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
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (184)

Search Parameters:
Keywords = oil spill detection

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 853 KiB  
Article
Detection of Nitrate-Reducing/Denitrifying Bacteria from Contaminated and Uncontaminated Tallgrass Prairie Soil: Limitations of PCR Primers
by Samer M. AbuBakr, Fares Z. Najar and Kathleen E. Duncan
Microorganisms 2024, 12(10), 1981; https://doi.org/10.3390/microorganisms12101981 - 30 Sep 2024
Viewed by 279
Abstract
Contamination of soil by spills of crude oil and oilfield brine is known to affect the species composition and functioning of soil microbial communities. However, the effect of such contamination on nitrogen cycling, an important biogeochemical cycle in tallgrass prairie soil, is less [...] Read more.
Contamination of soil by spills of crude oil and oilfield brine is known to affect the species composition and functioning of soil microbial communities. However, the effect of such contamination on nitrogen cycling, an important biogeochemical cycle in tallgrass prairie soil, is less well known. Detecting nitrate-reducing (NR) and denitrifying (DN) bacteria via PCR amplification of the genes essential for these processes depends on how well PCR primers match the sequences of these bacteria. In this study, we enriched for NR and DN bacteria from oil/brine tallgrass prairie soil contaminated 5–10 years previously versus those cultured from uncontaminated soil, confirmed the capacity of 75 strains isolated from the enrichments to reduce nitrate and/nitrite, then screened the strains with primers specific to seven nitrogen cycle functional genes. The strains comprised a phylogenetically diverse group of NR and DN bacteria, with proportionately more γ-Proteobacteria in oil-contaminated sites and more Bacilli in brine-contaminated sites, suggesting some residual effect of the contaminants on the NR and DN species distribution. Around 82% of the strains shown to reduce nitrate/nitrite would not be identified as NR and DN bacteria by the battery of NR and DN primers used. Our results indicate an urgent need to expand the NR/DN functional gene primer database by first identifying novel NR/DN strains through their capacity to reduce nitrate/nitrite. Full article
(This article belongs to the Section Environmental Microbiology)
Show Figures

Figure 1

Figure 1
<p>Taxonomic Affiliation of Strains. The number of strains affiliated with different higher taxonomic level bacterial groups found in oil-, brine-, and prairie-soils. The 16S rRNA gene sequences of seventy-five strains were classified by the RDP Classifier tool. Brine: isolated from sites contaminated primarily by oil field brine but also by crude oil (10:1 vol/vol), G5 and G7. Oil: isolated from sites contaminated by crude oil, J6-F, J6-NF, and LF. Prairie: isolated from uncontaminated sites, G5P, G7P, J6P, and LFP. “All” indicates the distribution of the 75 strains isolated after enrichment in nitrate broth, and “NR/DN” indicates the distribution of 54 strains confirmed as nitrate/nitrite reducers after incubation in nitrate broth as described in Materials and Methods.</p>
Full article ">Figure 2
<p>Phylogenetic tree of the 16S rRNA gene of 75 bacterial strains isolated from tallgrass prairie soils with respect to related sequences. There were 29 OTUs (indicated in bold letters) at the 98% level of similarity. One strain representing each OTU was used for phylogenetic analysis. The number of strains contained in each OTU is included between parentheses next to the representative strain. The numbers of strains of NR (nitrate-reducing) bacteria, DN (nitrite-reducing) bacteria, and None (strains that did not reduce nitrate or nitrite) are also included next to the total number of strains included in the OTU. The tree is constructed from approximately 1400 bp 16S rRNA gene sequence using the neighbor-joining algorithm. One thousand bootstrap replications were performed; only values greater than 700 are shown. (<b>a</b>) (Bar: 0.02 nucleotide substitutions per nucleotide) shows the γ-Proteobacteria where the sequence of <span class="html-italic">Ensifer adhaerens</span> strain WJB69 (KU877644) was included as the outgroup, while (<b>b</b>) (Bar: 0.03 nucleotide substitutions per nucleotide) shows the remaining groups where the sequence of <span class="html-italic">Archaeoglobus infectus</span> DSM 18877 strain Arc51 (NR_028166) was included as the outgroup.</p>
Full article ">Figure 2 Cont.
<p>Phylogenetic tree of the 16S rRNA gene of 75 bacterial strains isolated from tallgrass prairie soils with respect to related sequences. There were 29 OTUs (indicated in bold letters) at the 98% level of similarity. One strain representing each OTU was used for phylogenetic analysis. The number of strains contained in each OTU is included between parentheses next to the representative strain. The numbers of strains of NR (nitrate-reducing) bacteria, DN (nitrite-reducing) bacteria, and None (strains that did not reduce nitrate or nitrite) are also included next to the total number of strains included in the OTU. The tree is constructed from approximately 1400 bp 16S rRNA gene sequence using the neighbor-joining algorithm. One thousand bootstrap replications were performed; only values greater than 700 are shown. (<b>a</b>) (Bar: 0.02 nucleotide substitutions per nucleotide) shows the γ-Proteobacteria where the sequence of <span class="html-italic">Ensifer adhaerens</span> strain WJB69 (KU877644) was included as the outgroup, while (<b>b</b>) (Bar: 0.03 nucleotide substitutions per nucleotide) shows the remaining groups where the sequence of <span class="html-italic">Archaeoglobus infectus</span> DSM 18877 strain Arc51 (NR_028166) was included as the outgroup.</p>
Full article ">
21 pages, 10576 KiB  
Article
Prediction and Management of the Groundwater Environmental Pollution Impact in Anning Refinery in Southern China
by Xiaoqi Fang, Shiyao Tang, Zhenru Niu and Juntao Tong
Water 2024, 16(19), 2713; https://doi.org/10.3390/w16192713 - 24 Sep 2024
Viewed by 403
Abstract
Anning Refinery, a large-scale joint venture in southern China, possesses significant potential in regard to polluting local groundwater environments due to its extensive petroleum raw materials. This study aims to mitigate the substantial risks associated with oil spills and prevent consequential groundwater pollution [...] Read more.
Anning Refinery, a large-scale joint venture in southern China, possesses significant potential in regard to polluting local groundwater environments due to its extensive petroleum raw materials. This study aims to mitigate the substantial risks associated with oil spills and prevent consequential groundwater pollution by developing a robust groundwater flow model using the MODFLOW module in GMS software that aligns closely with natural and pumping test conditions. Furthermore, by integrating the MT3DMS model, a groundwater solute transport model is constructed and calibrated using sodium chloride tracer dispersion data. Notably, the wax hydrocracking unit and aviation coal finished product tank area are identified as key pollution sources warranting attention. By considering local constraints such as karst collapse, ground subsidence, and single-well water output capacity, the study introduces a tailored groundwater pollution management model. The research simulates various scenarios of petroleum pollutant migration in groundwater and proposes multi-objective emergency response optimization plans. In Scenario 1, simulations show that petroleum pollutants migrate within the unconfined aquifer and enter the karst aquifer as low-concentration plumes over an extended period. Detection of these plumes in karst water monitoring wells indicates upstream unconfined aquifer contamination at higher concentrations, necessitating immediate activation of the nearest monitoring or emergency wells in both layers. Conversely, in Scenario 2, pollutants reside briefly in the unconfined aquifer before entering the karst aquifer at relatively higher concentrations. Here, low-efficiency pollutant discharge through unconfined aquifer monitoring wells prompts the activation of nearby karst aquifer monitoring or emergency wells for effective pollution control. This model underscores the necessity for proactive monitoring and validates the efficacy of coupled numerical modeling in understanding pollutant behavior, offering valuable insights into pollution control scenario assessments. In summary, the study emphasizes the importance of targeted monitoring and emergency protocols, demonstrating the benefits of integrated modeling approaches in industrial areas prone to pollution risks, and provides critical theoretical and practical guidance for groundwater protection and pollution management, offering transferable insights for similar industrial settings worldwide. Full article
Show Figures

Figure 1

Figure 1
<p>Hydrogeological profile.</p>
Full article ">Figure 2
<p>Distribution map of groundwater monitoring points in the refinery area.</p>
Full article ">Figure 3
<p>Conceptual model of the unconsolidated layer.</p>
Full article ">Figure 4
<p>Conceptualization of Boundary Conditions.</p>
Full article ">Figure 5
<p>Error bar illustration.</p>
Full article ">Figure 6
<p>Natural field fitting graph.</p>
Full article ">Figure 7
<p>Pollution source location distribution map.</p>
Full article ">Figure 8
<p>Contaminant plume on the 360th day in Scenario 1.</p>
Full article ">Figure 9
<p>Scenario 2, contaminant plume at 75 days.</p>
Full article ">Figure 10
<p>Migration of the pollution plume under continuous pumping conditions at JC02.</p>
Full article ">Figure 11
<p>Trend chart of the center concentration of pollutants at the pollution source when pumping directly from JC02.</p>
Full article ">Figure 12
<p>The movement of the pollution plume under continuous pumping conditions at the pollution source.</p>
Full article ">Figure 13
<p>Migration of the pollution plume under continuous pumping conditions at JC03.</p>
Full article ">
18 pages, 3626 KiB  
Article
Detection of Oil Spill in SAR Image Using an Improved DeepLabV3+
by Jiahao Zhang, Pengju Yang and Xincheng Ren
Sensors 2024, 24(17), 5460; https://doi.org/10.3390/s24175460 - 23 Aug 2024
Viewed by 437
Abstract
Oil spill SAR images are characterized by high noise, low contrast, and irregular boundaries, which lead to the problems of overfitting and insufficient capturing of detailed features of the oil spill region in the current method when processing oil spill SAR images. An [...] Read more.
Oil spill SAR images are characterized by high noise, low contrast, and irregular boundaries, which lead to the problems of overfitting and insufficient capturing of detailed features of the oil spill region in the current method when processing oil spill SAR images. An improved DeepLabV3+ model is proposed to address the above problems. First, the original backbone network Xception is replaced by the lightweight MobileNetV2, which significantly improves the generalization ability of the model while drastically reducing the number of model parameters and effectively addresses the overfitting problem. Further, the spatial and channel Squeeze and Excitation module (scSE) is introduced and the joint loss function of Bce + Dice is adopted to enhance the sensitivity of the model to the detailed parts of the oil spill area, which effectively solves the problem of insufficient capture of the detailed features of the oil spill area. The experimental results show that the mIOU and F1-score of the improved model in an oil spill region in the Gulf of Mexico reach 80.26% and 88.66%, respectively. In an oil spill region in the Persian Gulf, the mIOU and F1-score reach 81.34% and 89.62%, respectively, which are better than the metrics of the control model. Full article
(This article belongs to the Special Issue Applications of Synthetic-Aperture Radar (SAR) Imaging and Sensing)
Show Figures

