[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (542)

Search Parameters:
Keywords = InSAR monitoring

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
25 pages, 3121 KiB  
Article
Analysing the Value of Digital Twinning Opportunities in Infrastructure Asset Management
by João Vieira, Nuno Marques de Almeida, João Poças Martins, Hugo Patrício and João Gomes Morgado
Infrastructures 2024, 9(9), 158; https://doi.org/10.3390/infrastructures9090158 - 11 Sep 2024
Viewed by 229
Abstract
Many studies and technology companies highlight the actual or potential value of Digital Twins, but they often fail to demonstrate this value or how it can be realised. This gap constitutes a barrier for infrastructure asset management organisations in their attempt to innovate [...] Read more.
Many studies and technology companies highlight the actual or potential value of Digital Twins, but they often fail to demonstrate this value or how it can be realised. This gap constitutes a barrier for infrastructure asset management organisations in their attempt to innovate and incorporate digital twinning opportunities in their decision-making processes and their asset management planning activities. Asset management planning activities often make use of existing value-based decision-support tools to select and prioritise investments in physical assets. However, these tools were not originally designed to consider digital twinning investments that also compete for funding. This paper addresses this gap and proposes a value-based analysis for digital twinning opportunities in infrastructure asset management. The proposed analysis method is tested with three rail and road infrastructure case studies: (i) real-time monitoring of a power transformer; (ii) BIM for the design, construction, and maintenance of a new railway line; and (iii) infrastructure displacement monitoring using satellite data (InSAR). The study shows that the proposed method provides a conceptual construct and a common language that facilitates the communication of digital twinning opportunities in terms of their relevance in different contexts. The proposed method can be used to support the investment decision-making process for investments in both physical and non-physical assets and help derive maximum value from the limited available resources. Full article
(This article belongs to the Special Issue Recent Progress in Transportation Infrastructures)
Show Figures

Figure 1

Figure 1
<p>Types of decisions within the typical hierarchy of an infrastructure asset management organisation (adapted from [<a href="#B14-infrastructures-09-00158" class="html-bibr">14</a>,<a href="#B15-infrastructures-09-00158" class="html-bibr">15</a>]).</p>
Full article ">Figure 2
<p>LoDT radar for rail and road case studies.</p>
Full article ">Figure 3
<p>Steps of the value-based analysis of digital twinning opportunities.</p>
Full article ">Figure 4
<p>Proposed value framework for supporting decision-making in IP (abbreviations in parentheses).</p>
Full article ">
22 pages, 10522 KiB  
Article
Application of PS-InSAR and Diagnostic Train Measurement Techniques for Monitoring Subsidence in High-Speed Railway in Konya, Türkiye
by Gokhan Kizilirmak and Ziyadin Cakir
Infrastructures 2024, 9(9), 152; https://doi.org/10.3390/infrastructures9090152 - 7 Sep 2024
Viewed by 348
Abstract
Large-scale man-made linear structures like high-speed railway lines have become increasingly important in modern life as a faster and more comfortable transportation option. Subsidence or longitudinal levelling deformation problems along these railway lines can prevent the line from operating effectively and, in some [...] Read more.
Large-scale man-made linear structures like high-speed railway lines have become increasingly important in modern life as a faster and more comfortable transportation option. Subsidence or longitudinal levelling deformation problems along these railway lines can prevent the line from operating effectively and, in some cases, require speed reduction, continuous maintenance or repairs. In this study, the longitudinal levelling deformation of the high-speed railway line passing through Konya province (Central Turkey) was analyzed for the first time using the Persistent Scatter Synthetic Aperture Radar Interferometry (PS-InSAR) technique in conjunction with diagnostic train measurements, and the correlation values between them were found. In order to monitor potential levelling deformation along the railway line, medium-resolution, free-of-charge C-band Sentinel-1 (S-1) data and high-resolution, but paid, X-band Cosmo-SkyMed (CSK) Synthetic Aperture Radar (SAR) data were analyzed from the diagnostic train and reports received from the relevant maintenance department. Comparison analyses of the results obtained from the diagnostic train and radar measurements were carried out for three regions with different deformation scenarios, selected from a 30 km railway line within the whole analysis area. PS-InSAR measurements indicated subsidence events of up to 40 mm/year along the railway through the alluvial sediments of the Konya basin, which showed good agreement with the diagnostic train. This indicates that the levelling deformation of the railway and its surroundings can be monitored efficiently, rapidly and cost-effectively using the InSAR technique. Full article
Show Figures

Figure 1

Figure 1
<p>Photograph depicting a section of the Ankara–Konya High-Speed Railways provided by the Gokhan Kizilirmak.</p>
Full article ">Figure 2
<p>Geological maps: (<b>a</b>) shows the 1st study area; (<b>b</b>) shows the 2nd study area.</p>
Full article ">Figure 3
<p>Roger-800 performing measurements on the Ankara–Konya high-speed railway. The image was provided by Gokhan Kizilirmak.</p>
Full article ">Figure 4
<p>Photo showing the position of the laser measurement sensors. The image is provided by the Gokhan Kizilirmak.</p>
Full article ">Figure 5
<p>Illustration showing the working principle of levelling on a diagnostic train.</p>
Full article ">Figure 6
<p>Displacement diagram of the railway in the line of sight and at multiple passes.</p>
Full article ">Figure 7
<p>Simplified workflow of PS-InSAR processing in SARPROZ© (adapted from [<a href="#B58-infrastructures-09-00152" class="html-bibr">58</a>,<a href="#B59-infrastructures-09-00152" class="html-bibr">59</a>]).</p>
Full article ">Figure 8
<p>Image graphs for each time-series data stack: (<b>a</b>) CSK ascending; (<b>b</b>) S1–B/T65 descending; and (<b>c</b>) S1–B /T160 ascending. They show the 2D spatiotemporal baseline (yyyymmdd) spaces. Each point displays a scene, and each line displays an interferogram concerning a single master, which is represented with a red color dot.</p>
Full article ">Figure 9
<p>Reflectivity map showing the reference point location, city center and railway with blue colored text from all radar images.</p>
Full article ">Figure 10
<p>PSC maps and scatter plots: (<b>a</b>) CSK; (<b>b</b>) S1–B/T65; (<b>c</b>) S1–B/T160. PSC maps (red line means the railway) and mean velocity maps for CSK and S-1 analyses in LOS direction (dark blue line represents the 30 km-long railway).</p>
Full article ">Figure 11
<p>Vertical accumulated subsidence profiles of the railway along the 1st and 2nd study areas: (<b>a</b>) CSK; (<b>b</b>) S1–B/T65; (<b>c</b>) S1–B/T160; and (<b>d</b>) diagnostic train measurement time-series graphs.</p>
Full article ">Figure 12
<p>Accumulated subsidence graphs for the clustered PSs, where blue color represents PSs from CSK, orange color signs PSs from S1–B /T160, and lastly, grey color denotes PSs from S1–B /T65: (<b>a</b>) Location#1; (<b>b</b>) Location#2; (<b>c</b>) Location#3; and (<b>d</b>) Location#4.</p>
Full article ">Figure 13
<p>Specialized workflow model.</p>
Full article ">Figure A1
<p>Map of Ankara–Konya High-Speed railway showing the study areas.</p>
Full article ">
25 pages, 8750 KiB  
Article
Liaohe Oilfield Reservoir Parameters Inversion Based on Composite Dislocation Model Utilizing Two-Dimensional Time-Series InSAR Observations
by Hang Jiang, Rui Zhang, Bo Zhang, Kangyi Chen, Anmengyun Liu, Ting Wang, Bing Yu and Lin Deng
Remote Sens. 2024, 16(17), 3314; https://doi.org/10.3390/rs16173314 - 6 Sep 2024
Viewed by 427
Abstract
To address the industry’s demand for sustainable oilfield development and safe production, it is crucial to enhance the scientific rigor and accuracy of monitoring ground stability and reservoir parameter inversion. For the above purposes, this paper proposes a technical solution that employs two-dimensional [...] Read more.
To address the industry’s demand for sustainable oilfield development and safe production, it is crucial to enhance the scientific rigor and accuracy of monitoring ground stability and reservoir parameter inversion. For the above purposes, this paper proposes a technical solution that employs two-dimensional time-series ground deformation monitoring based on ascending and descending Interferometric Synthetic Aperture Radar (InSAR) technique first, and the composite dislocation model (CDM) is utilized to achieve high-precision reservoir parameter inversion. To validate the feasibility of this method, the Liaohe Oilfield is selected as a typical study area, and the Sentinel-1 ascending and descending Synthetic Aperture Radar (SAR) images obtained from January 2020 to December 2023 are utilized to acquire the ground deformation in various line of sight (LOS) directions based on Multitemporal Interferometric Synthetic Aperture Radar (MT-InSAR). Subsequently, by integrating the ascending and descending MT-InSAR observations, we solved for two-dimensional ground deformation, deriving a time series of vertical and east-west deformations. Furthermore, reservoir parameter inversion and modeling in the subsidence trough area were conducted using the CDM and nonlinear Bayesian inversion method. The experimental results indicate the presence of uneven subsidence troughs in the Shuguang and Huanxiling oilfields within the study area, with a continuous subsidence trend observed in recent years. Among them, the subsidence of the Shuguang oilfield is more significant and shows prominent characteristics of single-source center subsidence accompanied by centripetal horizontal displacement, the maximum vertical subsidence rate reaches 221 mm/yr, and the maximum eastward and westward deformation is more than 90 mm/yr. Supported by the two-dimensional deformation field, we conducted a comparative analysis between the Mogi, Ellipsoidal, and Okada models in terms of reservoir parameter inversion, model fitting efficacy, and residual distribution. The results confirmed that the CDM offers the best adaptability and highest accuracy in reservoir parameter inversion. The proposed technical methods and experimental results can provide valuable references for scientific planning and production safety assurance in related oilfields. Full article
(This article belongs to the Special Issue Synthetic Aperture Radar Interferometry Symposium 2024)
Show Figures

