Boxplot shows the distributions of the ~5% random selected pixels from R-factor values for the six equations used in this study, which is very rough.</p>
Full article ">Figure 6
<p>The distribution of the (<b>A</b>) LS factors and (<b>B</b>) C factors maps within the DLB for the period until 2 May 2008.</p>
Full article ">Figure 7
<p>The very high C-factor values were distributed in the (<b>A</b>) landslide (<b>B</b>) and agricultural areas overlayed by the QuickBird image (R3:G2:B1).</p>
Full article ">Figure 8
<p>Boxplot shows the distributions of the ~5% random selected pixels from C-factor values for the three equations used in this study.</p>
Full article ">Figure 9
<p>The distribution of the (<b>A</b>) P factors and (<b>B</b>) RUSLE maps within the DLB for the period until 2 May 2008.</p>
Full article ">Figure 10
<p>Boxplot shows the distributions of the ~5% random selected pixels from the 18 RUSLE scenarios used in this study (outlier pixels were removed).</p>
Full article ">Figure 11
<p>The distribution of the (<b>A</b>) sediment delivery ratio and (<b>B</b>) reservoir sedimentation maps within the DLB for the period until 2 May 2008.</p>
Full article ">Figure 12
<p>Evaluation of the reservoir sedimentation scenarios tested with the actual sedimentation in the Darbandikhan Lake Basin.</p>
Full article ">Figure 13
<p>The average distribution of the (<b>A</b>) RUSLE maps and (<b>B</b>) reservoir sedimentation maps within the DLB for the period until 2019.</p>
Full article ">Figure 14
<p>Uncertainty plots show the mean value of the factors estimates (x-axis) against two standard deviations of the factors estimates (y-axis) for (<b>A</b>) R factor and (<b>B</b>) C factor models.</p>
Full article ">Figure 15
<p>Uncertainty plots show the mean value of the eighteen RUSLE scenarios (x-axis) against two standard deviations of the eighteen RUSLE scenarios (y-axis); 124 outlier pixels of the RUSLE were removed, which are &gt;5000 t·ha<sup>−1</sup>·y<sup>−1</sup> for (<b>A</b>) ~1% and (<b>B</b>) ~5% of the total pixels was used.</p>
Full article ">
21 pages, 15414 KiB  
Article
A New Global Bathymetry Model: STO_IEU2020
by Diao Fan, Shanshan Li, Jinkai Feng, Yongqi Sun, Zhenbang Xu and Zhiyong Huang
Remote Sens. 2022, 14(22), 5744; https://doi.org/10.3390/rs14225744 - 13 Nov 2022
Cited by 7 | Viewed by 1771
Abstract
To address the limitations in global seafloor topography model construction, a scheme is proposed that takes into account the efficiency of seafloor topography prediction, the applicability of inversion methods, the heterogeneity of seafloor environments, and the inversion advantages of sea surface gravity field [...] Read more.