Figure 1

Figure 1
<p>Structure of SAR image oil spill detection model.</p>
Full article ">Figure 2
<p>Structure of the scSE module.</p>
Full article ">Figure 3
<p>Structure of scSE–MobileNetV2: (<b>a</b>) Inverted residual block with a stride of 1; (<b>b</b>) Inverted residual block with a stride of 2.</p>
Full article ">Figure 4
<p>Structure of scSE–ASPP.</p>
Full article ">Figure 5
<p>Oil spill SAR images from the ALOS satellite: (<b>a</b>) with filename 10777_sat in dataset; (<b>b</b>) with filename 10794_sat; (<b>c</b>) with filename 11148_sat; (<b>d</b>) with filename 11064_sat; (<b>e</b>) with filename 11168_sat.</p>
Full article ">Figure 6
<p>Prediction results of different backbone networks in the Gulf of Mexico oil spill area: (<b>a</b>) SAR image; (<b>b</b>) Ground truth; (<b>c</b>) Xception–DeepLabV3+ model; (<b>d</b>) MobileNetV2–DeepLabV3+ model.</p>
Full article ">Figure 7
<p>Prediction results of different backbone networks in the Persian Gulf oil spill area: (<b>a</b>) SAR image; (<b>b</b>) Ground truth; (<b>c</b>) Xception–DeepLabV3+ model; (<b>d</b>) MobileNetV2–DeepLabV3+ model.</p>
Full article ">Figure 8
<p>Prediction results of different models in the Gulf of Mexico oil spill area: (<b>a</b>) SAR image; (<b>b</b>) Ground truth; (<b>c</b>) ABCNet; (<b>d</b>) CGNet; (<b>e</b>) DFANet; (<b>f</b>) LEDNet; (<b>g</b>) MANet; (<b>h</b>) UNet; (<b>i</b>) Ours.</p>
Full article ">Figure 9
<p>Prediction results of different models in the Persian Gulf oil spill area: (<b>a</b>) SAR image; (<b>b</b>) Ground truth; (<b>c</b>) ABCNet; (<b>d</b>) CGNet; (<b>e</b>) DFANet; (<b>f</b>) LEDNet; (<b>g</b>) MANet; (<b>h</b>) UNet; (<b>i</b>) Ours.</p>
Full article ">
32 pages, 7438 KiB  
Article
Monitoring of Spatio-Temporal Variations of Oil Slicks via the Collocation of Multi-Source Satellite Images
by Tran Vu La, Ramona-Maria Pelich, Yu Li, Patrick Matgen and Marco Chini
Remote Sens. 2024, 16(16), 3110; https://doi.org/10.3390/rs16163110 - 22 Aug 2024
Viewed by 753
Abstract
Monitoring oil drift by integrating multi-source satellite imagery has been a relatively underexplored practice due to the limited time-sampling of datasets. However, this limitation has been mitigated by the emergence of new satellite constellations equipped with both Synthetic Aperture Radar (SAR) and optical [...] Read more.
Monitoring oil drift by integrating multi-source satellite imagery has been a relatively underexplored practice due to the limited time-sampling of datasets. However, this limitation has been mitigated by the emergence of new satellite constellations equipped with both Synthetic Aperture Radar (SAR) and optical sensors. In this manuscript, we take advantage of multi-temporal and multi-source satellite imagery, incorporating SAR (Sentinel-1 and ICEYE-X) and optical data (Sentinel-2/3 and Landsat-8/9), to provide insights into the spatio-temporal variations of oil spills. We also analyze the impact of met–ocean conditions on oil drift, focusing on two specific scenarios: marine floating oil slicks off the coast of Qatar and oil spills resulting from a shipwreck off the coast of Mauritius. By overlaying oils detected from various sources, we observe their short-term and long-term evolution. Our analysis highlights the finding that changes in oil structure and size are influenced by strong surface winds, while surface currents predominantly affect the spread of oil spills. Moreover, to detect oil slicks across different datasets, we propose an innovative unsupervised algorithm that combines a Bayesian approach used to detect oil and look-alike objects with an oil contours approach distinguishing oil from look-alikes. This algorithm can be applied to both SAR and optical data, and the results demonstrate its ability to accurately identify oil slicks, even in the presence of oil look-alikes and under varying met–ocean conditions. Full article
(This article belongs to the Special Issue Marine Ecology and Biodiversity by Remote Sensing Technology)
Show Figures