Figure 1

Figure 1
<p>The overall technical flow char.</p>
Full article ">Figure 2
<p>Diagram of vertical and horizontal deformation and slope of the subsidence trough.</p>
Full article ">Figure 3
<p>Simplified schematics of two<span class="html-italic">-</span>dimensional time series. The blue circles represent ascending and descending SAR data at time <span class="html-italic">t<sub>i</sub></span>. The horizontal solid line <span class="html-italic">I<sub>i</sub></span> between two points indicates the interferogram, while Δ<span class="html-italic">t<sub>i</sub></span> denotes the time interval between adjacent images.</p>
Full article ">Figure 4
<p>Diagram of the composite dislocation model (CDM).</p>
Full article ">Figure 5
<p>Study area and image coverage. The red and blue boxes indicate the coverage areas of the ascending and descending SAR data, respectively.</p>
Full article ">Figure 6
<p>The spatio-temporal baseline of the (<b>a</b>) ascending and (<b>b</b>) descending interferometric pairs.</p>
Full article ">Figure 7
<p>LOS deformation velocities for ascending (<b>a</b>) and descending (<b>b</b>) datasets. The red box highlights areas of significant subsidence, and the red star denotes the location of the chosen reference point.</p>
Full article ">Figure 8
<p>Profile deformation distribution. (<b>a</b>,<b>c</b>) represent the deformation velocity results for the Shuguang and Huanxiling oilfields from the ascending and descending datasets, including the selected profile lines and feature points. (<b>b</b>,<b>d</b>) represent the same for the deformation velocity results and profile lines for the ascending and descending datasets. (<b>e</b>–<b>h</b>) depict the deformation distribution along the profile lines AA’, BB’, CC’, and DD’ for the ascending and descending datasets.</p>
Full article ">Figure 9
<p>Time-series deformation for feature Points P1 (<b>a</b>) and P2 (<b>b</b>) in the ascending and descending datasets.</p>
Full article ">Figure 10
<p>2D deformation velocity maps. (<b>a</b>,<b>b</b>) represent the vertical and horizontal deformation velocities for the Shuguang oilfield area. (<b>c</b>,<b>d</b>) show the vertical and horizontal deformation velocities for the Huanxiling oilfield area. Positive and negative values for vertical deformation velocities indicate uplift and subsidence, respectively, while positive and negative values for horizontal deformation velocities represent eastward and westward deformation, respectively.</p>
Full article ">Figure 11
<p>Vertical and horizontal deformation distributions along the L1 profile line for the Shuguang oilfield (<b>a</b>) and the L2 profile line for the Huanxiling oilfield (<b>b</b>). (<b>c</b>,<b>d</b>) show the 3D effects of vertical deformation velocities for the Shuguang and Huanxiling oilfield areas, respectively.</p>
Full article ">Figure 12
<p>Vertical (<b>a</b>) and horizontal (<b>b</b>) time-series deformation characteristics from January 2021 to December 2021.</p>
Full article ">Figure 13
<p>3D results of reservoir parameter inversion for the Shuguang oilfield using the CDM.</p>
Full article ">Figure 14
<p>The observed deformation field in the (<b>a</b>) vertical and (<b>d</b>) horizontal directions. The modeled deformation field from the CDM parameter inversion in the (<b>b</b>) vertical and (<b>e</b>) horizontal directions. The residuals in the (<b>c</b>) vertical and (<b>f</b>) horizontal directions.</p>
Full article ">Figure 15
<p>The observed deformation field in the (<b>a</b>) vertical and (<b>d</b>) horizontal directions. The modeled deformation field from the Mogi parameter inversion in the (<b>b</b>) vertical and (<b>e</b>) horizontal directions. The residuals in the (<b>c</b>) vertical and (<b>f</b>) horizontal directions.</p>
Full article ">Figure 16
<p>The observed deformation field in the (<b>a</b>) vertical and (<b>d</b>) horizontal directions. The modeled deformation field from the Ellipsoidal parameter inversion in the (<b>b</b>) vertical and (<b>e</b>) horizontal directions; the residuals in the (<b>c</b>) vertical and (<b>f</b>) horizontal directions.</p>
Full article ">Figure 17
<p>The observed deformation field in the (<b>a</b>) vertical and (<b>d</b>) horizontal directions. The modeled deformation field from the Okada parameter inversion in the (<b>b</b>) vertical and (<b>e</b>) horizontal directions. The residuals in the (<b>c</b>) vertical and (<b>f</b>) horizontal directions.</p>
Full article ">Figure 18
<p>Vertical (<b>a</b>) and horizontal (<b>b</b>) observed deformation and the L1 profile line distribution. Fitting results of the four inversion models compared to the observed deformation along the vertical (<b>c</b>) and horizontal (<b>d</b>) profile lines.</p>
Full article ">Figure 19
<p>Histograms and statistics of residuals for the four models in the vertical (<b>a</b>) and horizontal (<b>b</b>) directions. The horizontal axis represents residual values in mm/yr, while the vertical axis indicates the number of residuals.</p>
Full article ">
19 pages, 16252 KiB  
Article
Method of Predicting Dynamic Deformation of Mining Areas Based on Synthetic Aperture Radar Interferometry (InSAR) Time Series Boltzmann Function
by Shenshen Chi, Xuexiang Yu and Lei Wang
Appl. Sci. 2024, 14(17), 7917; https://doi.org/10.3390/app14177917 - 5 Sep 2024
Viewed by 273
Abstract
The movement and deformation of rock strata and the ground surface is a dynamic deformation process that occurs as underground mining progresses. Therefore, the dynamic prediction of three-dimensional surface deformation caused by underground mining is of great significance for assessing potential geological disasters. [...] Read more.
The movement and deformation of rock strata and the ground surface is a dynamic deformation process that occurs as underground mining progresses. Therefore, the dynamic prediction of three-dimensional surface deformation caused by underground mining is of great significance for assessing potential geological disasters. Synthetic aperture radar interferometry (InSAR) has been introduced into the field of mine deformation monitoring as a new mapping technology, but it is affected by many factors, and it cannot monitor the surface deformation value over the entire mining period, making it impossible to accurately predict the spatiotemporal evolution characteristics of the surface. To overcome this limitation, we propose a new dynamic prediction method (InSAR-DIB) based on a combination of InSAR and an improved Boltzmann (IB) function model. Theoretically, the InSAR-DIB model can use information on small dynamic deformation during mining to obtain surface prediction parameters and further realize a dynamic prediction of the surface. The method was applied to the 1613 (1) working face in the Huainan mining area. The results showed that the estimated mean error of the predicted surface deformation during mining was between 80.2 and 112.5 mm, and the estimated accuracy met the requirements for mining subsidence monitoring. The relevant research results are of great significance, and they support expanding the application of InSAR in mining areas with large deformation gradients. Full article
Show Figures

Figure 1

Figure 1
<p>Principle diagram of arbitrary point movement deformation calculation.</p>
Full article ">Figure 2
<p>Technical principle diagram of the InSAR-DIB dynamic prediction model.</p>
Full article ">Figure 3
<p>Simulated LOS deformation map.</p>
Full article ">Figure 4
<p>Comparison of the measured and fitted LOS deformation values: (<b>a</b>) 12~36 days; (<b>b</b>) 36~60 days.</p>
Full article ">Figure 5
<p>A comparison of the predicted and measured values when the working face advanced to the 84th day.</p>
Full article ">Figure 6
<p>A comparison of the predicted and measured values when the working face advanced to the 108th day.</p>
Full article ">Figure 7
<p>1613 working face overview.</p>
Full article ">Figure 8
<p>Surface deformation map in the LOS direction from 4 November 2017 to 28 November 2017.</p>
Full article ">Figure 9
<p>Surface deformation map in the LOS direction from 10 December 2017 to 3 January 2018.</p>
Full article ">Figure 10
<p>Comparison of the measured and fitted values of LOS from 4 November 2017 to 28 November 2017.</p>
Full article ">Figure 11
<p>Comparison of the measured and fitted values of LOS from 10 December 2017 to 3 January 2018.</p>
Full article ">Figure 12
<p>Comparison of the measured subsidence values and InSAR-DIB-predicted subsidence values. From (<b>a</b>–<b>c</b>) are the monitoring points MS29–MS48; in (<b>d</b>) are the monitoring points MS01–MS62.</p>
Full article ">Figure 13
<p>A 1222 (1) comparison of the measured and fitted subsidence values of the working face.</p>
Full article ">Figure 14
<p>Variation law of the predicted parameters with mining degree (The pink triangle is represented as a parameter value).</p>
Full article ">Figure 15
<p>Comparison of the fitting subsidence factor and measured.</p>
Full article ">
23 pages, 19899 KiB  
Article
InSAR-Driven Dynamic Landslide Hazard Mapping in Highly Vegetated Area
by Liangxuan Yan, Qianjin Xiong, Deying Li, Enok Cheon, Xiangjie She and Shuo Yang
Remote Sens. 2024, 16(17), 3229; https://doi.org/10.3390/rs16173229 - 31 Aug 2024
Viewed by 543
Abstract
Landslide hazard mapping is important to urban construction and landslide risk management. Dynamic landslide hazard mapping considers landslide deformation with changes in the environment. It can show more details of the landslide process state. Landslides in highly vegetated areas are difficult to observe [...] Read more.
Landslide hazard mapping is important to urban construction and landslide risk management. Dynamic landslide hazard mapping considers landslide deformation with changes in the environment. It can show more details of the landslide process state. Landslides in highly vegetated areas are difficult to observe directly, which makes landslide hazard mapping much more challenging. The application of multi-InSAR opens new ideas for dynamic landslide hazard mapping. Specifically, landslide susceptibility mapping reflects the spatial probability of landslides. For rainfall-induced landslides, the scale exceedance probability reflects the temporal probability. Based on the coupling of them, dynamic landslide hazard mapping further considers the landslide deformation intensity at different times. Zigui, a highly vegetation-covered area, was taken as the study area. The landslide displacement monitoring effect of different band SAR datasets (ALOS-2, Sentinel-1A) and different interpretation methods (D-InSAR, PS-InSAR, SBAS-InSAR) were studied to explore a combined application method. The deformation interpreted by SBAS-InSAR was taken as the main part, PS-InSAR data were used in towns and villages, and D-InSAR was used for the rest. Based on the preliminary evaluation and the displacement interpreted by fusion InSAR, the dynamic landslide hazard mappings of the study area from 2019 to 2021 were finished. Compared with the preliminary evaluation, the dynamic mapping approach was more focused and accurate in predicting the deformation of landslides. The false positives in very-high-hazard zones were reduced by 97.8%, 60.4%, and 89.3%. Dynamic landslide hazard mapping can summarize the development of and change in landslides very well, especially in highly vegetated areas. Additionally, it can provide trend prediction for landslide early warning and provide a reference for landslide risk management. Full article
(This article belongs to the Special Issue Application of Remote Sensing Approaches in Geohazard Risk)
Show Figures