To address the limitations in global seafloor topography model construction, a scheme is proposed that takes into account the efficiency of seafloor topography prediction, the applicability of inversion methods, the heterogeneity of seafloor environments, and the inversion advantages of sea surface gravity field element. Using the South China Sea as a study area, we analyzed and developed the methodology in modeling the seafloor topography, and then evaluated the feasibility and effectiveness of the modeling strategy. Based on the proposed modeling approach, the STO_IEU2020 global bathymetry model was constructed using various input data, including the SIO V29.1 gravity anomaly (GA) and vertical gravity gradient anomaly (VGG), as well as bathymetric data from multiple sources (single beam, multi-beam, seismic, Electronic Navigation Chart, and radar sensor). Five evaluation areas located in the Atlantic and Indian Oceans were used to assess the performance of the generated model. The results showed that 79%, 89%, 72%, 92% and 93% of the checkpoints were within the ±100 m range for the five evaluation areas, and with average relative accuracy better than 6%. The generated STO_IEU2020 model correlates well with the SIO V20.1 model, indicating that the proposed construction strategy for global seafloor topography is feasible. Full article
Show Figures

Figure 1

Figure 1
<p>The South China Sea Area is shown in the red box.</p>
Full article ">Figure 2
<p>Distribution of bathymetric data. (<b>a</b>) The black dots represnt bathymetric data from GEBCO_2019. (<b>b</b>) The black dots represnt bathymetric data from NGDC. (<b>c</b>) Distribution comparison between NGDC and GEBCO_2019, where the black dots represent data from GEBCO_2019 and the red dots represent data from NGDC.</p>
Full article ">Figure 3
<p>Distribution of shipborn depth data. (<b>a</b>) Distribution of control points and checkpoints; the red dots represent the control points and the black dots represent the checkpoints. (<b>b</b>) The black dots represent fusion results of control point and GEBCO_2019.</p>
Full article ">Figure 4
<p>Sea surface gravity data in the South China Sea. (<b>a</b>) Gravity anomaly, called GA_29.1; (<b>b</b>) Vertical gravity gradient anomaly, called VGG_29.1.</p>
Full article ">Figure 5
<p>ST models. (<b>a</b>) BAT_SCS model. (<b>b</b>) The colored part in the figure represents ST derived from gravity data, and the black dots represent the shipborne bathymetric data.</p>
Full article ">Figure 6
<p>ST models. (<b>a</b>) SIO V20.1 bathymetry model; (<b>b</b>) DTU18 bathymetry model; (<b>c</b>) ETOPO1 bathymetry model; and (<b>d</b>) BAT_VGG bathymetry model.</p>
Full article ">Figure 6 Cont.
<p>ST models. (<b>a</b>) SIO V20.1 bathymetry model; (<b>b</b>) DTU18 bathymetry model; (<b>c</b>) ETOPO1 bathymetry model; and (<b>d</b>) BAT_VGG bathymetry model.</p>
Full article ">Figure 7
<p>Distribution of checkpoints in the test areas. (<b>a</b>) Verification Sea Area A; (<b>b</b>) Verification Sea Area B.</p>
Full article ">Figure 8
<p>SIO V29.1 gravity anomaly.</p>
Full article ">Figure 9
<p>SIO V29.1 vertical gravity gradient anomaly.</p>
Full article ">Figure 10
<p>The distribution of shipborne depth data.</p>
Full article ">Figure 11
<p>STO_IEU2020 global ST model.</p>
Full article ">Figure 12
<p>Schematic diagram of assessing sea areas.</p>
Full article ">Figure 13
<p>Distribution of checkpoints in the sea area. (<b>a</b>) Sea Area 1; (<b>b</b>) Sea Area 2; (<b>c</b>) Sea Area 3; (<b>d</b>) Sea Area 4; and (<b>e</b>) Sea Area 5.</p>
Full article ">Figure 13 Cont.
<p>Distribution of checkpoints in the sea area. (<b>a</b>) Sea Area 1; (<b>b</b>) Sea Area 2; (<b>c</b>) Sea Area 3; (<b>d</b>) Sea Area 4; and (<b>e</b>) Sea Area 5.</p>
Full article ">Figure 14
<p>Number of checkpoints in different difference ranges (<b>a</b>) Sea Area 1; (<b>b</b>) Sea Area 2; (<b>c</b>) Sea Area 3; (<b>d</b>) Sea Area 4; and (<b>e</b>) Sea Area 5.</p>
Full article ">Figure 15
<p>Difference between the SIO V20.1 and STO_IEU2020 at different ranges (<b>a</b>) Sea Area 1; (<b>b</b>) Sea Area 2; (<b>c</b>) Sea Area 3; (<b>d</b>) Sea Area 4; and (<b>e</b>) Sea Area 5.</p>
Full article ">
17 pages, 8655 KiB  
Article
Shallow Sea Topography Detection from Multi-Source SAR Satellites: A Case Study of Dazhou Island in China
by Longyu Huang, Junmin Meng, Chenqing Fan, Jie Zhang and Jingsong Yang
Remote Sens. 2022, 14(20), 5184; https://doi.org/10.3390/rs14205184 - 17 Oct 2022
Cited by 1 | Viewed by 2548
Abstract
Accurate measurement of underwater topography in the coastal zone is essential for human marine activities, and the synthetic aperture radar (SAR) presents a completely new solution. However, underwater topography detection using a single SAR image is vulnerable to the interference of sea state [...] Read more.