Figure 1

Figure 1
<p>Footprints of multi-sensor and multi-temporal images for observing oil spills. (<b>a</b>) Offshore Qatar, as covered by Sentinel-1 (28 March 2021, 14:33:06), Sentinel-2 (27 March 2021, 06:56:21), and Sentinel-3 (28 March 2021, 06:34:43). (<b>b</b>) Mauritius Island, as covered by Sentinel-1 IW (10 August 2020, 01:37:55), Sentinel-1 EW (10 August 2020, 14:36:16), and ICEYE-X (11 August 2020, 11:12:41).</p>
Full article ">Figure 2
<p>Oil slicks with low NRCS (dark objects) observed on the extracted scenes of (<b>a</b>) a Sentinel-1 image, 28 March 2021, 14:33:06; and (<b>b</b>) an ICEYE-X image, 6 August 2020, 18:33:23.</p>
Full article ">Figure 3
<p>Oil slick observed on the extracted scene of a Sentinel-2 image, offshore Qatar, 3 September 2021, 06:56:21: (<b>a</b>) RGB image; (<b>b</b>) oil index (in dB) calculated from the averages of RGB bands.</p>
Full article ">Figure 4
<p>Oil slick observed on the Sentinel-3 image, offshore Qatar, 28 March 2021, 06:34:43: (<b>a</b>) Sentinel-3 OLCI tristimulus image (Sentinel-3 User Handbook); (<b>b</b>) T865 variable (in dB) from Sentinel-3 Level-2 data.</p>
Full article ">Figure 5
<p>Flowchart of oil-slick detection from Sentinel-1/ICEYE-X SAR, Sentinel-2/Landsat-8 optical, and Sentinel-3 visible/near-infrared data.</p>
Full article ">Figure 6
<p>An example of the HSBA algorithm’s results for a Sentinel-1 SAR image, 5 July 2021, 02:23:25. One can find a detailed description of the HSBA algorithm in [<a href="#B33-remotesensing-16-03110" class="html-bibr">33</a>]. The purple box displays the histogram of backscattering values for the complete scene, in which the bimodality is less noticeable, making it difficult to identify T and BG. The red box presents the backscattering value histogram for the areas selected by HSBA, where oil slick is present, clearly highlighting a bimodal behavior. The green box is a histogram of the sea’s backscattering value.</p>
Full article ">Figure 7
<p>Floating-oil-slick detection for Case #Q1, 27–28 March 2021. (<b>a</b>–<b>c</b>) Extracted scenes from Sentinel-2 (27 March, 06:56:21), Sentinel-3 (28 March, 06:34:43), and Sentinel-1 (28 March, 14:33:06) images, respectively. (<b>Left</b>) Sentinel-2 RGB, Sentinel-3 OLCI tristimulus, and Sentinel-1 NRCS images, respectively. (<b>Right</b>) Oil slicks detected from Sentinel-2/3/1 images (<b>left</b>), respectively.</p>
Full article ">Figure 8
<p>Collocation of Sentinel-2/3/1 images (Case #Q1, 27–28 March 2021) for observations of oil-slick evolution in periods of about (<b>a</b>) 24 h (between Sentinel-2/3), (<b>b</b>) 8 h (between Sentinel-3/1), and (<b>c</b>) 32 h (between Sentinel-2/1).</p>
Full article ">Figure 9
<p>Surface wind and current speed and direction (indicated by the arrows), corresponding to the extracted Sentinel-2/3/1 scenes (<a href="#remotesensing-16-03110-f007" class="html-fig">Figure 7</a>a–c), respectively. (<b>i</b>) ERA-5 wind vectors for (<b>a</b>) 06:00, 27 March 2021; (<b>b</b>) 06:00, 28 March; and (<b>c</b>) 14:00, 28 March. (<b>ii</b>) CMEMS current vectors for (<b>d</b>) 06:30, 27 March; (<b>e</b>) 06:30, 28 March; and (<b>f</b>) 14:30, 28 March. (<b>iii</b>) Mean values of wind and current fields from 06:00, 27 March, to 14:00 28 March.</p>
Full article ">Figure 9 Cont.
<p>Surface wind and current speed and direction (indicated by the arrows), corresponding to the extracted Sentinel-2/3/1 scenes (<a href="#remotesensing-16-03110-f007" class="html-fig">Figure 7</a>a–c), respectively. (<b>i</b>) ERA-5 wind vectors for (<b>a</b>) 06:00, 27 March 2021; (<b>b</b>) 06:00, 28 March; and (<b>c</b>) 14:00, 28 March. (<b>ii</b>) CMEMS current vectors for (<b>d</b>) 06:30, 27 March; (<b>e</b>) 06:30, 28 March; and (<b>f</b>) 14:30, 28 March. (<b>iii</b>) Mean values of wind and current fields from 06:00, 27 March, to 14:00 28 March.</p>
Full article ">Figure 10
<p>Floating-oil-slick detection for Case #Q2, 5–6 July 2021. (<b>a</b>–<b>c</b>) Extracted scenes from Sentinel-1 (July 5, 02:23:25), Sentinel-2 (5 July, 06:56:21), and Landsat-8 (6 July, 06:58:26), respectively. (<b>Left</b>) Sentinel-1 NRCS, Sentinel-2 RGB, and Landsat-8 RGB, respectively. (<b>Right</b>) Oil slicks detected from Sentinel-1/2 and Landsat-8 images (<b>left</b>), respectively.</p>
Full article ">Figure 10 Cont.
<p>Floating-oil-slick detection for Case #Q2, 5–6 July 2021. (<b>a</b>–<b>c</b>) Extracted scenes from Sentinel-1 (July 5, 02:23:25), Sentinel-2 (5 July, 06:56:21), and Landsat-8 (6 July, 06:58:26), respectively. (<b>Left</b>) Sentinel-1 NRCS, Sentinel-2 RGB, and Landsat-8 RGB, respectively. (<b>Right</b>) Oil slicks detected from Sentinel-1/2 and Landsat-8 images (<b>left</b>), respectively.</p>
Full article ">Figure 11
<p>Collocation of Sentinel-1/2 and Landsat-8 images (Case #Q2, 5–6 July 2021) for observations of oil-slick evolution after about (<b>a</b>) 4 h (between Sentinel-1/2), (<b>b</b>) 24 h (between Sentinel-2 and Landsat-8), and (<b>c</b>) 28 h (between Sentinel-1 and Landsat-8).</p>
Full article ">Figure 11 Cont.
<p>Collocation of Sentinel-1/2 and Landsat-8 images (Case #Q2, 5–6 July 2021) for observations of oil-slick evolution after about (<b>a</b>) 4 h (between Sentinel-1/2), (<b>b</b>) 24 h (between Sentinel-2 and Landsat-8), and (<b>c</b>) 28 h (between Sentinel-1 and Landsat-8).</p>
Full article ">Figure 12
<p>Surface wind and current speed and direction (indicated by the arrows), corresponding to the extracted Sentinel-1/2 and Landsat-8 scenes (<a href="#remotesensing-16-03110-f010" class="html-fig">Figure 10</a>a–c), respectively. (<b>i</b>) ERA-5 wind vectors for (<b>a</b>) 02:00, 5 July 2021; (<b>b</b>) 07:00, 5 July; and (<b>c</b>) 07:00, 6 July. (<b>ii</b>) CMEMS current vectors on (<b>d</b>) 02:30, 5 July; (<b>e</b>) 07:30, 5 July; and (<b>f</b>) 07:30, 6 July. (<b>iii</b>) Mean values of wind and current fields from 02:00, 5 July, to 09:00, 6 July.</p>
Full article ">Figure 12 Cont.
<p>Surface wind and current speed and direction (indicated by the arrows), corresponding to the extracted Sentinel-1/2 and Landsat-8 scenes (<a href="#remotesensing-16-03110-f010" class="html-fig">Figure 10</a>a–c), respectively. (<b>i</b>) ERA-5 wind vectors for (<b>a</b>) 02:00, 5 July 2021; (<b>b</b>) 07:00, 5 July; and (<b>c</b>) 07:00, 6 July. (<b>ii</b>) CMEMS current vectors on (<b>d</b>) 02:30, 5 July; (<b>e</b>) 07:30, 5 July; and (<b>f</b>) 07:30, 6 July. (<b>iii</b>) Mean values of wind and current fields from 02:00, 5 July, to 09:00, 6 July.</p>
Full article ">Figure 13
<p>Floating-oil-slick detection for Case #Q3, 3 September 2021. (<b>a</b>,<b>c</b>) Three extracted scenes (#S1–3, <b>left</b>–<b>right</b>) from the Sentinel-1 NRCS (3 September, 02:23:54) and Sentinel-2 RGB (3 September, 06:56:21) images, respectively. (<b>b</b>,<b>d</b>) Oil slicks detected from the Sentinel-1/2 scenes #S1–3, respectively.</p>
Full article ">Figure 13 Cont.
<p>Floating-oil-slick detection for Case #Q3, 3 September 2021. (<b>a</b>,<b>c</b>) Three extracted scenes (#S1–3, <b>left</b>–<b>right</b>) from the Sentinel-1 NRCS (3 September, 02:23:54) and Sentinel-2 RGB (3 September, 06:56:21) images, respectively. (<b>b</b>,<b>d</b>) Oil slicks detected from the Sentinel-1/2 scenes #S1–3, respectively.</p>
Full article ">Figure 14
<p>Collocation of Sentinel-1/2 images (Case #Q3, 3 September 2021) for observations of oil-slick evolution after about 4 h. (<b>a</b>–<b>c</b>) Oil slicks detected from the extracted Sentinel-1/2 scenes #S1–3 (<a href="#remotesensing-16-03110-f013" class="html-fig">Figure 13</a>), respectively.</p>
Full article ">Figure 14 Cont.
<p>Collocation of Sentinel-1/2 images (Case #Q3, 3 September 2021) for observations of oil-slick evolution after about 4 h. (<b>a</b>–<b>c</b>) Oil slicks detected from the extracted Sentinel-1/2 scenes #S1–3 (<a href="#remotesensing-16-03110-f013" class="html-fig">Figure 13</a>), respectively.</p>
Full article ">Figure 15
<p>(<b>Left</b>–<b>Right</b>) Surface wind and current speed and direction (indicated by the arrows), corresponding to the extracted Sentinel-1/2 scenes #S1–3, respectively. (<b>a</b>) ERA-5 wind vectors for 02:00, 3 September 2021. (<b>b</b>) CMEMS current vectors for 02:30, 3 September 2021. (<b>c</b>) Mean values of wind and current fields from 02:00 to 07:00, 3 September.</p>
Full article ">Figure 15 Cont.
<p>(<b>Left</b>–<b>Right</b>) Surface wind and current speed and direction (indicated by the arrows), corresponding to the extracted Sentinel-1/2 scenes #S1–3, respectively. (<b>a</b>) ERA-5 wind vectors for 02:00, 3 September 2021. (<b>b</b>) CMEMS current vectors for 02:30, 3 September 2021. (<b>c</b>) Mean values of wind and current fields from 02:00 to 07:00, 3 September.</p>
Full article ">Figure 16
<p>Oil-slick observation on Sentinel-1 IW/EW and ICEYE-X images, offshore Mauritius, 10–11 August 2020. (<b>a</b>–<b>c</b>) Extracted scenes corresponding to the MV Wakashio oil spill, from Sentinel-1 IW (10 August, 01:37:55); Sentinel-1 EW (10 August, 14:36:16); and ICEYE-X (11 August, 11:12:41), respectively. (<b>d</b>–<b>f</b>) Oil spill, as detected by HSBA from the extracted scenes (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 17
<p>Surface wind and current speed and direction (indicated by the arrows), corresponding to the extracted Sentinel-1 IW, EW, and ICEYE-X scenes (<a href="#remotesensing-16-03110-f016" class="html-fig">Figure 16</a>a–c), respectively. (<b>i</b>) ERA-5 wind vectors (<b>a</b>) 01:00, 10 August 2020, and (<b>b</b>) 14:00 and (<b>c</b>) 11:00, 11 August. (<b>ii</b>) CMEMS current vectors on (<b>d</b>) 01:30, 10 August, and (<b>e</b>) 14:30 and (<b>f</b>) 11:30, 11 August. (<b>iii</b>) Mean values of wind and current fields from 01:00, 10 August, to 11:00, 11 August.</p>
Full article ">Figure 18
<p>(<b>i</b>,<b>ii</b>) Comparison between the oil slicks identified from the Sentinel-2 (scene S#3–<a href="#remotesensing-16-03110-f013" class="html-fig">Figure 13</a>c) and Sentinel-1 (scene S#2–<a href="#remotesensing-16-03110-f013" class="html-fig">Figure 13</a>a) images, respectively, for Case Q#3, 3 September 2021, by only HSBA (<a href="#remotesensing-16-03110-f018" class="html-fig">Figure 18</a>i–a, <a href="#remotesensing-16-03110-f018" class="html-fig">Figure 18</a>ii–d), those identified from HSBA plus oil contour (<a href="#remotesensing-16-03110-f018" class="html-fig">Figure 18</a>i–b, <a href="#remotesensing-16-03110-f018" class="html-fig">Figure 18</a>ii–e), and those manually segmented (<a href="#remotesensing-16-03110-f018" class="html-fig">Figure 18</a>i–c, <a href="#remotesensing-16-03110-f018" class="html-fig">Figure 18</a>ii–f).</p>
Full article ">Figure 19
<p>Comparison between the oil slicks (<b>a</b>) detected by HSBA plus oil contour and (<b>b</b>) those manually segmented, from the Sentinel-2 image, 5 July 2021, 06:56:21. (<b>c</b>) Difference between the detected pixels (<b>a</b>) and ground truth (<b>b</b>).</p>
Full article ">Figure 20
<p>Comparison between the oil slicks (<b>a</b>) detected by HSBA plus oil contour and (<b>b</b>) those manually segmented, from the Landsat-8 image, 6 July 2021, 06:58:26. (<b>c</b>) Difference between the detected pixels (<b>a</b>) and ground truth (<b>b</b>).</p>
Full article ">
22 pages, 9248 KiB  
Article
Developing a Comprehensive Oil Spill Detection Model for Marine Environments
by Farkhod Akhmedov, Rashid Nasimov and Akmalbek Abdusalomov
Remote Sens. 2024, 16(16), 3080; https://doi.org/10.3390/rs16163080 - 21 Aug 2024
Cited by 1 | Viewed by 1128
Abstract
Detecting oil spills in marine environments is crucial for avoiding environmental damage and facilitating rapid response efforts. In this study, we propose a robust method for oil spill detection leveraging state-of-the-art (SOTA) deep learning techniques. We constructed an extensive dataset comprising images and [...] Read more.
Detecting oil spills in marine environments is crucial for avoiding environmental damage and facilitating rapid response efforts. In this study, we propose a robust method for oil spill detection leveraging state-of-the-art (SOTA) deep learning techniques. We constructed an extensive dataset comprising images and frames extracted from video sourced from Google, significantly augmenting the dataset through frame extraction techniques. Each image is meticulously labeled to ensure high-quality training data. Utilizing the Yolov8 segmentation model, we trained our oil spill detection model to accurately identify and segment oil spills in ocean environments. K-means and Truncated Linear Stretching algorithms are combined with trained model weight to increase model detection accuracy. The model demonstrated exceptional performance, yielding high detection accuracy and precise segmentation capabilities. Our results indicate that this approach is highly effective for real-time oil spill detection, offering a promising tool for environmental monitoring and disaster management. In training metrics, the model reached over 97% accuracy in 100 epochs. In evaluation, model achieved its best detection rates by 94% accuracy in F1, 93.9% accuracy in Precision, and 95.5% [email protected] accuracy in Recall curves. Full article
Show Figures