Figure 1

Figure 1
<p>Dynamic LHM workflow.</p>
Full article ">Figure 2
<p>(<b>a</b>) Zigui County in China; (<b>b</b>) study area in Zigui County; (<b>c</b>) landslide map of study area.</p>
Full article ">Figure 3
<p>Evaluation factors of LSM: (<b>a</b>) Elevation; (<b>b</b>) Slope; (<b>c</b>) Aspect; (<b>d</b>) TRI; (<b>e</b>) Curvature; (<b>f</b>) Plane Curvature; (<b>g</b>) Section Curvature; (<b>h</b>) Lithology; (<b>i</b>) Distance to Fault; (<b>j</b>) TWI; (<b>k</b>) SPI; (<b>l</b>) Distance to River.</p>
Full article ">Figure 4
<p>IGR of LSM factors.</p>
Full article ">Figure 5
<p>LSM of the study area.</p>
Full article ">Figure 6
<p>Preliminary LHM of the study area in different scenarios (<b>a</b>) Scenario A; (<b>b</b>) Scenario B; (<b>c</b>) Scenario C; (<b>d</b>) Scenario D.</p>
Full article ">Figure 7
<p>Visibility graph of the SAR datasets.</p>
Full article ">Figure 8
<p>Deformation maps of the study area from 2019 to 2021 by multi-InSAR: (<b>a</b>) 2019; (<b>b</b>) 2020; (<b>c</b>) 2021.</p>
Full article ">Figure 9
<p>Dynamic LHM in the study area from 2019 to 2021: (<b>a</b>) 2019; (<b>b</b>) 2020; (<b>c</b>) 2021.</p>
Full article ">Figure 10
<p>Photos of Xiaoyantou landslide: (<b>a</b>) landslide image on 12 November 2020, by GaoFeng-1; (<b>b</b>) landslide failure image on 27 November 2021, by GaoFeng-2; (<b>c</b>) landslide failure photo on August 20, 2021; (<b>d</b>) scarp on landslide crown; (<b>e</b>) cracks on landslide right edge.</p>
Full article ">Figure 11
<p>Time series InSAR spatiotemporal connection diagram: (<b>a</b>) PS-InSAR (34 scenes); (<b>b</b>) PS-InSAR (81 scenes); (<b>c</b>) SBAS-InSAR (32 scenes); (<b>d</b>) SBAS-InSAR (81 scenes).</p>
Full article ">Figure 12
<p>Comparison of Sentinel-1A time series InSAR interpretation results: (<b>a</b>) PS-InSAR (34 scenes); (<b>b</b>) PS-InSAR (81 scenes); (<b>c</b>) SBAS-InSAR (32 scenes); (<b>d</b>) SBAS-InSAR (81 scenes).</p>
Full article ">Figure 13
<p>Comparison of Sentinel-1A time series InSAR interpretation results of Lianhuatuo landslide: (<b>a</b>) PS-InSAR (34 scenes); (<b>b</b>) PS-InSAR (81 scenes); (<b>c</b>) SBAS-InSAR (32 scenes); (<b>d</b>) SBAS-InSAR (81 scenes).</p>
Full article ">Figure 14
<p>Deformation of ALOS-2 data by D-InSAR.</p>
Full article ">Figure 15
<p>Interpretation result comparison of Sentinel-1A and ALOS-2.</p>
Full article ">
18 pages, 32081 KiB  
Article
Monitoring and Law Analysis of Secondary Deformation on the Surface of Multi-Coal Seam Mining in Closed Mines
by Xiaofei Liu, Jiangtao Wang, Sen Du, Kazhong Deng, Guoliang Chen and Xipeng Qin
Remote Sens. 2024, 16(17), 3223; https://doi.org/10.3390/rs16173223 - 30 Aug 2024
Viewed by 435
Abstract
A large number of mines have been closed due to resource depletion, failure to meet safety production requirements, and other reasons. To effectively ensure the safety of the ecological environment above these closed mines along with the safety of engineering construction, it is [...] Read more.
A large number of mines have been closed due to resource depletion, failure to meet safety production requirements, and other reasons. To effectively ensure the safety of the ecological environment above these closed mines along with the safety of engineering construction, it is necessary to monitor the secondary deformation of closed mines. Based on TerraSAR-X, Sentinel-1A data, and InSAR technology, this study obtained high-density secondary surface deformation data on the Jiahe Coal Mine and Pangzhuang Coal Mine in the western Xuzhou area. Combining mining geological data, we analyzed the spatiotemporal variation patterns and mechanisms of secondary deformation in multi-seam mining of closed mines. It was found that when mining multiple seams involves large interlayer spacing, the secondary deformation pattern shows a “W” shape. In this situation, the deformation can be divided into five stages: subsidence, uplift, re-subsidence, re-uplift, and relative stability. This study provides technical support for the evaluation and prevention of secondary deformation hazards in closed mines. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of study area and mine closing time; (<b>b</b>) overlay map of the working face in Jiahe coal mine; and (<b>c</b>) overlay map of the working face in Pangzhuang coal mine.</p>
Full article ">Figure 2
<p>(<b>a</b>) Interferometric time-space baseline diagram of TerraSAR-X data; (<b>b</b>) interferometric time-space baseline diagram of Sentinel-1A data.</p>
Full article ">Figure 3
<p>Processing flow of DS-InSAR technology.</p>
Full article ">Figure 4
<p>(<b>a</b>) Profile of settlement rate based on TerraSAR-X data of the study area by SBAS-InSAR from 17 January 2014 to 8 January 2018; (<b>b</b>) Jiahe Coal Mine AA’ subsidence rate profile; and (<b>c</b>) Pangzhuang Coal Mine BB’ subsidence rate profile.</p>
Full article ">Figure 5
<p>Part of the time series cumulative deformation diagram based on TerraSAR-X data of the study area by SBAS-InSAR.</p>
Full article ">Figure 6
<p>Time series cumulative deformation diagram based on Sentinel-1A data of the study area by DS-InSAR.</p>
Full article ">Figure 7
<p>(<b>a</b>) Benchmark location map and location map of same-name points (yellow asterisks in the figure) of Terrasar-X data and Sentinel-1A data; (<b>b</b>) comparison of monitoring results and leveling data of SBAS-InSAR and DS-InSAR.</p>
Full article ">Figure 8
<p>Comparative analysis of monitoring results of Terrasar-X data and Sentinel-1A data for the same name points in overlapping time periods: (<b>a</b>) Point E of Jiahe Coal Mine and (<b>b</b>) Point F of Pangzhuang Coal Mine.</p>
Full article ">Figure 9
<p>Contrast curve of multi-seam selection in Jiahe Coal Mine at P1–P4 points.</p>
Full article ">Figure 10
<p>Contrast curve of multi-seam selection in Pangzhuang Coal Mine at points P5–P6.</p>
Full article ">
29 pages, 38452 KiB  
Article
Integration of Multi-Source Datasets for Assessing Ground Swelling/Shrinking Risk in Cyprus: The Case Studies of Pyrgos–Parekklisia and Moni
by Athanasios V. Argyriou, Maria Prodromou, Christos Theocharidis, Kyriaki Fotiou, Stavroula Alatza, Constantinos Loupasakis, Zampela Pittaki-Chrysodonta, Charalampos Kontoes, Diofantos G. Hadjimitsis and Marios Tzouvaras
Remote Sens. 2024, 16(17), 3185; https://doi.org/10.3390/rs16173185 - 28 Aug 2024
Viewed by 632
Abstract
The determination of swelling/shrinking phenomena, from natural and anthropogenic activity, is examined in this study through the synergy of various remote sensing methodologies. For the period of 2016–2022, a time-series InSAR analysis of Sentinel-1 satellite images, with a Coherent Change Detection procedure, was [...] Read more.