Accurate measurement of underwater topography in the coastal zone is essential for human marine activities, and the synthetic aperture radar (SAR) presents a completely new solution. However, underwater topography detection using a single SAR image is vulnerable to the interference of sea state and sensor noise, which reduces the detection accuracy. A new underwater topography detection method based on multi-source SAR (MSSTD) was proposed in this study to improve the detection precision. GF-3, Sentinel-1, ALOS PALSAR, and ENVISAT ASAR data were used to verify the sea area of Dazhou Island. The detection result was in good agreement with the chart data (MAE of 2.9 m and correlation coefficient of 0.93), and the detection accuracy was improved over that of a single SAR image. GF-3 image with 3 m spatial resolution performed best in bathymetry among the four SAR images. Additionally, the resolution of the SAR image had greater influence on bathymetry compared with polarization and radar band. The ability of MSSTD has been proved in our work. Collaborative multi-source satellite observation is a feasible and effective scheme in marine research, but its application potential in underwater topography detection still requires further exploration. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of the study area, the detection area is the right figure. The left figure was produced from ArcGIS 10.2, and the right figure was produced from Google Earth.</p>
Full article ">Figure 2
<p>The pre-processed SAR image of the Dazhou Island: (<b>a</b>) GF-3 HH-polarized; (<b>b</b>) Sentinel-1 VV-polarized; (<b>c</b>) ALOS PALSAR HH-polarized; (<b>d</b>) ENVISAT ASAR HH-polarized. The corresponding clipped images were shown as (<b>e</b>–<b>h</b>).</p>
Full article ">Figure 3
<p>The chart data of the study area: (<b>a</b>) the water depths of sea area around Dazhou Island on the Chart Sever website, the selected depth points were in the red box; (<b>b</b>) collected water depths ranged from 0 to 90 m.</p>
Full article ">Figure 4
<p>The locations of sub-regions. The yellow boxes represented the sub-region locations of GF-3, and the red, green, and black boxes represented Sentinel-1, ALOS PALSAR, and ENVISAT ASAR, respectively.</p>
Full article ">Figure 5
<p>The wavelength results. The blank areas are land or wavelengths greater than the wavelength maximum. (<b>a</b>) GF-3, (<b>b</b>) Sentinel-1, (<b>c</b>) ALOS PALSAR, (<b>d</b>) ENVISAT ASAR.</p>
Full article ">Figure 6
<p>Bathymetry results. GF-3, Sentinel-1, ALOS PALSAR, and ENVISAT ASAR were shown in the (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 7
<p>Error of retrieval depths, GF-3, Sentinel-1, ALOS PALSAR, and ENVISAT ASAR were shown in the (<b>a</b>–<b>d</b>). The blue dots indicated that the MRE was less than 10%, and the red dots indicated that the MRE was greater than 50%.</p>
Full article ">Figure 8
<p>Integrated schematic of valid water depth, the green, blue, yellow, brown dots present the locations of valid water depth from ENVISAT ASAR, ALOS PALSAR, Sentinel-1, GF-3 images, respectively.</p>
Full article ">Figure 9
<p>The topography retrieved from MSSTD. It was produced by interpolation of the valid water depths. The upper left and center blank areas were land and Dazhou Island, respectively.</p>
Full article ">Figure 10
<p>Scatter plots of retrieved topography and chart data. The red dash line is the 1:1 line and red solid line is the linear fit line. The index R is the linear correlation coefficient.</p>
Full article ">Figure 11
<p>Line chart of bathymetry error of different FFT sizes.</p>
Full article ">Figure 12
<p>Error distribution in different water depth ranges. The black bar is the MAE and red bar is the MRE.</p>
Full article ">
17 pages, 30345 KiB  
Article
On the Effect of Interferences on X-Band Radar Wave Measurements
by Pavel Chernyshov, Katrin Hessner, Andrey Zavadsky and Yaron Toledo
Sensors 2022, 22(10), 3818; https://doi.org/10.3390/s22103818 - 18 May 2022
Cited by 2 | Viewed by 2549
Abstract
X-band radars are in growing use for various oceanographic purposes, providing spatial real-time information about sea state parameters, surface elevations, currents, and bathymetry. Therefore, it is very appealing to use such systems as operational aids to harbour management. In an installation of such [...] Read more.