Figure 1

Figure 1
<p>YOLO-v8 architecture. The design integrates a modified CSP Darknet53 framework, where the CSPLayer from YOLOv5 is substituted with the C2F module. To speed up computation, an SPPF module is utilized for pooling features into a standardized map. Each convolution module incorporates a batch normalization (BN) layer and SiLU activation.</p>
Full article ">Figure 2
<p>Proposed oil spill detection approach flowchart.</p>
Full article ">Figure 3
<p>Internet source oil spill image examples.</p>
Full article ">Figure 4
<p>Video frame generated image example with diverse scenarios.</p>
Full article ">Figure 5
<p>Lookalike images in ocean environment.</p>
Full article ">Figure 6
<p>Image annotation using LabelMe.</p>
Full article ">Figure 7
<p>Image processing example in single image case.</p>
Full article ">Figure 8
<p>RGB color change before and after image processing.</p>
Full article ">Figure 9
<p>Example detections while model training with oil spill images.</p>
Full article ">Figure 10
<p>Example detection while model validating with oil spill images.</p>
Full article ">Figure 11
<p>Data distribution with class match.</p>
Full article ">Figure 12
<p>Confusion matrix (<b>a</b>) and collegram (<b>b</b>).</p>
Full article ">Figure 13
<p>Model training and validation losses, metrics with comparative mAP scores.</p>
Full article ">Figure 14
<p>Line curve representation of metrics.</p>
Full article ">Figure 15
<p>Oil spill detection test results.</p>
Full article ">
16 pages, 5724 KiB  
Article
Automatic Identification of Sunken Oil in Homogeneous Information Perturbed Environment through Fusion Image Enhancement with Convolutional Neural Network
by Jinfeng Cao, Mingzhong Gao, Jihong Guo, Haichun Hao, Yongjun Zhang, Peng Liu and Nan Li
Sustainability 2024, 16(15), 6665; https://doi.org/10.3390/su16156665 - 4 Aug 2024
Viewed by 762
Abstract
With the development of the marine oil industry, leakage accidents are one of the most serious problems threatening maritime and national security. The spilt crude oil can float and sink in the water column, posing a serious long-term threat to the marine environment. [...] Read more.
With the development of the marine oil industry, leakage accidents are one of the most serious problems threatening maritime and national security. The spilt crude oil can float and sink in the water column, posing a serious long-term threat to the marine environment. High-frequency sonar detection is currently the most efficient method for identifying sunken oil. However, due to the complicated environment of the deep seabed and the interference of the sunken oil signals with homogeneous information, sonar detection data are usually difficult to interpret, resulting in low efficiency and a high failure rate. Previous works have focused on features designed by experts according to the detection environments and the identification of sunken oil targets regardless of the feature extraction step. To automatically identify sunken oil targets without a prior knowledge of the complex seabed conditions during the image acquisition process for sonar detection, a systematic framework is contrived for identifying sunken oil targets that combines image enhancement with a convolutional neural network (CNN) classifier for the final decision on sunken oil targets examined in this work. Case studies are conducted using datasets obtained from a sunken oil release experiment in an outdoor water basin. The experimental results show that (i) the method can effectively distinguish between the sunken oil target, the background, and the interference target; (ii) it achieved an identification accuracy of 83.33%, outperforming feature-based recognition systems, including SVM; and (iii) it provides important information about sunken oil such as the location of the leak, which is useful for decision-making in emergency response to oil spills at sea. This line of research offers a more robust and, above all, more objective option for the difficult task of automatically identifying sunken oils under complex seabed conditions. Full article
(This article belongs to the Section Waste and Recycling)
Show Figures