The determination of swelling/shrinking phenomena, from natural and anthropogenic activity, is examined in this study through the synergy of various remote sensing methodologies. For the period of 2016–2022, a time-series InSAR analysis of Sentinel-1 satellite images, with a Coherent Change Detection procedure, was conducted to calculate the Normalized Coherence Difference. These were combined with Sentinel-2 multispectral data by exploiting the Normalized Difference Vegetation Index to create multi-temporal image composites. In addition, ALOS-Palsar DEM derivatives highlighted the geomorphological characteristics, which, in conjunction with the satellite imagery outcomes and other auxiliary spatial datasets, were embedded within a Multi-Criteria Decision Analysis (MCDA) model. The synergy of the remote sensing and GIS techniques’ applicability within the MCDA model highlighted the zones undergoing seasonal swelling/shrinking processes in Pyrgos–Parekklisia and Moni regions in Cyprus. The accuracy assessment of the produced final MCDA outcome provided an overall accuracy of 72.4%, with the Kappa statistic being 0.66, indicating substantial agreement of the MCDA outcome with the results from a Persistent Scatterer Interferometry analysis and ground-truth observations. Thus, this study offers decision-makers a powerful procedure to monitor longer- and shorter-term swelling/shrinking phenomena. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The Pyrgos–Parekklisia, Moni, and Monagroulli deforming sites in Limassol, Cyprus.</p>
Full article ">Figure 2
<p>The Pyrgos Lemesou–Parekklisia and Moni–Monagroulli geology.</p>
Full article ">Figure 3
<p>Sentinel-1 satellite passes in ascending and descending tracks and satellite image details.</p>
Full article ">Figure 4
<p>The Coherent Change Detection workflow methodology. The step that provides the coherence values is marked in red.</p>
Full article ">Figure 5
<p>Pyrgos–Parekklisia area. (<b>a</b>) Coherence difference and (<b>b</b>) Normalized Coherence difference from descending Sentinel-1 satellite images during 12 February 2021–8 March 2021. (<b>c</b>) Coherence difference and (<b>d</b>) Normalized Coherence difference from ascending Sentinel-1 satellite images during 23 February 2021–7 March 2021.</p>
Full article ">Figure 6
<p>Moni–Monagroulli area. (<b>a</b>) Coherence difference and (<b>b</b>) Normalized Coherence difference from descending Sentinel-1 satellite images during 12 February 2021–8 March 2021. (<b>c</b>) Coherence difference and (<b>d</b>) Normalized Coherence difference from ascending Sentinel-1 satellite images during 23 February 2021–7 March 2021.</p>
Full article ">Figure 7
<p>Annual NDVI variations and corresponding masked areas excluded from further analysis.</p>
Full article ">Figure 8
<p>CCD of the Area of Interest showing the changes that occurred between 2016 and 2022.</p>
Full article ">Figure 9
<p>The TWI with dark bluish hues highlighting the high moisture accumulation.</p>
Full article ">Figure 10
<p>Landform type classification, showing valleys, semi-mountainous, and mountainous zones.</p>
Full article ">Figure 11
<p>The determined precipitation derived from the weather stations.</p>
Full article ">Figure 12
<p>The soil texture map of the AoI.</p>
Full article ">Figure 13
<p>The reclassified soil texture map, highlighting the degree of the swelling/shrinking effect.</p>
Full article ">Figure 14
<p>The reclassified hydrogeological map highlights the swelling degree.</p>
Full article ">Figure 15
<p>The GIS-based MCDA swelling and shrinking effect outcome based on the acknowledged variables of CCD, soil texture, hydrogeology, TWI, landforms, and rainfall. High-risk zones are presented in orange and very high-risk zones in red.</p>
Full article ">Figure 16
<p>(<b>a</b>) Sentinel-1 LoS displacements in Pyrgos–Parekklisia for descending satellite pass and (<b>b</b>) interpolated Sentinel-1 LOS displacements in Pyrgos–Parekklisia for descending satellite pass.</p>
Full article ">Figure 17
<p>The MCDA swelling and shrinking effect outcome with the overlaid ground-truth locations with verified deformed structures, indicated with red arrows, from ground-truth surveys.</p>
Full article ">Figure A1
<p>The distribution of the accuracy assessment points across the final MCDA swelling/shrinking effect outcome.</p>
Full article ">
23 pages, 9165 KiB  
Article
Leveraging Multi-Temporal InSAR Technique for Long-Term Structural Behaviour Monitoring of High-Speed Railway Bridges
by Winter Kim, Changgil Lee, Byung-Kyu Kim, Kihyun Kim and Ilwha Lee
Remote Sens. 2024, 16(17), 3153; https://doi.org/10.3390/rs16173153 - 26 Aug 2024
Viewed by 540
Abstract
The effective monitoring of railway facilities is crucial for safety and operational efficiency. This study proposes an enhanced remote monitoring technique for railway facilities, specifically bridges, using satellite radar InSAR (Interferometric Synthetic Aperture Radar) technology. Previous studies faced limitations such as insufficient data [...] Read more.
The effective monitoring of railway facilities is crucial for safety and operational efficiency. This study proposes an enhanced remote monitoring technique for railway facilities, specifically bridges, using satellite radar InSAR (Interferometric Synthetic Aperture Radar) technology. Previous studies faced limitations such as insufficient data points and challenges with topographical and structural variations. Our approach addresses these issues by analysing displacements from 30 images captured by the X-band SAR satellite, TerraSAR-X, over two years. We tested each InSAR parameter to develop an optimal set of parameters, applying the technique to a post-tensioned PSC (pre-stressed concrete) box bridge. Our findings revealed a recurring arch-shaped elevation along the bridge, attributed to temporal changes and long-term deformation. Further analysis showed a strong correlation between this deformation pattern and average surrounding temperature. This indicates that our technique can effectively identify micro-displacements due to temperature changes and structural deformation. Thus, the technique provides a theoretical foundation for improved SAR monitoring of large-scale social overhead capital (SOC) facilities, ensuring efficient maintenance and management. Full article
(This article belongs to the Special Issue Remote Sensing in Urban Infrastructure and Building Monitoring)
Show Figures