X-band radars are in growing use for various oceanographic purposes, providing spatial real-time information about sea state parameters, surface elevations, currents, and bathymetry. Therefore, it is very appealing to use such systems as operational aids to harbour management. In an installation of such a remote sensing system in Haifa Port, consistent radially aligned spikes of brightness randomly distributed with respect to azimuth were identified. These streak noise patterns were found to be interfering with the common approach of oceanographic analysis. Harbour areas are regularly frequented with additional electromagnetic transmissions from other ship and land-based radars, which may serve as a source of such interference. A new approach is proposed for the filtering of such undesirable interference patterns from the X-band radar images. It was verified with comparison to in-situ measurements of a nearby wave buoy. Regardless of the actual source of the corresponding pseudo-wave energy, it was found to be crucial to apply such filtration in order to improve the performance of the standard oceanographic parameter retrieval algorithm. This results in better estimation of the mean sea state parameters towards lower values of the significant wave height. For the commercial WaMoSII system this enhancement was clearly apparent in the improvement of the built-in quality control criteria marks. The developed prepossessing procedure improves the robustness of the directional spectra estimation practically eliminating pseudo-wave energy components. It also extends the system’s capability to measure storm events earlier on, a fact that is of high importance for harbour operational decision making. Full article
(This article belongs to the Special Issue Advanced Remote Sensing Technologies in Ocean Observations)
Show Figures

Figure 1

Figure 1
<p>Google map showing the position of the WaMoSII radar system in Haifa port and the associated radar range of 5 km. The overlay radar image shows signatures of ships and breakwater as well as sea clutter. Furthermore the location of the reference buoy offshore is marked with a yellow pin near the southern cape of the bay.</p>
Full article ">Figure 2
<p>Examples of the WaMoSII X-Band radar images of a low (<math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mi>s</mi> </msub> <mo>≈</mo> <mn>0.6</mn> </mrow> </semantics></math> m, wind speed 3.2 m/s) and moderate sea states (<math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mi>s</mi> </msub> <mo>≈</mo> <mn>1</mn> </mrow> </semantics></math> m, wind speed 5.8 m/s), acquired at Haifa port given before (left column) and after (right column) streak filtration procedure application. Zoomed areas correspond to the analysis area of box 1 (red rectangle). Ships are clearly visible as a bright spots within the radar footprint.</p>
Full article ">Figure 3
<p>Comparison of the radar image sequence derived sea state parameters (<math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>T</mi> <mi>p</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>D</mi> <mi>p</mi> </msub> </semantics></math>), calculated for the analysis box 2 (see the box definition in <a href="#sensors-22-03818-f004" class="html-fig">Figure 4</a>) and the corresponding wave rider buoy measurements for the mention period of 26–29 May 2018 (times are given in UTC). Streak filtration procedure is not introduced to the preprocessing. Reliable data and insufficient sea clutter radar data are marked based on the inner WaMoSII quality control criteria.</p>
Full article ">Figure 4
<p>WaMoSII radar image acquired in Haifa bay, showing the main and lee breakwaters and location of the WaMoSII wave analysis areas at 2 km off the radar.</p>
Full article ">Figure 5
<p>Photo of FURUNO navigational radar Far 2117 XN-24AF antenna connected to the WaMoSII system situated inside the control tower building of the Haifa port. Ultrasonic wind sensor installed as a support part for the WaMoSII processing system. In the far field, the main breakwater of the harbor is visible.</p>
Full article ">Figure 6
<p>Geometry of the analysis box situation to understand the directional spikes component distribution in the spectral domain.</p>
Full article ">Figure 7
<p>Streaks spectra examples from box 1 (344° N) (<b>left panel</b>) and box 2 (14° N) (<b>right panel</b>). For the positioning of the corresponding analysis boxes on an actual radar image, please refer to the <a href="#sensors-22-03818-f004" class="html-fig">Figure 4</a>. Both spectra are measured at the calm sea state (<math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mi>s</mi> </msub> <mo>≈</mo> <mn>0.6</mn> </mrow> </semantics></math> m) 27 May 2018 04:02 a.m. (refer to <a href="#sensors-22-03818-f003" class="html-fig">Figure 3</a>).</p>