Figure 1

Figure 1
<p>Fusion image enhancement using the CNN method.</p>
Full article ">Figure 2
<p>Gaussian filtering. I–IV are identified targets from Gaussian filtering results: (<b>a</b>) original image; (<b>b</b>) filtered image.</p>
Full article ">Figure 3
<p>Histogram equalization.</p>
Full article ">Figure 4
<p>Binary image result after applying threshold segmentation.</p>
Full article ">Figure 5
<p>Extraction of the edge of the target area.</p>
Full article ">Figure 6
<p>Extraction of the target area.</p>
Full article ">Figure 7
<p>Schematic diagram of experimental setup.</p>
Full article ">Figure 8
<p>Sonar detection image.</p>
Full article ">Figure 9
<p>Illustrative examples of cropped image samples. (<b>a</b>) Submerged oil; (<b>b</b>) interference targets; (<b>c</b>) background.</p>
Full article ">Figure 10
<p>Loss curves of training and validation sets during training.</p>
Full article ">Figure 11
<p>Comparison of filters (<b>a</b>) before training; (<b>b</b>) after training.</p>
Full article ">Figure 12
<p>Precision and recall rates for each category.</p>
Full article ">Figure 13
<p>Examples of identification results for test images. Targets marked in red boxes are identified sunken oil. Targets marked in yellow boxes are false alarms. Digits 1-4 are target numbers. (<b>a</b>) Original images; (<b>b</b>) identification results.</p>
Full article ">
14 pages, 12697 KiB  
Communication
Deep Learning-Based Detection of Oil Spills in Pakistan’s Exclusive Economic Zone from January 2017 to December 2023
by Abdul Basit, Muhammad Adnan Siddique, Salman Bashir, Ehtasham Naseer and Muhammad Saquib Sarfraz
Remote Sens. 2024, 16(13), 2432; https://doi.org/10.3390/rs16132432 - 2 Jul 2024
Viewed by 964
Abstract
Oil spillages on a sea’s or an ocean’s surface are a threat to marine and coastal ecosystems. They are mainly caused by ship accidents, illegal discharge of oil from ships during cleaning and oil seepage from natural reservoirs. Synthetic-Aperture Radar (SAR) has proved [...] Read more.
Oil spillages on a sea’s or an ocean’s surface are a threat to marine and coastal ecosystems. They are mainly caused by ship accidents, illegal discharge of oil from ships during cleaning and oil seepage from natural reservoirs. Synthetic-Aperture Radar (SAR) has proved to be a useful tool for analyzing oil spills, because it operates in all-day, all-weather conditions. An oil spill can typically be seen as a dark stretch in SAR images and can often be detected through visual inspection. The major challenge is to differentiate oil spills from look-alikes, i.e., low-wind areas, algae blooms and grease ice, etc., that have a dark signature similar to that of an oil spill. It has been noted over time that oil spill events in Pakistan’s territorial waters often remain undetected until the oil reaches the coastal regions or it is located by concerned authorities during patrolling. A formal remote sensing-based operational framework for oil spills detection in Pakistan’s Exclusive Economic Zone (EEZ) in the Arabian Sea is urgently needed. In this paper, we report the use of an encoder–decoder-based convolutional neural network trained on an annotated dataset comprising selected oil spill events verified by the European Maritime Safety Agency (EMSA). The dataset encompasses multiple classes, viz., sea surface, oil spill, look-alikes, ships and land. We processed Sentinel-1 acquisitions over the EEZ from January 2017 to December 2023, and we thereby prepared a repository of SAR images for the aforementioned duration. This repository contained images that had been vetted by SAR experts, to trace and confirm oil spills. We tested the repository using the trained model, and, to our surprise, we detected 92 previously unreported oil spill events within those seven years. In 2020, our model detected 26 oil spills in the EEZ, which corresponds to the highest number of spills detected in a single year; whereas in 2023, our model detected 10 oil spill events. In terms of the total surface area covered by the spills, the worst year was 2021, with a cumulative 395 sq. km covered in oil or an oil-like substance. On the whole, these are alarming figures. Full article
(This article belongs to the Special Issue Deep Learning for Satellite Image Segmentation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Our study area in the Arabian sea, showing the extent of Pakistan’s Area of Responsibility (AOR). It contains 240,000 km<sup>2</sup> of Exclusive Economic Zone and 50,000 km<sup>2</sup> of extended continental shelf. Background: Google Maps.</p>
Full article ">Figure 2
<p>The flowchart illustrates the steps followed for the processing of Ground Range Detected High-resolution (GRDH) Sentinel-1 acquisitions, including orbit correction, thermal noise removal, radiometric calibration, conversion to decibels, terrain correction and speckle filtering.</p>
Full article ">Figure 3
<p>Schematic representation of ReU-Net architecture for semantic segmentation. The encoder features a ResNet-101 backbone with a <math display="inline"><semantics> <mrow> <mn>7</mn> <mo>×</mo> <mn>7</mn> </mrow> </semantics></math> convolutional layer, two batch normalization layers, ReLU activation, max-pooling and 33 residual blocks. In the decoder, adhering to a UNet framework, multi-scale features are integrated via skip connections (shown in dark-blue color). The number presented on the <span class="html-italic">Res-x</span> blocks represents the number of residual blocks cascaded within. The functionalities of the residual block (<span class="html-italic">Res-B</span>) and convolutional block (<span class="html-italic">Conv-B</span>) are expanded at the bottom-right and bottom-left corners of the figure, respectively. Each layer’s function is color-coded as per the legend.</p>
Full article ">Figure 4
<p>Three SAR images (<b>left</b>) from 110 test SAR images, with ground truth masks (<b>center</b>) and predicted class labels (<b>right</b>) detected by ReU-Net. In columns 2 and 3, black color shows the sea surface, green color shows land area, cyan color shows an oil spill area, red color is assigned to look-alikes and brown color shows ships. The dataset was prepared by Krestenitis et al. [<a href="#B20-remotesensing-16-02432" class="html-bibr">20</a>] from the MKLab ITI-CERTH, Greece.</p>
Full article ">Figure 5
<p>Three SAR images (<b>top row</b>) from the test set prepared by acquiring imagery over the Arabian sea along with predicted class labels (<b>bottom row</b>) containing potential oil spills in Pakistan territorial waters. Black color shows sea surface, cyan color shows oil spill and brown color shows ships.</p>
Full article ">Figure 6
<p>Yearly breakdown of the oil spill incidents in Pakistan’s Exclusive Economic Zone (EEZ) from January 2017 to December 2023. Overall, 92 incidents were identified. The yearly figure for each year is stated above each bar.</p>
Full article ">Figure 7
<p>(<b>Top</b>): Spatial spread of the oil spills detected in Pakistan’s EEZ from January 2017 to December 2023. (<b>Bottom</b>): The size of the spills (in terms of their length) and their distance from the shoreline.</p>
Full article ">
12 pages, 3531 KiB  
Communication
Experiment for Oil Spill Detection Based on Dual-Frequency QZSS Reflected Signals Using Drone-Borne GNSS-R
by Runqi Liu, Fan Gao, Cheng Jing, Xiao Li, Dongmei Song, Bin Wang, Huyu Sun, Yahui Kong, Zhenyao Zhong, Shuo Gu, Cong Yin and Weihua Bai
Remote Sens. 2024, 16(13), 2346; https://doi.org/10.3390/rs16132346 - 27 Jun 2024
Viewed by 671
Abstract
Oil spill detection plays an important role in marine environment protection. The technique of global navigation satellite system-reflectometry (GNSS-R) has the advantage of a short revisit time, which could help with timely cleanup of marine oil pollution. The conventional GNSS-R oil spill detection [...] Read more.
Oil spill detection plays an important role in marine environment protection. The technique of global navigation satellite system-reflectometry (GNSS-R) has the advantage of a short revisit time, which could help with timely cleanup of marine oil pollution. The conventional GNSS-R oil spill detection algorithm can resolve only the dielectric constant of oil based on power ratio measurements, while that of water cannot be realized. This is because the dielectric constant of water is much larger than that of oil such that the range of the equation used in the conventional algorithm is inadequate. To resolve this problem, we proposed a new algorithm containing a new equation with a larger scope, which has never been applied previously to GNSS-R oil spill detection. We derived a lookup method to resolve the dielectric constant of both oil and water. To validate our method, a drone-borne GNSS-R experiment based on dual-frequency QZSS reflection signals was conducted on 17 July 2023 using experimental pools simulating oil spills. Raw IF data in the L1 and L5 bands, collected using dual antennas and a data recorder, were processed using a software-defined receiver to deduce the power ratios and SNR of the GNSS signals. Results showed that the proposed algorithm is capable of resolving the dielectric constants of the reflected surface. In addition, the L5 signal was found to provide more detail and better contrast than the L1 C/A signal. Full article
Show Figures

Figure 1

Figure 1
<p>Dielectric constant–reflectance plots computed using Equations (1) and (2). The function images are multivalued. For example, the reflectance of 0.7 in the figure corresponds to the two dielectric constants of Equation (1). A program was used to retain the portion of the dielectric constant in the range of 0.01~100 based on the actual dielectric constant of oil and water. The lower part of the curve starts with a dielectric constant of 0.01 and the upper part of the curve has a dielectric constant of 100. The two curves intersect at the coordinate origin, which has been magnified in the diagram.</p>
Full article ">Figure 2
<p>Basic principles in distinguishing oil and water.</p>
Full article ">Figure 3
<p>(<b>a</b>) The drone hovering above the experimental pools whose serial number is marked in the figure received a steady stream of data, (<b>b</b>) the reference station, and (<b>c</b>) the drone and its accessories.</p>
Full article ">Figure 4
<p>Geometric relationship between the drone, satellite, and reflecting surface. The arrow represents the direction of signal propagation.</p>
Full article ">Figure 5
<p>(<b>a</b>) Locations and heights of the drone, and (<b>b</b>) the trajectory of the reflection points and the corresponding height of the drone. The drone was in a hovering (moving) state when the distribution of the colored dots is dense (sparse). The ‘*’ represents the contours of theexperimental pools. <a href="#remotesensing-16-02346-f006" class="html-fig">Figure 6</a> and <a href="#remotesensing-16-02346-f007" class="html-fig">Figure 7</a> follow the same logic.</p>
Full article ">Figure 6
<p>Inversion diagram of the dielectric constant of (<b>a</b>) the L1 C/A signal and (<b>b</b>) the L5 signal.</p>
Full article ">Figure 7
<p>Inversion diagram of the SNR of (<b>a</b>) the L1 C/A signal and (<b>b</b>) the L5 signal.</p>
Full article ">Figure 8
<p>(<b>a</b>) Relationship between the dielectric constant and the flight height of the drone for the L1 C/A signal, (<b>b</b>) relationship between the SNR and the flight height of the drone for the L1 C/A signal, (<b>c</b>) relationship between the dielectric constant and the flight height of the drone for the L5 signal, and (<b>d</b>) relationship between the SNR and the flight height of the drone for the L5 signal.</p>
Full article ">
20 pages, 4267 KiB  
Article
Comparative Bioremediation of Tetradecane, Cyclohexanone and Cyclohexane by Filamentous Fungi from Polluted Habitats in Kazakhstan
by Mariam Gaid, Wiebke Jentzsch, Hannah Beermann, Anne Reinhard, Mareike Meister, Ramza Berzhanova, Togzhan Mukasheva, Tim Urich and Annett Mikolasch
J. Fungi 2024, 10(6), 436; https://doi.org/10.3390/jof10060436 - 19 Jun 2024
Cited by 1 | Viewed by 916
Abstract
Studying the fates of oil components and their interactions with ecological systems is essential for developing comprehensive management strategies and enhancing restoration following oil spill incidents. The potential expansion of Kazakhstan’s role in the global oil market necessitates the existence of land-specific studies [...] Read more.
Studying the fates of oil components and their interactions with ecological systems is essential for developing comprehensive management strategies and enhancing restoration following oil spill incidents. The potential expansion of Kazakhstan’s role in the global oil market necessitates the existence of land-specific studies that contribute to the field of bioremediation. In this study, a set of experiments was designed to assess the growth and biodegradation capacities of eight fungal strains sourced from Kazakhstan soil when exposed to the hydrocarbon substrates from which they were initially isolated. The strains were identified as Aspergillus sp. SBUG-M1743, Penicillium javanicum SBUG-M1744, SBUG-M1770, Trichoderma harzianum SBUG-M1750 and Fusarium oxysporum SBUG-1746, SBUG-M1748, SBUG-M1768 and SBUG-M1769 using the internal transcribed spacer (ITS) region. Furthermore, microscopic and macroscopic evaluations agreed with the sequence-based identification. Aspergillus sp. SBUG-M1743 and P. javanicum SBUG-M1744 displayed remarkable biodegradation capabilities in the presence of tetradecane with up to a 9-fold biomass increase in the static cultures. T. harzianum SBUG-M1750 exhibited poor growth, which was a consequence of its low efficiency of tetradecane degradation. Monocarboxylic acids were the main degradation products by SBUG-M1743, SBUG-M1744, SBUG-M1750, and SBUG-M1770 indicating the monoterminal degradation pathway through β-oxidation, while the additional detection of dicarboxylic acid in SBUG-M1768 and SBUG-M1769 cultures was indicative of the fungus’ ability to undertake both monoterminal and diterminal degradation pathways. F. oxysporum SBUG-M1746 and SBUG-M1748 in the presence of cyclohexanone showed a doubling of the biomass with the ability to degrade the substrate almost completely in shake cultures. F. oxysporum SBUG-M1746 was also able to degrade cyclohexane completely and excreted all possible metabolites of the degradation pathway. Understanding the degradation potential of these fungal isolates to different hydrocarbon substrates will help in developing effective bioremediation strategies tailored to local conditions. Full article
(This article belongs to the Special Issue Bioremediation of Contaminated Soil by Fungi)
Show Figures