Figure 1

Figure 1
<p>Reinforced concrete railway track.</p>
Full article ">Figure 2
<p>Structural details of PSC box bridge: (<b>a</b>) cross-section of the bridge; (<b>b</b>) view of the bridge and surrounding environment.</p>
Full article ">Figure 3
<p>Mean annual air temperature of target region.</p>
Full article ">Figure 4
<p>Deflecting modes of bridge deck.</p>
Full article ">Figure 5
<p>Components forming the total displacement at time <math display="inline"><semantics> <mrow> <mi mathvariant="normal">t</mi> </mrow> </semantics></math>: (<b>a</b>) total sum of displacement versus time; (<b>b</b>) cyclic pattern of thermal expansion; (<b>c</b>) long-term deformation of reinforced concrete.</p>
Full article ">Figure 6
<p>Typical bridge cross-section and temperature gradient: (<b>a</b>) reference bridge cross-section; (<b>b</b>) actual temperature distribution; (<b>c</b>) simplified model.</p>
Full article ">Figure 7
<p>Typical bridge cross-section and temperature gradient: (<b>a</b>) a view from the top; (<b>b</b>) a horizontal view of the same model.</p>
Full article ">Figure 8
<p>Stages of PS-InSAR analysis.</p>
Full article ">Figure 9
<p>Connection graph of SLC images.</p>
Full article ">Figure 10
<p>PS-density-related parameters: (<b>a</b>) excess amount of PSs; (<b>b</b>) adequate amount of PSs; (<b>c</b>) insufficient amount of PSs.</p>
Full article ">Figure 11
<p>Comparison of parametric analysis results with survey data.</p>
Full article ">Figure 12
<p>PS-InSAR results of selected bridges at geocoded state: (<b>a</b>) Bridge A (<b>b</b>) Bridge B.</p>
Full article ">Figure 13
<p>PS-InSAR result of Bridge A.</p>
Full article ">Figure 14
<p>PS-InSAR result of Bridge B.</p>
Full article ">Figure 15
<p>PS-InSAR results of three consecutive spans from Bridge A: (<b>a</b>) span ‘a’, (<b>b</b>) span ‘b’, (<b>c</b>) span ‘c’.</p>
Full article ">Figure 16
<p>Time-series data of featured point clouds from the peak of each span: (<b>a</b>) span ‘a’, (<b>b</b>) span ‘b’, (<b>c</b>) span ‘c’.</p>
Full article ">Figure 17
<p>PS-InSAR results of three consecutive spans from Bridge B: (<b>a</b>) span ‘i’, (<b>b</b>) span ‘ii’, (<b>c</b>) span ‘iii’.</p>
Full article ">Figure 18
<p>Time-series data of featured point clouds from the peak of each span: (<b>a</b>) span ‘i’, (<b>b</b>) span ‘ii’, (<b>c</b>) span ‘iii’.</p>
Full article ">Figure 19
<p>Linear trend in concrete deformation based on time: (<b>a</b>) Bridge A; (<b>b</b>) Bridge B.</p>
Full article ">Figure 20
<p>Repeating trend in deformation data: (<b>a</b>) Bridge A; (<b>b</b>) Bridge B.</p>
Full article ">Figure 21
<p>Deformation of the bridge at corresponding temperature: (<b>a</b>) Bridge A; (<b>b</b>) Bridge B.</p>
Full article ">
17 pages, 13031 KiB  
Article
Accurate Deformation Retrieval of the 2023 Turkey–Syria Earthquakes Using Multi-Track InSAR Data and a Spatio-Temporal Correlation Analysis with the ICA Method
by Yuhao Liu, Songbo Wu, Bochen Zhang, Siting Xiong and Chisheng Wang
Remote Sens. 2024, 16(17), 3139; https://doi.org/10.3390/rs16173139 - 26 Aug 2024
Viewed by 638
Abstract
Multi-track synthetic aperture radar interferometry (InSAR) provides a good approach for the monitoring of long-term multi-dimensional earthquake deformation, including pre-, co-, and post-seismic data. However, the removal of atmospheric errors in both single- and multi-track InSAR data presents significant challenges. In this paper, [...] Read more.
Multi-track synthetic aperture radar interferometry (InSAR) provides a good approach for the monitoring of long-term multi-dimensional earthquake deformation, including pre-, co-, and post-seismic data. However, the removal of atmospheric errors in both single- and multi-track InSAR data presents significant challenges. In this paper, a method of spatio-temporal correlation analysis using independent component analysis (ICA) is proposed, which can extract multi-track deformation components for the accurate retrieval of earthquake deformation time series. Sentinel-1 data covering the double earthquakes in Turkey and Syria in 2023 are used to demonstrate the effectiveness of the proposed method. The results show that co-seismic displacement in the east–west and up–down directions ranged from −114.7 cm to 82.8 cm and from −87.0 cm to 63.9 cm, respectively. Additionally, the deformation rates during the monitoring period ranged from −137.9 cm/year to 123.3 cm/year in the east–west direction and from −51.8 cm/year to 45.7 cm/year in the up–down direction. A comparative validation experiment was conducted using three GPS stations. Compared with the results of the original MSBAS method, the proposed method provides results that are smoother and closer to those of the GPS data, and the average optimization efficiency is 43.08% higher. The experiments demonstrated that the proposed method could provide accurate two-dimensional deformation time series for studying the pre-, co-, and post-earthquake events of the 2023 Turkey–Syria Earthquakes. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart of the proposed atmospheric delay removal algorithm.</p>
Full article ">Figure 2
<p>Schematic diagram of the spatial and temporal ICA on InSAR time series.</p>
Full article ">Figure 3
<p>Locations of 2023 Turkey–Syria earthquake sequences. SAR dataset coverage is outlined in green rectangles. Locations of epicenters and GPS stations are marked with circles and triangles, respectively.</p>
Full article ">Figure 4
<p>Co-seismic deformation from (<b>a</b>) descending and (<b>b</b>) ascending SAR data. The epicenters of the Mw 7.8 and Mw 7.6 earthquakes are marked with red diamonds.</p>
Full article ">Figure 5
<p>Deformation velocity from (<b>a</b>) descending and (<b>b</b>) ascending SAR data.</p>
Full article ">Figure 6
<p>Independent component analysis results of ascending data. The left column contains the spatial patterns that illustrate the spatial distributions of signals, and the right column shows the corresponding temporal feature vectors, representing the temporal contribution of each source.</p>
Full article ">Figure 7
<p>The ICA results of descending data. The left column contains the spatial patterns that illustrate the spatial distributions of signals, and the right column shows the corresponding temporal feature vectors, representing the temporal contribution of each source.</p>
Full article ">Figure 8
<p>Spatio-temporal analysis matrix: (<b>a</b>) spatial correlation matrix; and (<b>b</b>) temporal correlation matrix.</p>
Full article ">Figure 9
<p>Two-dimensional co-seismic deformation maps of (<b>a</b>) original west–east, (<b>b</b>) original up–down, (<b>c</b>) ICA west–east, and (<b>d</b>) ICA up–down directions.</p>
Full article ">Figure 10
<p>Two-dimensional deformation velocity: (<b>a</b>) original west–east; (<b>b</b>) original up–down; (<b>c</b>) ICA west–east; and (<b>d</b>) ICA up–down.</p>
Full article ">Figure 11
<p>Two-dimensional deformation velocity during pre-seismic period in (<b>a</b>) east–west and (<b>b</b>) up–down directions, and during post-seismic period in (<b>c</b>) east–west and (<b>d</b>) up–down directions.</p>
Full article ">Figure 12
<p>Time series deformation in LOS direction from InSAR and GPS datasets.</p>
Full article ">Figure 13
<p>Time series deformation in horizontal and vertical directions from InSAR, traditional elevation-dependent atmospheric phase trend fitting method, proposed method, and GPS datasets.</p>
Full article ">
17 pages, 34782 KiB  
Article
Non-Tectonic Geohazards of Guangdong Province, China, Monitored Using Sentinel-1A/B from 2015 to 2022
by Jincang Liu, Zhenhua Fu, Lipeng Zhou, Guangcai Feng, Yilin Wang and Wulinhong Luo
Sensors 2024, 24(16), 5449; https://doi.org/10.3390/s24165449 - 22 Aug 2024
Viewed by 401
Abstract
Guangdong Province, home to 21 cities and a permanent population of 127.06 million people, boasts the largest provincial economy in China, contributing 11.76% to the national GDP in 2023. However, it is prone to geological hazards due to its geological conditions, extreme weather, [...] Read more.
Guangdong Province, home to 21 cities and a permanent population of 127.06 million people, boasts the largest provincial economy in China, contributing 11.76% to the national GDP in 2023. However, it is prone to geological hazards due to its geological conditions, extreme weather, and extensive human activities. Geohazards not only endanger lives but also hinder regional economic development. Monitoring surface deformation regularly can promptly detect geological hazards and allow for effective mitigation strategies. Traditional ground subsidence monitoring methods are insufficient for comprehensive surveys and rapid monitoring of geological hazards in the whole province. Interferometric Synthetic Aperture Radar (InSAR) technology using satellite images can achieve wide-area geohazard monitoring. However, current geological hazard monitoring in Guangdong Province based on InSAR technology lacks regional analysis and statistics of surface deformation across the entire province. Furthermore, such monitoring fails to analyze the spatial–temporal characteristics of surface deformation and disaster evolution mechanisms by considering the local geological features. To address these issues, current work utilizes Sentinel-1A/B satellite data covering Guangdong Province from 2015 to 2022 to obtain the wide-area surface deformation in the whole province using the multi-temporal (MT) InSAR technology. Based on the deformation results, a wide-area deformation region automatic identification method is used to identify the surface deformation regions and count the deformation area in each city of Guangdong Province. By analyzing the results, we obtained the following findings: (1) Using the automatic identification algorithm we identified 2394 deformation regions. (2) Surface subsidence is concentrated in the delta regions and reclamation areas; over a 4 cm/year subsidence rate is observed in the hilly regions of northern Guangdong, particularly in mining areas. (3) Surface deformation is closely related to geological structures and human activities. (4) Sentinel-1 satellite C-band imagery is highly effective for wide-area geological hazard monitoring, but has limitations in monitoring small-area geological hazards. In the future, combining the high-spatial–temporal-resolution L-band imagery from the NISAR satellite with Sentinel-1 imagery will allow for comprehensive monitoring and early warning of geological hazards, achieving multiple geometric and platform perspectives for geological hazard monitoring and management in Guangdong Province. The findings of this study have significant reference value for the monitoring and management of geological disasters in Guangdong Province. Full article
(This article belongs to the Section Environmental Sensing)
Show Figures