Full article ">Figure 8
<p>WaMoSII frequency spectra for insufficient sea clutter (upper row) marked as unreliable data and insufficient sea clutter marked as reliable signal (bottom row). The left column shows the standard results including artefact of the interferences. In right column the corresponding spectra after interference removal are shown. Streaks spectrum is well-pronounced in its theoretical location. (<b>a</b>) insufficient sea clutter (II) with artefacts before filtering. (<b>b</b>) insufficient sea clutter after streaks filtration. (<b>c</b>) insufficient sea clutter marked as reliable data (IR) with artefacts before filtering. (<b>d</b>) insufficient sea clutter marked as reliable data (IR) after streaks filtration.</p>
Full article ">Figure 9
<p>WaMoSII frequency spectra for reliable data marked as unreliable one (upper row) and reliable data (bottom row) before and after filtration. The left column shows the standard results including artefacts of the interferences. In right column the corresponding spectra after interference removal are shown. Streaks spectrum is still noticeable in its theoretical location. Right column panels demonstrate necessity of the streak filtration (weak streak patterns are efficiently removed). (<b>a</b>) Reliable data marked as one with insufficient sea clutter before filtering. (<b>b</b>) Reliable data marked as one with insufficient sea clutter after streaks’ filtration. (<b>c</b>) Reliable data marked correctly before filtration. (<b>d</b>) Reliable data marked correctly after streaks’ filtration.</p>
Full article ">Figure 10
<p>Comparison of the radar image sequence derived <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math>, calculated for the analysis box 2 and the corresponding wave rider buoy measurements for the mentioned period of the storm event before (<b>upper panel</b>) and after (<b>lower panel</b>) the streak filtration procedure. All the special cases, such as (II), (IR), (RI), and (RR) are marked in the time series (for the description of the cases refer also to the <a href="#sensors-22-03818-t002" class="html-table">Table 2</a>).</p>
Full article ">Figure 11
<p>Time series of the significant wave height before (<b>upper panel</b>) and after (<b>lower panel</b>) streak filtration given together the moving average and std with the 20 min window size.</p>
Full article ">Figure 12
<p>Time series of the peak period before (<b>upper panel</b>) and after (<b>lower panel</b>) streak filtration given together the moving average and std, calculated with the 20 min window size.</p>
Full article ">Figure 13
<p>Time series of the peak direction before (<b>upper panel</b>) and after (<b>lower panel</b>) streak filtration given together the moving average and std, calculated with the 20 min window size.</p>
Full article ">Figure 14
<p>Histograms of the mean values of the <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math> and its standard deviation given in a relative fraction to the total number of points, bin width for mean value is 0.05 m and for standard deviation 0.01 m.</p>
Full article ">Figure 15
<p>Normalized histograms of the mean values of the <math display="inline"><semantics> <msub> <mi>T</mi> <mi>p</mi> </msub> </semantics></math> and its standard deviation given in a relative fraction to the total number of points, bin width for mean value is 0.05 s and for standard deviation 0.01 s.</p>
Full article ">Figure 16
<p>Normalized histograms of the mean values of the <math display="inline"><semantics> <msub> <mi>D</mi> <mi>p</mi> </msub> </semantics></math> and its standard deviation given in a relative fraction to the total number of points, bin width for mean value is 1 deg and for standard deviation 1 deg.</p>
Full article ">
15 pages, 8390 KiB  
Technical Note
Depth Inversion from Wave Frequencies in Temporally Augmented Satellite Video
by Matthijs Gawehn, Rafael Almar, Erwin W. J. Bergsma, Sierd de Vries and Stefan Aarninkhof
Remote Sens. 2022, 14(8), 1847; https://doi.org/10.3390/rs14081847 - 12 Apr 2022
Cited by 2 | Viewed by 2294
Abstract
Optical satellite images of the nearshore water surface offer the possibility to invert water depths and thereby constitute the underlying bathymetry. Depth inversion techniques based on surface wave patterns can handle clear and turbid waters in a variety of global coastal environments. Common [...] Read more.