Figure 1

Figure 1
<p>An overview of the methodology and approaches used in the current study.</p>
Full article ">Figure 2
<p>Growth of fungal strains on 0.5% tetradecane and the remaining substrate after 7 days of incubation. (<b>a</b>) <span class="html-italic">Aspergillus</span> sp. St1, (<b>b</b>) <span class="html-italic">P. javanicum</span> St2, and (<b>c</b>) <span class="html-italic">T. harzianum</span> St3. Residual growth in cell controls was due to the pre-cultivation in malt broth. Bars presenting the mean values ± SD (n = 2).</p>
Full article ">Figure 3
<p>Yield biomass-substrate (Yx/S) of <span class="html-italic">Aspergillus</span> sp. St1, <span class="html-italic">P. javanicum</span> St2, <span class="html-italic">T. harzianum</span> St3, <span class="html-italic">F. oxysporum</span> St4, and <span class="html-italic">F. oxysporum</span> St5 after incubation with tetradecane or cyclohexanone. Data are the division results from values presented in <a href="#jof-10-00436-f002" class="html-fig">Figure 2</a>, <a href="#jof-10-00436-f004" class="html-fig">Figure 4</a> and <a href="#jof-10-00436-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 4
<p>Growth of fungal strains on 0.25% cyclohexanone and the remaining substrate after 7 days of incubation. (<b>a</b>) <span class="html-italic">F. oxysporum</span> St4, and (<b>b</b>) <span class="html-italic">F. oxysporum</span> St5. Residual growth in cell controls was due to the pre-cultivation in malt broth. Bars presenting the mean values ± SD (n = 2).</p>
Full article ">Figure 5
<p>Growth of fungal strains on 0.25% cyclohexane and the remaining substrate after 7 days of incubation with <span class="html-italic">F. oxysporum</span> St4. Residual growth in cell controls was due to the pre-cultivation in malt broth. Bars presenting the mean values ± SD (n = 2).</p>
Full article ">Figure 6
<p>Monoterminal and diterminal degradation pathways of tetradecane by <span class="html-italic">Aspergillus</span> sp. St1, <span class="html-italic">P. javanicum</span> St2, St8, <span class="html-italic">T. harzianum</span> St3, and <span class="html-italic">F. oxysporum</span> St6, St7. Structures that were not detected in the current study are marked by brackets. Pn<sub>T</sub> refers to the products’ number as per its appearance in <a href="#jof-10-00436-t002" class="html-table">Table 2</a> and <a href="#jof-10-00436-t003" class="html-table">Table 3</a>.</p>
Full article ">Figure 7
<p>Degradation pathways of cyclohexanone by <span class="html-italic">F. oxysporum</span> St4 and St5 compiled from pathways of Morgan and Watkinson [<a href="#B75-jof-10-00436" class="html-bibr">75</a>] and Mandal et al. [<a href="#B76-jof-10-00436" class="html-bibr">76</a>]. Pn<sub>con</sub> refers to the products’ number as per its appearance in <a href="#jof-10-00436-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 8
<p>Degradation pathways of cyclohexane by <span class="html-italic">F. oxysporum</span> St4. Pn<sub>C</sub> refers to the products’ number as per its appearance in <a href="#jof-10-00436-t005" class="html-table">Table 5</a>.</p>
Full article ">
18 pages, 14276 KiB  
Article
Marine Radar Oil Spill Detection Method Based on YOLOv8 and SA_PSO
by Jin Xu, Yuanyuan Huang, Haihui Dong, Lilin Chu, Yuqiang Yang, Zheng Li, Sihan Qian, Min Cheng, Bo Li, Peng Liu and Jianning Wu
J. Mar. Sci. Eng. 2024, 12(6), 1005; https://doi.org/10.3390/jmse12061005 - 16 Jun 2024
Cited by 2 | Viewed by 1081
Abstract
In the midst of a rapidly evolving economic landscape, the global demand for oil is steadily escalating. This increased demand has fueled marine extraction and maritime transportation of oil, resulting in a consequential and uneven surge in maritime oil spills. Characterized by their [...] Read more.
In the midst of a rapidly evolving economic landscape, the global demand for oil is steadily escalating. This increased demand has fueled marine extraction and maritime transportation of oil, resulting in a consequential and uneven surge in maritime oil spills. Characterized by their abrupt onset, rapid pollution dissemination, prolonged harm, and challenges in short-term containment, oil spill accidents pose significant economic and environmental threats. Consequently, it is imperative to adopt effective and reliable methods for timely detection of oil spills to minimize the damage inflicted by such incidents. Leveraging the YOLO deep learning network, this paper introduces a methodology for the automated detection of oil spill targets. The experimental data pre-processing incorporated denoise, grayscale modification, and contrast boost. Subsequently, realistic radar oil spill images were employed as extensive training samples in the YOLOv8 network model. The trained detection model demonstrated rapid and precise identification of valid oil spill regions. Ultimately, the oil films within the identified spill regions were extracted utilizing the simulated annealing particle swarm optimization (SA-PSO) algorithm. The proposed method for offshore oil spill survey presented here can offer immediate and valid data support for regular patrols and emergency reaction efforts. Full article
Show Figures

Figure 1

Figure 1
<p>The experimental marine radar remote-sensing image in the polar coordinate system.</p>
Full article ">Figure 2
<p>Experimental process.</p>
Full article ">Figure 3
<p>Data preprocessing scheme.</p>
Full article ">Figure 4
<p>The preprocess: (<b>a</b>) Cartesian coordinate system conversion; (<b>b</b>) noise reduction; (<b>c</b>) gray correction; and (<b>d</b>) local contrast enhancement.</p>
Full article ">Figure 4 Cont.
<p>The preprocess: (<b>a</b>) Cartesian coordinate system conversion; (<b>b</b>) noise reduction; (<b>c</b>) gray correction; and (<b>d</b>) local contrast enhancement.</p>
Full article ">Figure 5
<p>The YOLOv8 model architecture.</p>
Full article ">Figure 6
<p>Sample labeling method.</p>
Full article ">Figure 7
<p>SA_PSO algorithmic process.</p>
Full article ">Figure 8
<p>YOLOv8 model training curve: (<b>a</b>) the Precision-Confidence Curve; and (<b>b</b>) the Recall-Confidence Curve.</p>
Full article ">Figure 9
<p>YOLOv8 model prediction results: (<b>a</b>) preliminary detection result; and (<b>b</b>) the oil film regions were preserved.</p>
Full article ">Figure 10
<p>The oil spill segmentation results: (<b>a</b>) the segmentation result of SA_PSO; and (<b>b</b>) the real oil pill result; (<b>c</b>) The polar coordinates result.</p>
Full article ">Figure 10 Cont.
<p>The oil spill segmentation results: (<b>a</b>) the segmentation result of SA_PSO; and (<b>b</b>) the real oil pill result; (<b>c</b>) The polar coordinates result.</p>
Full article ">Figure 11
<p>The prediction results of different training models in red color: (<b>a</b>) YOLOv8s model; and (<b>b</b>) YOLOv8l model.</p>
Full article ">Figure 11 Cont.
<p>The prediction results of different training models in red color: (<b>a</b>) YOLOv8s model; and (<b>b</b>) YOLOv8l model.</p>
Full article ">Figure 12
<p>The YOLOv5n model prediction result in red color.</p>
Full article ">Figure 13
<p>The U-Net segmentation results: (<b>a</b>) Cartesian coordinate system; and (<b>b</b>) Polar coordinate system.</p>
Full article ">Figure 14
<p>Contrast enhanced image with no-wave-region information removed.</p>
Full article ">Figure 15
<p>Comparison of four machine learning threshold segmentation methods: (<b>a</b>) SA_PSO; (<b>b</b>) PSO; (<b>c</b>) FCM; and (<b>d</b>) K-means.</p>
Full article ">Figure 16
<p>Auxiliary verification methods for oil spill detection: (<b>a</b>) visible light data can verify the performance of marine radar oil spill monitoring; and (<b>b</b>) Infrared data can be used to verify the oil spill detection results of maritime radar at night.</p>
Full article ">
18 pages, 6129 KiB  
Article
OptimalNN: A Neural Network Architecture to Monitor Chemical Contamination in Cancer Alley
by Uchechukwu Leo Udeji and Martin Margala
J. Low Power Electron. Appl. 2024, 14(2), 33; https://doi.org/10.3390/jlpea14020033 - 10 Jun 2024
Viewed by 1439
Abstract
The detrimental impact of toxic chemicals, gas, and oil spills in aquatic environments poses a severe threat to plants, animals, and human life. Regions such as Cancer Alley exemplify the profound consequences of inadequately controlled chemical spills, significantly affecting the local community. Given [...] Read more.
The detrimental impact of toxic chemicals, gas, and oil spills in aquatic environments poses a severe threat to plants, animals, and human life. Regions such as Cancer Alley exemplify the profound consequences of inadequately controlled chemical spills, significantly affecting the local community. Given the far-reaching effects of these spills, it has become imperative to devise an efficient method for early monitoring, estimation, and cleanup, utilizing affordable and effective techniques. In this research, we explore the application of U-shaped neural Network (UNET) and U-shaped neural network transformer (UNETR) neural network models designed for the image segmentation of chemical and oil spills. Our models undergo training using the Commonwealth Scientific and Industrial Research Organization (CSIRO) dataset and the Oil Spill Detection dataset, employing a specialized filtering technique to enhance detection accuracy. We achieved training accuracies of 95.35% and 91% by applying UNET on the Oil Spill and the CSIRO datasets after 50 epochs of training, respectively. We also achieved a training accuracy of 75% by applying UNETR to the Oil Spill dataset. Additionally, we integrated mixed precision to expedite the model training process, thus maximizing data throughput. To further accelerate our implementation, we propose the utilization of the Field Programmable Gate Array (FPGA) architecture. The results obtained from our study demonstrate improvements in inference latency on FPGA. Full article
Show Figures