Figure 1

Figure 1
<p>The spatial distribution, level, and types of geological hazards in Guangdong Province. (<b>a</b>) Geological background and disaster distribution of the study area on a color-shaded elevation map; (<b>b</b>) the pie chart of geological hazard level; (<b>c</b>) the pie chart of geological hazard types.</p>
Full article ">Figure 2
<p>Coverage of the Sentinel-1 images over Guangdong Province from 2015 to 2022 including 11 frames in 5 tracks. The green line presents the footprint of Sentinel 1A/B data and the red line presents the boundary of Guangdong Province.</p>
Full article ">Figure 3
<p>The interferogram network figures are based on spatial–temporal baselines.</p>
Full article ">Figure 4
<p>Surface deformation velocity map of Guangdong Province for the period 2015–2022. A–F are six mining sites with large deformation; (<b>a</b>–<b>c</b>) are the zoom-ins of the Leizhou Peninsula, Pearl River Delta, and Hanjiang Delta.</p>
Full article ">Figure 5
<p>The uncertainty of the surface deformation velocity map of Guangdong Province for the period 2015–2022.</p>
Full article ">Figure 6
<p>Surface deformation velocity of the six selected mining areas in northern Guangdong Province. The figures (<b>a</b>–<b>f</b>) correspond to the six selected mining areas A–F in <a href="#sensors-24-05449-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 7
<p>Surface deformation sequence results of (<b>a</b>) the Zhuhai reclamation area and (<b>b</b>) the uplift region of Puning, Jieyang.</p>
Full article ">Figure 8
<p>Surface deformation rate map of coastal reclamation areas in Guangdong Province.</p>
Full article ">Figure 9
<p>Field survey photos of ground subsidence in Zhuhai. (<b>a</b>,<b>b</b>) A water pump house in Pingsha Town, Jinwan District, Zhuhai City; (<b>c</b>,<b>d</b>) an elliptical-shaped building in the Fourteenth Village of Tanzhou Town, Zhongshan City; (<b>e</b>,<b>f</b>) a stilted house in Ma’an Village, Nanlang Town, Zhongshan City.</p>
Full article ">Figure 10
<p>Deformation zones in Guangdong Province were identified by the automatic deformation detection method. The red points denote the location of deformation areas, and the blue lines are the administrative boundary.</p>
Full article ">Figure 11
<p>Identified deformation zones in (<b>a</b>) Zhanjiang City, (<b>b</b>) Zhuhai City, and (<b>c</b>) Jiangmen City.</p>
Full article ">
19 pages, 98931 KiB  
Article
Semi-Automatic Detection of Ground Displacement from Multi-Temporal Sentinel-1 Synthetic Aperture Radar Interferometry Analysis and Density-Based Spatial Clustering of Applications with Noise in Xining City, China
by Dianqiang Chen, Qichen Wu, Zhongjin Sun, Xuguo Shi, Shaocheng Zhang, Yi Zhang and Yunlong Wu
Remote Sens. 2024, 16(16), 3066; https://doi.org/10.3390/rs16163066 - 21 Aug 2024
Viewed by 563
Abstract
The China Loess Plateau (CLP) is the world’s most extensive and thickest region of loess deposits. The inherently loose structure of loess makes the CLP particularly vulnerable to geohazards such as landslides, collapses, and subsidence, resulting in substantial geological and environmental challenges. Xining [...] Read more.
The China Loess Plateau (CLP) is the world’s most extensive and thickest region of loess deposits. The inherently loose structure of loess makes the CLP particularly vulnerable to geohazards such as landslides, collapses, and subsidence, resulting in substantial geological and environmental challenges. Xining City, situated at the northwest edge of the CLP, is especially prone to frequent geological hazards due to intensified human activities and natural forces. Synthetic Aperture Radar Interferometry (InSAR) has become a widely used tool for identifying landslide hazards and displacement monitoring because of its high accuracy, low cost, and wide coverage. In this study, we utilized the small baseline subset (SBAS) InSAR technique to derive the line of sight (LOS) displacements of Xining City using Sentinel-1 datasets from ascending and descending orbits between October 2014 and September 2022. By integrating LOS displacements from the two datasets, we retrieved the eastward and vertical displacements to characterize the kinematics of active slopes. To identify the active areas semi-automatically, we applied the Density-Based Spatial Clustering of Applications with Noise (DBSCAN) algorithm to cluster InSAR measurement points (IMPs). Forty-eight active slopes with areas ranging from 0.0049 to 0.5496 km2 and twenty-five subsidence-dominant areas ranging from 0.023 to 3.123 km2 were identified across Xining City. Kinematics analysis of the Jiujiawan landslide indicated that acceleration started in August 2016, likely triggered by rainfall, and continued until the landslide. The extreme rainfall in August 2022 may have pushed the Jiujiawan landslide beyond its critical threshold, leading to instability. Additionally, the study identified nine active slopes that threaten the normal operation of the Lanzhou–Xinjiang High-Speed Railway, with kinematic analysis suggesting rainfall-related accelerations. The influence of anthropogenic activities on ground displacements in loess areas was also confirmed through time series displacement analysis. Our results can be leveraged for geohazard prevention and management in Xining City. As SAR image data continue to accumulate, InSAR can serve as a regular tool for maintaining up-to-date landslide inventories, thereby contributing to more sustainable geohazard management. Full article
(This article belongs to the Special Issue Advances in Remote Sensing for Land Subsidence Monitoring)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of our study area. (<b>b</b>) Geological map of Xining City [<a href="#B49-remotesensing-16-03066" class="html-bibr">49</a>].</p>
Full article ">Figure 2
<p>Sentinel-1 InSAR image pairs of (<b>a</b>) ascending and (<b>b</b>) descending orbit datasets.</p>
Full article ">Figure 3
<p>Workflow of semi-automatic detection of ground displacement in this study.</p>
Full article ">Figure 4
<p>Displacement rate maps in the LOS directions of the (<b>a</b>) ascending and (<b>b</b>) descending Sentinel-1 datasets and in the (<b>c</b>) eastward and (<b>d</b>) vertical directions.</p>
Full article ">Figure 5
<p>DBSCAN maps generated from the (<b>a</b>) ascending and (<b>b</b>) descending displacement rate maps, (<b>c</b>) identifying the cluster results by combining (<b>a</b>,<b>b</b>), and (<b>d</b>) the enlarged map.</p>
Full article ">Figure 6
<p>(<b>a</b>) Eastward and (<b>b</b>) vertical displacement rate maps of the Jiujiawan landslide.</p>
Full article ">Figure 7
<p>Time series eastward (E) and vertical (V) cumulative displacements of P1 and P2 in the Jiujiawan landslide.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>c</b>) Eastward and (<b>b</b>,<b>d</b>) vertical displacement rate maps of five typical landslides along the LXHR.</p>
Full article ">Figure 9
<p>Time series (<b>a</b>) eastward and (<b>b</b>) vertical cumulative displacements of P3-P6 in <a href="#remotesensing-16-03066-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 10
<p>Subsidence rate maps (<b>a1</b>–<b>f1</b>) and corresponding Google Earth optical images (<b>a2</b>–<b>f2</b>) of six typical subsidence zones.</p>
Full article ">Figure 11
<p>Cumulative vertical displacements of P7 and P8.</p>
Full article ">Figure 12
<p>Displacement rate maps (<b>a1</b>–<b>c1</b>) and corresponding Google Earth optical images (<b>a2</b>–<b>c2</b>) of typical active slopes affected by anthropogenic activities.</p>
Full article ">
22 pages, 11811 KiB  
Article
Research on the Application of Dynamic Process Correlation Based on Radar Data in Mine Slope Sliding Early Warning
by Yuejuan Chen, Yang Liu, Yaolong Qi, Pingping Huang, Weixian Tan, Bo Yin, Xiujuan Li, Xianglei Li and Dejun Zhao
Sensors 2024, 24(15), 4976; https://doi.org/10.3390/s24154976 - 31 Jul 2024
Viewed by 581
Abstract
With the gradual expansion of mining scale in open-pit coal mines, slope safety problems are increasingly diversified and complicated. In order to reduce the potential loss caused by slope sliding and reduce the major threat to the safety of life and property of [...] Read more.
With the gradual expansion of mining scale in open-pit coal mines, slope safety problems are increasingly diversified and complicated. In order to reduce the potential loss caused by slope sliding and reduce the major threat to the safety of life and property of residents in the mining area, this study selected two mining areas in Xinjiang as cases and focused on the relationship between phase noise and deformation. The study predicts the specific time point of slope sliding by analyzing the dynamic history correlation tangent angle between the two. Firstly, the time series data of the micro-variation monitoring radar are used to obtain the small deformation of the study area by differential InSAR (D-InSAR), and the phase noise is extracted from the radar echo in the sequence data. Then, the volume of the deformation body is calculated by analyzing the small deformation at each time point, and the standard deviation of the phase noise is calculated accordingly. Finally, the sliding time of the deformation body is predicted by combining the tangent angle of the ratio of the volume of the deformation body to the standard deviation of the phase noise. The results show that the maximum deformation rates of the deformation bodies in the studied mining areas reach 10.1 mm/h and 6.65 mm/h, respectively, and the maximum deformation volumes are 2,619,521.74 mm3 and 2,503,794.206 mm3, respectively. The predicted landslide time is earlier than the actual landslide time, which verifies the effectiveness of the proposed method. This prediction method can effectively identify the upcoming sliding events and the characteristics of the slope, provide more accurate and reliable prediction results for the slope monitoring staff, and significantly improve the efficiency of slope monitoring and early warning. Full article
(This article belongs to the Section Remote Sensors)
Show Figures