Optical satellite images of the nearshore water surface offer the possibility to invert water depths and thereby constitute the underlying bathymetry. Depth inversion techniques based on surface wave patterns can handle clear and turbid waters in a variety of global coastal environments. Common depth inversion algorithms require video from shore-based camera stations, UAVs or Xband-radars with a typical duration of minutes and at framerates of 1–2 fps to find relevant wave frequencies. These requirements are often not met by satellite imagery. In this paper, satellite imagery is augmented from a sequence of 12 images of Capbreton, France, collected over a period of ∼1.5 min at a framerate of 1/8 fps by the Pleiades satellite, to a pseudo-video with a framerate of 1 fps. For this purpose, a recently developed method is used, which considers spatial pathways of propagating waves for temporal video reconstruction. The augmented video is subsequently processed with a frequency-based depth inversion algorithm that works largely unsupervised and is openly available. The resulting depth estimates approximate ground truth with an overall depth bias of −0.9 m and an interquartile range of depth errors of 5.1 m. The acquired accuracy is sufficiently high to correctly predict wave heights over the shoreface with a numerical wave model and to find hotspots where wave refraction leads to focusing of wave energy that has potential implications for coastal hazard assessments. A more detailed depth inversion analysis of the nearshore region furthermore demonstrates the possibility to detect sandbars. The combination of image augmentation with a frequency-based depth inversion method shows potential for broad application to temporally sparse satellite imagery and thereby aids in the effort towards globally available coastal bathymetry data. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Pleiades satellite collecting 12 images of the field-site Capbreton, France. The observed area and its location (Lat°, Lon°) are depicted in the top right.</p>
Full article ">Figure 2
<p>First four frames of the reconstructed video at times <math display="inline"><semantics> <mrow> <mi mathvariant="normal">t</mi> <mn>0</mn> <mo>=</mo> <mn>0</mn> <mrow/> </mrow> </semantics></math> s to <math display="inline"><semantics> <mrow> <mrow> <mi mathvariant="normal">t</mi> <mn>0</mn> <mo>+</mo> <mn>3</mn> </mrow> <mo>=</mo> <mspace width="3.33333pt"/> <mn>3</mn> <mrow/> </mrow> </semantics></math> s. Wave movement is highlighted by zooming in on two example regions (white boxes), and looking at the difference with respect to <math display="inline"><semantics> <mrow> <mi mathvariant="normal">t</mi> <mn>0</mn> </mrow> </semantics></math> (color scale). Yellow, positive differences point out rising water levels due to incoming wave fronts. Red, negative differences point out the associated falling water levels at the back of the wave. For clarity, only differences &gt;10% are depicted.</p>
Full article ">Figure 3
<p>Comparison of an in situ measured (<b>left</b>, red ‘x’) variance density spectrum from a local buoy (<b>top right</b>) against a corresponding optical variance density spectrum from reconstructed satellite video (<b>bottom right</b>) of a representative area (<math display="inline"><semantics> <mrow> <mn>2</mn> <mo> </mo> <mi>km</mi> <mo> </mo> <mo>×</mo> <mo> </mo> <mn>2</mn> <mo> </mo> <mi>km</mi> <mrow/> </mrow> </semantics></math>) around the buoy location (<b>left</b>, white box). Both spectra are normalized to unit magnitude for comparison.</p>
Full article ">Figure 4