Figure 1

Figure 1
<p>RGB (Red, Green, Blue) image and dataset sample across the RGB and HSV (Hue, Saturation, Value) color spaces.</p>
Full article ">Figure 2
<p>U-Net architecture.</p>
Full article ">Figure 3
<p>UNETR architecture.</p>
Full article ">Figure 4
<p>Architecture of the vision transformer.</p>
Full article ">Figure 5
<p>Diagram showing the implementation of mixed precision using the PyTorch apex amp framework.</p>
Full article ">Figure 6
<p>Diagram showing the concept of the mixed precision architecture.</p>
Full article ">Figure 7
<p>Preprocessing architecture.</p>
Full article ">Figure 8
<p>Labeling convention used to create labels for the dataset.</p>
Full article ">Figure 9
<p>Plot of training vs. testing accuracy using the UNET model on the Oil Spill (<b>left</b>) and CSIRO datasets (<b>right</b>).</p>
Full article ">Figure 10
<p>Plot of training vs. testing accuracy using the UNETR model on the Oil Spill (<b>left</b>) and CSIRO datasets (<b>right</b>).</p>
Full article ">Figure 11
<p>Diagram showing the generated FPGA architecture using Vivado HLS.</p>
Full article ">Figure 12
<p>Architectures to handle the transformer component of our UNETR model.</p>
Full article ">Figure 13
<p>Architecture that handles the CNN components of our UNETR model.</p>
Full article ">Figure 14
<p>Distribution of data classes within the Oil Spill dataset.</p>
Full article ">Figure 15
<p>Confusion matrix showing the performance of our UNET model.</p>
Full article ">Figure 16
<p>Showing the test mask and the predicted mask.</p>
Full article ">Figure 17
<p>FPGA setup.</p>
Full article ">
24 pages, 18033 KiB  
Article
Full-Scale Aggregated MobileUNet: An Improved U-Net Architecture for SAR Oil Spill Detection
by Yi-Ting Chen, Lena Chang and Jung-Hua Wang
Sensors 2024, 24(12), 3724; https://doi.org/10.3390/s24123724 - 7 Jun 2024
Cited by 1 | Viewed by 748
Abstract
Oil spills are a major threat to marine and coastal environments. Their unique radar backscatter intensity can be captured by synthetic aperture radar (SAR), resulting in dark regions in the images. However, many marine phenomena can lead to erroneous detections of oil spills. [...] Read more.
Oil spills are a major threat to marine and coastal environments. Their unique radar backscatter intensity can be captured by synthetic aperture radar (SAR), resulting in dark regions in the images. However, many marine phenomena can lead to erroneous detections of oil spills. In addition, SAR images of the ocean include multiple targets, such as sea surface, land, ships, and oil spills and their look-alikes. The training of a multi-category classifier will encounter significant challenges due to the inherent class imbalance. Addressing this issue requires extracting target features more effectively. In this study, a lightweight U-Net-based model, Full-Scale Aggregated MobileUNet (FA-MobileUNet), was proposed to improve the detection performance for oil spills using SAR images. First, a lightweight MobileNetv3 model was used as the backbone of the U-Net encoder for feature extraction. Next, atrous spatial pyramid pooling (ASPP) and a convolutional block attention module (CBAM) were used to improve the capacity of the network to extract multi-scale features and to increase the speed of module calculation. Finally, full-scale features from the encoder were aggregated to enhance the network’s competence in extracting features. The proposed modified network enhanced the extraction and integration of features at different scales to improve the accuracy of detecting diverse marine targets. The experimental results showed that the mean intersection over union (mIoU) of the proposed model reached more than 80% for the detection of five types of marine targets including sea surface, land, ships, and oil spills and their look-alikes. In addition, the IoU of the proposed model reached 75.85 and 72.67% for oil spill and look-alike detection, which was 18.94% and 25.55% higher than that of the original U-Net model, respectively. Compared with other segmentation models, the proposed network can more accurately classify the black regions in SAR images into oil spills and their look-alikes. Furthermore, the detection performance and computational efficiency of the proposed model were also validated against other semantic segmentation models. Full article
(This article belongs to the Special Issue Intelligent SAR Target Detection and Recognition)
Show Figures

Figure 1

Figure 1
<p>Samples of SAR oil spill images from the MKLab dataset. Cyan, red, brown, green, and black correspond to oil spills, look-alikes, ships, land, and sea surface, respectively. (<b>a</b>) SAR images. (<b>b</b>) RGB masks.</p>
Full article ">Figure 2
<p>The collected SAR images corresponding to oil spill events in the Mediterranean Sea. The sampling dates from left to right are 25 February 2021, 5 September 2021, and 5 September 2021. Cyan, red, brown, green, and black correspond to oil spills, look-alikes, ships, land, and sea surface, respectively. (<b>a</b>) SAR images. (<b>b</b>) RGB masks.</p>
Full article ">Figure 3
<p>The architecture of the proposed FA-MobileUNet model.</p>
Full article ">Figure 4
<p>The U-Net structure proposed by Ronneberger et al. [<a href="#B20-sensors-24-03724" class="html-bibr">20</a>].</p>
Full article ">Figure 5
<p>The block structure of MobileNetv3.</p>
Full article ">Figure 6
<p>The structure of the CBAM.</p>
Full article ">Figure 7
<p>The structure of the ASPP module.</p>
Full article ">Figure 8
<p>Full-scale aggregation example of stage 4 (44 × 44) of the decoder layer in <a href="#sensors-24-03724-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 9
<p>The training process of the proposed FA-MobileUNet model using augmented MKLab dataset. (<b>a</b>) Loss. (<b>b</b>) Accuracy.</p>
Full article ">Figure 10
<p>The segmentation results of the 55th image in the MKLab dataset: (<b>a</b>) original SAR image, (<b>b</b>) the corresponding ground truth data, and results from (<b>c</b>) U-Net model, (<b>d</b>) LinkNet model, (<b>e</b>) PSPNet model, (<b>f</b>) DeepLabv2 model, (<b>g</b>) DeepLabv3+ model, (<b>h</b>) FA-MobileUNet model. Black, cyan, red, brown, and green represent the sea surface, oil spills, look-alikes, ships, and land, respectively.</p>
Full article ">Figure 10 Cont.
<p>The segmentation results of the 55th image in the MKLab dataset: (<b>a</b>) original SAR image, (<b>b</b>) the corresponding ground truth data, and results from (<b>c</b>) U-Net model, (<b>d</b>) LinkNet model, (<b>e</b>) PSPNet model, (<b>f</b>) DeepLabv2 model, (<b>g</b>) DeepLabv3+ model, (<b>h</b>) FA-MobileUNet model. Black, cyan, red, brown, and green represent the sea surface, oil spills, look-alikes, ships, and land, respectively.</p>
Full article ">Figure 11
<p>The segmentation results of 71st image in the MKLab dataset: (<b>a</b>) original SAR image, (<b>b</b>) the corresponding ground truth data, and results from (<b>c</b>) U-Net model, (<b>d</b>) LinkNet model, (<b>e</b>) PSPNet model, (<b>f</b>) DeepLabv2 model, (<b>g</b>) DeepLabv3+ model, (<b>h</b>) FA-MobileUNet model. Black, cyan, red, brown, and green represent the sea surface, oil spills, look-alikes, ships, and land, respectively.</p>
Full article ">Figure 12
<p>The segmentation results of 106th image in the MKLab dataset: (<b>a</b>) original SAR image, (<b>b</b>) the corresponding ground truth data, and results from (<b>c</b>) U-Net model, (<b>d</b>) LinkNet model, (<b>e</b>) PSPNet model, (<b>f</b>) DeepLabv2 model, (<b>g</b>) DeepLabv3+ model, (<b>h</b>) FA-MobileUNet model. Black, red, brown, and green represent the sea surface, look-alikes, ships, and land, respectively.</p>
Full article ">Figure 13
<p>The segmentation results of U-Net model with different modules: (<b>a</b>) original SAR image, (<b>b</b>) the corresponding ground truth data, and results from (<b>c</b>) U-Net model, (<b>d</b>) U-Net model with CBAM, (<b>e</b>) U-Net model with ASPP module, (<b>f</b>) U-Net with FA module. Black, cyan, red and brown represent the sea surface, oil spills, look-alikes and ships, respectively.</p>
Full article ">Figure 13 Cont.
<p>The segmentation results of U-Net model with different modules: (<b>a</b>) original SAR image, (<b>b</b>) the corresponding ground truth data, and results from (<b>c</b>) U-Net model, (<b>d</b>) U-Net model with CBAM, (<b>e</b>) U-Net model with ASPP module, (<b>f</b>) U-Net with FA module. Black, cyan, red and brown represent the sea surface, oil spills, look-alikes and ships, respectively.</p>
Full article ">Figure 14
<p>The incorrect ground truth data in the dataset. A1 and A2 are the 111th and 275th training images, respectively. Black, cyan, red, brown, and green represent the sea surface, oil spills, look-alikes, ships, and land, respectively. (<b>a</b>) SAR image. (<b>b</b>) Ground truth data.</p>
Full article ">Figure 15
<p>Example of the revised ground truth data in the dataset. B1 and B2 are the 140th and 157th training images, respectively. Black, cyan, red, brown, and green represent the sea surface, oil spills, look-alikes, ships, and land, respectively. (<b>a</b>) SAR image. (<b>b</b>) Original ground truth data. (<b>c</b>) Revised ground truth data.</p>
Full article ">
20 pages, 25890 KiB  
Article
Charge-Coupled Frequency Response Multispectral Inversion Network-Based Detection Method of Oil Contamination on Airport Runway
by Shuanfeng Zhao, Zhijian Luo, Li Wang, Xiaoyu Li and Zhizhong Xing
Sensors 2024, 24(12), 3716; https://doi.org/10.3390/s24123716 - 7 Jun 2024
Viewed by 657
Abstract
Aircraft failures can result in the leakage of fuel, hydraulic oil, or other lubricants onto the runway during landing or taxiing. Damage to fuel tanks or oil lines during hard landings or accidents can also contribute to these spills. Further, improper maintenance or [...] Read more.
Aircraft failures can result in the leakage of fuel, hydraulic oil, or other lubricants onto the runway during landing or taxiing. Damage to fuel tanks or oil lines during hard landings or accidents can also contribute to these spills. Further, improper maintenance or operational errors may leave oil traces on the runway before take-off or after landing. Identifying oil spills in airport runway videos is crucial to flight safety and accident investigation. Advanced image processing techniques can overcome the limitations of conventional RGB-based detection, which struggles to differentiate between oil spills and sewage due to similar coloration; given that oil and sewage have distinct spectral absorption patterns, precise detection can be performed based on multispectral images. In this study, we developed a method for spectrally enhancing RGB images of oil spills on airport runways to generate HSI images, facilitating oil spill detection in conventional RGB imagery. To this end, we employed the MST++ spectral reconstruction network model to effectively reconstruct RGB images into multispectral images, yielding improved accuracy in oil detection compared with other models. Additionally, we utilized the Fast R-CNN oil spill detection model, resulting in a 5% increase in Intersection over Union (IOU) for HSI images. Moreover, compared with RGB images, this approach significantly enhanced detection accuracy and completeness by 25.3% and 26.5%, respectively. These findings clearly demonstrate the superior precision and accuracy of HSI images based on spectral reconstruction in oil spill detection compared with traditional RGB images. With the spectral reconstruction technique, we can effectively make use of the spectral information inherent in oil spills, thereby enhancing detection accuracy. Future research could delve deeper into optimization techniques and conduct extensive validation in real airport environments. In conclusion, this spectral reconstruction-based technique for detecting oil spills on airport runways offers a novel and efficient approach that upholds both efficacy and accuracy. Its wide-scale implementation in airport operations holds great potential for improving aviation safety and environmental protection. Full article
(This article belongs to the Section Environmental Sensing)
Show Figures