Figure 1

Figure 1
<p>Geographic location and site image of Research Area 1: (<b>a</b>) Altay prefecture in Xinjiang Province, (<b>b</b>) Fuyun County in Altay Region, (<b>c</b>) Post-slope slide site image.</p>
Full article ">Figure 2
<p>Geographic location of study area two: (<b>a</b>) Aksu prefecture in Xinjiang Province, (<b>b</b>) Baycheng county in Aksu prefecture.</p>
Full article ">Figure 3
<p>Micro-variation monitoring radar measurement diagram.</p>
Full article ">Figure 4
<p>MPDMR-LSA radar system.</p>
Full article ">Figure 5
<p>Radar slices of two research areas: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 6
<p>Saito curve model and tangent angle model.</p>
Full article ">Figure 7
<p>Three types of displacement–time curves of slope deformation.</p>
Full article ">Figure 8
<p>Slope slide early warning flowchart.</p>
Full article ">Figure 9
<p>Working principle of deformation monitoring with micro-variation monitoring radar.</p>
Full article ">Figure 10
<p>Interferogram of the study area: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 11
<p>Cumulative displacement deformation: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 12
<p>Coherence map of the study area: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 13
<p>The deformed pixels are finally selected: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 14
<p>Cumulative displacement curve: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 15
<p>Cumulative volume change: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 16
<p>Volume change during the acceleration deformation stage: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 17
<p>Phase noise standard deviation curve during the deformation stage: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 18
<p>Dynamic process correlation curve of the deformation area: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 19
<p>Tangent angle of displacement in the deformation area: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">Figure 20
<p>Tangent angle of dynamic process correlation in the deformation area: (<b>a</b>) Study area one, (<b>b</b>) Study area two.</p>
Full article ">
23 pages, 30405 KiB  
Article
North American Circum-Arctic Permafrost Degradation Observation Using Sentinel-1 InSAR Data
by Shaoyang Guan, Chao Wang, Yixian Tang, Lichuan Zou, Peichen Yu, Tianyang Li and Hong Zhang
Remote Sens. 2024, 16(15), 2809; https://doi.org/10.3390/rs16152809 - 31 Jul 2024
Viewed by 529
Abstract
In the context of global warming, the accelerated degradation of circum-Arctic permafrost is releasing a significant amount of carbon. InSAR can indirectly reflect the degradation of permafrost by monitoring its deformation. This study selected three typical permafrost regions in North America: Alaskan North [...] Read more.
In the context of global warming, the accelerated degradation of circum-Arctic permafrost is releasing a significant amount of carbon. InSAR can indirectly reflect the degradation of permafrost by monitoring its deformation. This study selected three typical permafrost regions in North America: Alaskan North Slope, Northern Great Bear Lake, and Southern Angikuni Lake. These regions encompass a range of permafrost landscapes, from tundra to needleleaf forests and lichen-moss, and we used Sentinel-1 SAR data from 2018 to 2021 to determine their deformation. In the InSAR process, due to the prolonged snow cover in the circum-Arctic permafrost, we used only SAR data collected during the summer and applied a two-stage interferogram selection strategy to mitigate the resulting temporal decorrelation. The Alaskan North Slope showed pronounced subsidence along the coastal alluvial plains and uplift in areas with drained thermokarst lake basins. Northern Great Bear Lake, which was impacted by wildfires, exhibited accelerated subsidence rates, revealing the profound and lasting impact of wildfires on permafrost degradation. Southern Angikuni Lake’s lichen and moss terrains displayed mild subsidence. Our InSAR results indicate that more than one-third of the permafrost in the North American study area is degrading and that permafrost in diverse landscapes has different deformation patterns. When monitoring the degradation of large-scale permafrost, it is crucial to consider the unique characteristics of each landscape. Full article
(This article belongs to the Special Issue Remote Sensing of the Cryosphere II)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The coverage of Sentinel-1 images in the study areas, where the background permafrost zonation is from [<a href="#B29-remotesensing-16-02809" class="html-bibr">29</a>]. (<b>b</b>) The land cover in the study areas, where detailed labeling information can be found in [<a href="#B30-remotesensing-16-02809" class="html-bibr">30</a>] for Alaska and [<a href="#B31-remotesensing-16-02809" class="html-bibr">31</a>] for Canada.</p>
Full article ">Figure 2
<p>Flowchart of InSAR processing.</p>
Full article ">Figure 3
<p>Examples of interferograms and interferometric coherence maps in end-to-begin and end-to-end. (<b>a</b>) Mean interferometric coherence. (<b>b</b>,<b>e</b>) Interferogram and coherence map in end-to-begin. (<b>c</b>,<b>f</b>) Interferogram and coherence map in end-to-end. (<b>d</b>,<b>g</b>) Interferogram and coherence map within a single thawing season.</p>
Full article ">Figure 4
<p>Coherence over the consecutive four thawing seasons. The blue lines represent the coherence calculated from the interferogram combinations of SAR images acquired at time t<sub>i</sub> and t<sub>i+1</sub> in chronological order.</p>
Full article ">Figure 5
<p>Average interferometric coherence of different landscape features.</p>
Full article ">Figure 6
<p>Mean annual deformation rates of the North Slope of Alaska from mid-June 2018 to the end of September 2021. Sector 1 shows an enlarged view of Utqiagvik (U1 and U2 representing specific active layer monitoring sites); Sector 2 shows an enlarged view of Prudhoe Bay; Sector 3 shows an enlarged view along the alignment of the Alaska Pipeline. The blue and red boxes are the uplift area in <a href="#remotesensing-16-02809-f007" class="html-fig">Figure 7</a> and the subsidence area in <a href="#remotesensing-16-02809-f008" class="html-fig">Figure 8</a>, respectively.</p>
Full article ">Figure 7
<p>Deformation results and displacement curves in a typical uplift area. (<b>a</b>) Mean annual deformation rates. (<b>b</b>) Enlarged deformation image of the typical uplift area. (<b>c</b>,<b>d</b>) Optical images of the typical uplift area in 1986 and 2020 (sourced from Google Earth). (<b>e</b>) Cumulative displacement curve at point P (154.155°W, 70.092°N), air and soil temperature data sourced from the ECMWF Reanalysis v5 (ERA5-Land) dataset.</p>
Full article ">Figure 8
<p>Deformation results and displacement curve in a typical subsidence area. (<b>a</b>) Mean annual deformation rates. (<b>b</b>,<b>c</b>) Enlarged deformation image and optical image of the typical subsidence area (optical image from Esri). (<b>d</b>) Cumulative displacement curve at point P (159.160°W, 70.668°N), air and soil temperature data sourced from the ERA5-Land dataset.</p>
Full article ">Figure 9
<p>Mean annual deformation rates of the Northern Great Bear Lake and an enlarged view of a wildfire area. (<b>a</b>) Mean annual deformation rates from 2018-06-10 to 2021-10-04. (<b>b</b>) Enlarged view of two wildfire areas. (<b>c</b>) Optical image of wildfire areas (sourced from Google Earth, 2020). In (<b>b</b>), the red box A represents the area where a wildfire occurred in 2012, the yellow box B represents the area where a wildfire occurred in 2017, and the blue box C represents the unburned area.</p>
Full article ">Figure 10
<p>The distribution of the LOS deformation rate in two separate wildfires and the surrounding area. In each box-and-whisker plot, the box boundaries are the 25th and 75th percentiles, the line inside the box is the mean, and the whiskers are the 5th and 95th percentiles.</p>
Full article ">Figure 11
<p>(<b>a</b>) Mean annual deformation rates of the Southern Angikuni Lake. (<b>b</b>) Enlarged view of the optical image of Hex Lake (sourced from Google Earth). (<b>c</b>) Detailed land cover types of the Southern Angikuni Lake.</p>
Full article ">Figure 12
<p>Density distribution of the surface deformation rates in three typical permafrost regions.</p>
Full article ">Figure 13
<p>Results of the reliability validation of permafrost deformation in the North Slope of Alaska. (<b>a</b>) The coverage and overlap area of the Sentinel-1 images. (<b>b</b>) The correlation result between orbit 44 and orbit 73 (Sector 1) based on the deformation results. (<b>c</b>) The correlation result between orbit 73 and orbit 102 (Sector 2) based on the deformation results.</p>
Full article ">Figure 14
<p>Relationship between the InSAR displacement and the ALT. For visualization, convert the negative subsidence values into positive values to represent the average displacement. Vertical bars represent the range of ALT values for each year.</p>
Full article ">
27 pages, 8943 KiB  
Article
How Phenology Shapes Crop-Specific Sentinel-1 PolSAR Features and InSAR Coherence across Multiple Years and Orbits
by Johannes Löw, Steven Hill, Insa Otte, Michael Thiel, Tobias Ullmann and Christopher Conrad
Remote Sens. 2024, 16(15), 2791; https://doi.org/10.3390/rs16152791 - 30 Jul 2024
Viewed by 575
Abstract
Spatial information about plant health and productivity are essential when assessing the progress towards Sustainable Development Goals such as life on land and zero hunger. Plant health and productivity are strongly linked to a plant’s phenological progress. Remote sensing, and since the launch [...] Read more.
Spatial information about plant health and productivity are essential when assessing the progress towards Sustainable Development Goals such as life on land and zero hunger. Plant health and productivity are strongly linked to a plant’s phenological progress. Remote sensing, and since the launch of Sentinel-1 (S1), specifically, radar-based frameworks have been studied for the purpose of monitoring phenological development. This study produces insights into how crop phenology shapes S1 signatures of PolSAR features and InSAR coherence of wheat, canola, sugar beet. and potato across multiple years and orbits. Hereby, differently smoothed time series and a base line of growing degree days are stacked to estimate the patterns of occurrence of extreme values and break points. These patterns are then linked to in situ observations of phenological developments. The comparison of patterns across multiple orbits and years reveals that a single optimized fit hampers the tracking capacities of an entire season monitoring framework, as does the sole reliance on extreme values. VV and VH backscatter intensities outperform all other features, but certain combinations of phenological stage and crop type are better covered by a complementary set of PolSAR features and coherence. With regard to PolSAR features, alpha and entropy can be replaced by the cross-polarization ratio for tracking certain stages. Moreover, a range of moderate incidence angles is better suited for monitoring crop phenology. Also, wheat and canola are favored by a late afternoon overpass. In sum, this study provides insights into phenological developments at the landscape level that can be of further use when investigating spatial and temporal variations within the landscape. Full article
(This article belongs to the Special Issue Cropland Phenology Monitoring Based on Cloud-Computing Platforms)
Show Figures