<p>Global one-component phase images (GOCPI) of dominant frequency components <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>0.09</mn> <mo>,</mo> <mn>0.11</mn> <mo>,</mo> <mrow> <mn>0.14</mn> </mrow> <mrow/> </mrow> </semantics></math> Hz in the reconstructed video. The phase images are naturally retrieved via the Dynamic Mode Decomposition as part of the depth inversion procedure [<a href="#B13-remotesensing-14-01847" class="html-bibr">13</a>]. In total, nine phase images are used for analysis, of which three are presented as example.</p>
Full article ">Figure 5
<p>Comparison of depths, <span class="html-italic">d</span>, from ground truth (<b>left</b>) against depths estimated from reconstructed satellite video (<b>centre</b>). Ground truth depth contours are superimposed on estimated depths for reference. Depths are indicated from red (shallow) to blue (deep) as of centre colour scale. The difference, <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>d</mi> </mrow> </semantics></math>, between estimated depths and ground truth, is presented in the right panel, with red/blue, respectively, denoting under-/overestimation of depth (<b>right</b> colour scale). Parts where <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>&gt;</mo> <mn>35</mn> <mrow/> </mrow> </semantics></math> m are masked and indicate the underwater canyon where waves are unaffected by the bathymetry.</p>
Full article ">Figure 6
<p>Direct comparison of inverted depths, <math display="inline"><semantics> <msub> <mi>d</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>v</mi> </mrow> </msub> </semantics></math> against ground truth, <math display="inline"><semantics> <msub> <mi>d</mi> <mn>0</mn> </msub> </semantics></math> (blue dots). The median is indicated in green and aims to approximate the black 1:1 line. The 25th–75th percentile is shaded red and superimposed on the 10th–90th percentile shaded orange.</p>
Full article ">Figure 7
<p>Predicted wave height distributions over ground truth bathymetry and estimated bathymetry of Capbreton using a numerical wave model: (<b>a</b>,<b>b</b>) ground-truth depths and estimated depths, respectively, used in the model, where red/blue indicate shallow/deep regions (top colorbar). Salt-and-pepper noise has been removed from (<b>b</b>) using a <math display="inline"><semantics> <mrow> <mn>3</mn> <mo>×</mo> <mn>3</mn> </mrow> </semantics></math> median filter; (<b>c</b>,<b>d</b>) significant wave height distribution associated to (<b>a</b>,<b>b</b>), respectively, where red/blue indicate high/low wave heights (bottom colorbar). The 35 m depth contour is superimposed to outline the location of the canyon. Hydrodynamic field conditions with <math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <mi>s</mi> </msub> <mo>=</mo> <mn>2</mn> <mrow/> </mrow> </semantics></math> m, <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>11</mn> <mrow/> </mrow> </semantics></math> s and wave direction = 315° measured by a local buoy (see <a href="#remotesensing-14-01847-f003" class="html-fig">Figure 3</a>) during satellite overpass are used as boundary forcing.</p>
Full article ">Figure 8
<p>The beach La Piste at Capbreton, France.</p>
Full article ">Figure 9
<p>Sandbars from depth inversion of a locally reconstructed video of the nearshore wave field: (<b>a</b>) reconstructed video with framerate of 2 fps; (<b>b</b>) satellite image showing the position of the sandbar; (<b>c</b>) depth estimates based on (<b>a</b>); For reference, dashed black contours outline the position of the sandbar. and the solid black line indicates the position of the coastline.</p>
Full article ">
Back to TopTop