Figure 1

Figure 1
<p>Oil characteristics.</p>
Full article ">Figure 2
<p>Partial dataset.</p>
Full article ">Figure 3
<p>Spectral response curve acquisition platform. (<b>a</b>) Component display diagram of the platform. (<b>b</b>) Schematic diagram of experimental collection process.</p>
Full article ">Figure 4
<p>Spectral response curve acquisition. The various colour curves displayed in the graph represent the differing wavelength response values of the red, green and blue RGB channels, respectively.</p>
Full article ">Figure 5
<p>Schematic diagram of oil spectral reconstruction and oil detection network architecture.</p>
Full article ">Figure 6
<p>MST++ network structure diagram. (<b>a</b>) Multi-stage spectral-wise Transformer. (<b>b</b>) Single-stage spectral-wise Transformer. (<b>c</b>) Spectral-wise attention block. (<b>d</b>) Feed forward network.</p>
Full article ">Figure 7
<p>Structure of Faster RCNN network.</p>
Full article ">Figure 8
<p>Comparison of spectral recovery of 4 bands in RGB image.</p>
Full article ">Figure 9
<p>Reconstructed pixel results with coordinates p1 (323,250), p2 (439,169) and the RMSE and RMAE error histogram calculated from the model reconstruction results and the ground truth.</p>
Full article ">Figure 10
<p>The effect of oil reconstruction.</p>
Full article ">Figure 11
<p>Reconstruction effect of different oil stains in different bands.</p>
Full article ">Figure 12
<p>RGB images of oil in different scenes.</p>
Full article ">Figure 13
<p>Schematic illustration of the effect of RGB and HSI image detection of oil on airport pavement in multiple scenarios.</p>
Full article ">Figure 14
<p>Results of oil IOUs in multiple scenarios.</p>
Full article ">
18 pages, 4019 KiB  
Article
Assessment of C-Band Polarimetric Radar for the Detection of Diesel Fuel in Newly Formed Sea Ice
by Leah Hicks, Mahdi Zabihi Mayvan, Elvis Asihene, Durell S. Desmond, Katarzyna Polcwiartek, Gary A. Stern and Dustin Isleifson
Remote Sens. 2024, 16(11), 2002; https://doi.org/10.3390/rs16112002 - 2 Jun 2024
Viewed by 436
Abstract
There is a heightened risk of an oil spill occurring in the Arctic, as climate change driven sea ice loss permits an increase in Arctic marine transportation. The ability to detect an oil spill and monitor its progression is key to enacting an [...] Read more.
There is a heightened risk of an oil spill occurring in the Arctic, as climate change driven sea ice loss permits an increase in Arctic marine transportation. The ability to detect an oil spill and monitor its progression is key to enacting an effective response. Microwave scatterometer systems may be used detect changes in sea ice thermodynamic and physical properties, so we examined the potential of C-band polarimetric radar for detecting diesel fuel beneath a thin sea ice layer. Sea ice physical properties, including thickness, temperature, and salinity, were measured before and after diesel addition beneath the ice. Time-series polarimetric C-band scatterometer measurements monitored the sea ice evolution and diesel migration to the sea ice surface. We characterized the temporal evolution of the diesel-contaminated seawater and sea ice by monitoring the normalized radar cross section (NRCS) and polarimetric parameters (conformity coefficient (μ), copolarization correlation coefficient (ρco)) at 20° and 25° incidence angles. We delineated three stages, with distinct NRCS and polarimetric results, which could be connected to the thermophysical state and the presence of diesel on the surface. Stage 1 described the initial formation of sea ice, while in Stage 2, we injected 20L of diesel beneath the sea ice. No immediate response was noted in the radar measurements. With the emergence of diesel on the sea ice surface, denoted by Stage 3, the NRCS dropped substantially. The largest response was for VV and HH polarizations at 20° incidence angle. Physical sampling indicated that diesel emerged to the surface of the sea ice and trended towards the tub edge and the polarimetric scatterometer was sensitive to these physical changes. This study contributes to a greater understanding of how C-band frequencies can be used to monitor oil products in the Arctic and act as a baseline for the interpretation of satellite data. Additionally, these findings will assist in the development of standards for oil and diesel fuel detection in the Canadian Arctic in association with the Canadian Standards Association Group. Full article
(This article belongs to the Section Environmental Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>SERF Site Overview Prior to Experimental Setup.</p>
Full article ">Figure 2
<p>Experimental Setup at SERF and Radar Footprint of C-Scat (outlined in red). The highlighted areas represent the swaths of incidence angles used in this study after a careful inspection was done to eliminate errors associated with the edge of the tub. The solid blue line represents the incidence angle of 20° while the dashed yellow line is 25°.</p>
Full article ">Figure 3
<p>Physical Sampling Locations. S1: 7 March before oil injection, S2: 7 March after diesel injection, S3: 8 March, S4: 9 March.</p>
Full article ">Figure 4
<p>Temporal Progression of Ice Growth and Presence of Diesel Throughout the Study Period. (<b>A</b>) Start of Study, (<b>B</b>) Ice Formed, (<b>C</b>) After S1 and After Diesel Injection, (<b>D</b>) After S2, (<b>E</b>) After S3, (<b>F</b>) After S4 and End of Study.</p>
Full article ">Figure 5
<p>Temporal Variation of Air Temperature, Relative Humidity, and Wind Speed Throughout the Study Period. The Time of Diesel Injection is Marked by a Blue Line.</p>
Full article ">Figure 6
<p>C-band NRCS Values for the study period at the 20° and 25° incidence angles. The red shaded area is Stage 1, the non-shaded area is Stage 2, and the blue shaded area is Stage 3. The blue line denotes the time of diesel injection.</p>
Full article ">Figure 7
<p>Copolarization correlation coefficient (ρco) and conformity coefficient (μ) for the 20° (upper panel) and 25° (lower panel) incidence angles. The red-shaded area is stage one, the non-shaded area is stage two, and the blue-shaded area is stage three. A solid vertical blue line denotes the time of diesel injection.</p>
Full article ">
128 KiB  
Abstract
Marine Oil Spill Detection with Deep Neural Networks
by Fatih Uysal, Mesut Güven and Fırat Hardalaç
Proceedings 2024, 105(1), 4; https://doi.org/10.3390/proceedings2024105004 - 28 May 2024
Viewed by 218
Abstract
Oil spills, primarily due to accidents involving pipelines, tankers, and storage facilities, significantly impact marine life, particularly fish and shellfish [...] Full article
Back to TopTop