Figure 1

Figure 1
<p>Map of InVeKoS data 2020 for DEMMIN and the selected crops: winter wheat, sugar beet, canola, and potato. Top right corner: extent of the AOI in Mecklenburg Western Pomerania. Center right: extent in relation to footprint of relative orbits.</p>
Full article ">Figure 2
<p>Essential steps of the analysis per orbit and year separated by field and landscape level.</p>
Full article ">Figure 3
<p>Schematic depiction of estimating temporal density (TSM occurrence plot) of TSM occurrence at the field scale. The dimensions of content of the analysis encompass five years (2017–2021), three relative orbits (146,168, 95), and seven S1 features. The smoothing span ranges from 0.05 to 0.5 in steps of 0.05, resulting in eleven (n = 11) time series per field.</p>
Full article ">Figure 4
<p>Exemplary yearly crop signatures for each crop type of VV backscatter with locations of their extrema. Signatures were smoothed by LOESS with span 0.2.</p>
Full article ">Figure 5
<p>Schematic depiction of the analyses at the landscape level containing the pattern extraction and the derivation of trackable stages. This was applied for time series originating from different years and/or orbits of the same crop type and S1 feature to enable the comparison of their respective TSM distributions. These comparisons allow for the derivation of common phenological patterns across years and orbits for each crop type and S1 feature.</p>
Full article ">Figure 6
<p>Orbit-specific patterns of major signal changes at landscape level tracked by break points according to day of year (DOY; <span class="html-italic">x</span>-axis) and artificial growing degree day (GDDsim) values (<span class="html-italic">y</span>-axis) in relation to the corresponding five-year mean GDDsim value of BBCH stadia observed by DWD at landscape level from 2018 and 2020. Exemplary illustration for fields of wheat. Temporal uncertainties around BBCH stadia are marked by grey areas. Exemplary illustration for fields of winter wheat.</p>
Full article ">Figure 7
<p>Year-wise count of S1 features producing break points (Y.) that closely track phenological stages by crop type and by their respective distribution of GDD values (GD.) at the landscape level which is overlaid by the GDD values of BBCH in situ observations (colored areas).</p>
Full article ">Figure 8
<p>Orbit, stage, and crop-specific offsets of break points at landscape level in days, displaying their mean deviation from in situ observations and temporal variance (standard deviation) by crop type and BBCH stage, containing only tracked events that were labeled reliable by the threshold approach.</p>
Full article ">Figure 9
<p>Orbit, stage, and crop-specific offsets of maxima at landscape level in days, displaying their mean deviation from in situ observations and temporal variance (standard deviation) by crop type and BBCH stage, containing only tracked events that were labeled reliable by the threshold approach.</p>
Full article ">Figure 10
<p>Orbit, stage, and crop-specific offsets of minima at landscape level in days, displaying their mean deviation from in situ observations and temporal variance (standard deviation) by crop type and BBCH stage, containing only tracked events that were labeled reliable by the threshold approach.</p>
Full article ">Figure A1
<p>Orbit-specific patterns of major signal changes at landscape level tracked by break points according to day of year (DOY; <span class="html-italic">x</span>-axis) and artificial growing degree day (GDDsim) values (<span class="html-italic">y</span>-axis) in relation to the corresponding five-year mean GDD value of BBCH stadia observed by DWD at landscape level from 2017 to 2021. Exemplary illustration for fields of winter wheat.</p>
Full article ">Figure A2
<p>Orbit-specific patterns of major signal changes by maxima by day of year (DOY; <span class="html-italic">x</span>-axis) and growing degree day (GDDsim) values (<span class="html-italic">y</span>-axis) in relation to the corresponding five-year mean GDDsim value of BBCH stadia observed by DWD at landscape level from 2017 to 2021. Exemplary illustration for fields of winter wheat.</p>
Full article ">Figure A3
<p>Year-wise count of S1 features producing maxima (Y.) that closely track phenological stages by crop type and by their respective distribution of GDD values (GD.)at the landscape level which is overlaid by the GDD values of BBCH in situ observations (colored areas).</p>
Full article ">Figure A4
<p>Orbit-specific patterns of major signal changes by minima by day of year (DOY; <span class="html-italic">x</span>-axis) and growing degree day (GDDsim) values (<span class="html-italic">y</span>-axis) in relation to the corresponding five-year mean GDDsim value of BBCH stadia observed by DWD at landscape level from 2017 to 2021. Exemplary illustration for fields of winter wheat.</p>
Full article ">Figure A5
<p>Year-wise count of S1 features producing minima (Y.) that closely track phenological stages by crop type and by their respective distribution of GDD values (GD.)at the landscape level which is overlaid by the GDD values of BBCH in situ observations (colored areas).</p>
Full article ">
23 pages, 14119 KiB  
Article
Construction of High-Precision and Complete Images of a Subsidence Basin in Sand Dune Mining Areas by InSAR-UAV-LiDAR Heterogeneous Data Integration
by Rui Wang, Shiqiao Huang, Yibo He, Kan Wu, Yuanyuan Gu, Qimin He, Huineng Yan and Jing Yang
Remote Sens. 2024, 16(15), 2752; https://doi.org/10.3390/rs16152752 - 27 Jul 2024
Viewed by 558
Abstract
Affected by geological factors, the scale of surface deformation in a hilly semi-desertification mining area varies. Meanwhile, there is certain dense vegetation on the ground, so it is difficult to construct a high-precision and complete image of a subsidence basin by using a [...] Read more.
Affected by geological factors, the scale of surface deformation in a hilly semi-desertification mining area varies. Meanwhile, there is certain dense vegetation on the ground, so it is difficult to construct a high-precision and complete image of a subsidence basin by using a single monitoring method, and hence the laws of the deformation and inversion of mining parameters cannot be known. Therefore, we firstly propose conducting collaborative monitoring by using InSAR (Interferometric Synthetic Aperture Radar), UAV (unmanned aerial vehicle), and 3DTLS (three-dimensional terrestrial laser scanning). The time-series complete surface subsidence basin is constructed by fusing heterogeneous data. In this paper, SBAS-InSAR (Small Baseline Subset) technology, which has the characteristics of reducing the time and space discorrelation, is used to obtain the small-scale deformation of the subsidence basin, oblique photogrammetry and 3D-TLS with strong penetrating power are used to obtain the anomaly and large-scale deformation, and the local polynomial interpolation based on the weight of heterogeneous data is used to construct a complete and high-precision subsidence basin. Compared with GNSS (Global Navigation Satellite System) monitoring data, the mean square errors of 1.442 m, 0.090 m, 0.072 m are obtained. The root mean square error of the high-precision image of the subsidence basin data is 0.040 m, accounting for 1.4% of the maximum subsidence value. The high-precision image of complete subsidence basin data can provide reliable support for the study of surface subsidence law and mining parameter inversion. Full article
(This article belongs to the Special Issue Synthetic Aperture Radar Interferometry Symposium 2024)
Show Figures

Figure 1

Figure 1
<p>An overview of the study area. (<b>a</b>) Specific coordinates of 2S201 working face. (<b>b</b>) The red box represents the range of the working face, and the black line represents the contours of the coal seam.</p>
Full article ">Figure 2
<p>Aerial mapping of 2S201 working face in Wangjiata Coal Mine (Inner Mongolia, China). (<b>a</b>) A photo of the UAV. (<b>b</b>) Flight range. (<b>c</b>) Track and image control point layout.</p>
Full article ">Figure 3
<p>Three-dimensional laser scanning applied to surface subsidence monitoring diagram.</p>
Full article ">Figure 4
<p>Layout of 3D laser scanning station: (<b>a</b>) 2S201 working face topographic map; (<b>b</b>) S* represents each station.</p>
Full article ">Figure 5
<p>Imaging equation coordinate system.</p>
Full article ">Figure 6
<p>Scanning sketch.</p>
Full article ">Figure 7
<p>Scanning coordinate system.</p>
Full article ">Figure 8
<p>Data fusion flowchart.</p>
Full article ">Figure 9
<p>SBASInSAR cumulative subsidence: (<b>a</b>) The surface deformation map from the beginning to the end of mining. (<b>b</b>) The surface deformation map from the beginning to the end of mining until stable subsidence.</p>
Full article ">Figure 10
<p>UAV timing DEM. (<b>a</b>) DEM 20180609; (<b>b</b>) DEM 20180904; (<b>c</b>) DEM 20181015; (<b>d</b>) DEM 20190416.</p>
Full article ">Figure 11
<p>UAV time-series subsidence basin: (<b>a</b>) 20180609–201800904 cumulative deformation data; (<b>b</b>) 20180609–20181015 cumulative deformation data; (<b>c</b>) 20180609–20190416 cumulative deformation data.</p>
Full article ">Figure 12
<p>Schematic diagram of progressive triangulation encryption algorithm.</p>
Full article ">Figure 13
<p>Adaptive grid diagram.</p>
Full article ">Figure 14
<p>The red circle area shows the result of different filtering algorithms. (<b>a</b>) the original surface point cloud data; (<b>b</b>) the result of the Morphological Filtering Method; (<b>c</b>) the result of CSF; (<b>d</b>) the result of the adaptive grid progressive TIN densification filtering encryption algorithm.</p>
Full article ">Figure 15
<p>Extraction of time sequence subsidence basins by three-dimensional laser scanning technique: (<b>a</b>) 20180609–20180730 cumulative deformation data; (<b>b</b>) 20180609–20180903 cumulative deformation data; (<b>c</b>) 20180609–20181016 cumulative deformation data; (<b>d</b>) 20180609–20190416 cumulative deformation data.</p>
Full article ">Figure 16
<p>The 2S201 working face sky and earth data fusion subsidence basin fused by the space–sky–surface integrated monitoring data.</p>
Full article ">Figure 17
<p>Four different methods of settlement map: (<b>a</b>) cumulative subsidence by InSAR; (<b>b</b>) cumulative subsidence by UAV; (<b>c</b>) cumulative subsidence by LiDAR; (<b>d</b>) cumulative subsidence by fusion method.</p>
Full article ">Figure 18
<p>Error analysis diagram of four different monitoring methods. (<b>a</b>) The comparison between the results obtained by different monitoring methods and GNSS monitoring results individually. (<b>b</b>) The error comparison between the results obtained by different monitoring methods and GNSS monitoring combined.</p>
Full article ">Figure 19
<p>Three-dimensional space fused by the space-sky-surface integrated monitoring data and the profile curve. (<b>a</b>) Three-dimensional subsidence map of 2S201 working face; (<b>b</b>) point d-e profile curve of 2S201 working face.</p>
Full article ">
Back to TopTop