Next Article in Journal
Influence of Image Compositing and Multisource Data Fusion on Multitemporal Land Cover Mapping of Two Philippine Watersheds
Next Article in Special Issue
Surface Subsidence Characteristics and Causes Analysis in Ningbo Plain by Sentinel-1A TS-InSAR
Previous Article in Journal
Cross-Border Sand and Dust Storms between Mongolia and Northern China in Spring and Their Driving Weather Systems
Previous Article in Special Issue
UAV, GNSS, and InSAR Data Analyses for Landslide Monitoring in a Mountainous Village in Western Greece
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Time-Series Analysis of Mining-Induced Subsidence in the Arid Region of Mongolia Based on SBAS-InSAR

1
School of Environmental and Geographical Sciences, Shanghai Normal University, Shanghai 200234, China
2
Institute of Geography and Geoecology, Mongolian Academy of Sciences, Ulaanbaatar 13330, Mongolia
3
Regional Environment Conservation Division, National Institute for Environmental Studies, Tsukuba 305-8506, Japan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Remote Sens. 2024, 16(12), 2166; https://doi.org/10.3390/rs16122166
Submission received: 30 April 2024 / Revised: 3 June 2024 / Accepted: 12 June 2024 / Published: 14 June 2024
(This article belongs to the Special Issue Remote Sensing in Urban Natural Hazards Monitoring)

Abstract

:
Mongolia’s substantial mineral resources have played a pivotal role in its economic progress, with mining activities significantly contributing to this development. However, these continuous mining operations, particularly at the Oyu Tolgoi copper and gold mine, have induced land subsidence that threatens both production activities and poses risks of geological and other natural disasters. This study employs the Small Baseline Subset Interferometric Synthetic Aperture Radar (SBAS-InSAR) technique to monitor and analyze time-series surface subsidence using 120 Sentinel-1A datasets from 2018 to 2022. The findings reveal that the SBAS-InSAR method successfully captures the subsidence and its spatial distribution at Oyu Tolgoi, with the maximum cumulative subsidence reaching −742.01 mm and the highest annual average subsidence rate at −158.11 mm/year. Key drivers identified for the subsidence include variations in groundwater levels, active mining operations, and changes in surface stress. This research underscores the ongoing subsidence issue at the Oyu Tolgoi mining area, providing crucial insights that could aid in enhancing mining safety and environmental conservation in the region.

Graphical Abstract

1. Introduction

The Mongolian economy is highly reliant on mineral extraction, focusing particularly on gold, copper, and coal, with over 250 active mining operations reported as of 2022 [1,2]. While the mining sector has propelled economic growth through job creation, foreign investment, and increased household incomes [3], it has also led to severe ecological issues. These include environmental degradation, water and air pollution, soil erosion, desertification, and, notably, surface subsidence, which are collectively contributing to significant geo-hazards [4,5,6,7,8,9,10,11].
Remote sensing, including traditional and advanced Interferometric Synthetic Aperture Radar (InSAR) technologies, has been pivotal in monitoring these impacts. Traditional methods such as GPS and leveling, despite their precision, are limited by high costs and labor intensity and lack the capability for extensive area coverage [12]. In contrast, InSAR technologies, including Differential Interference Radar (DInSAR), Small Baseline Subset InSAR (SBAS-InSAR), and Persistent Scatterer InSAR (PS-InSAR), offer expansive and high-resolution monitoring capabilities, proving essential in arid and mining regions [13,14,15,16,17,18,19].
Oyu Tolgoi (OT), one of the world’s largest copper and gold mines, is located in Mongolia’s Gobi region, which is currently facing severe droughts and water shortages exacerbated by mining activities. These conditions not only lead to vegetation loss and increased sandstorms but also pose a risk of land subsidence, threatening local ecosystems and the safety of the area’s inhabitants [20,21,22].
Given the lack of extensive research on the Oyu Tolgoi mine’s subsidence issues, this study employs the SBAS-InSAR technique to monitor and analyze the spatiotemporal evolution of surface subsidence from 2018 to 2022 using Sentinel-1A data. This research aims to elucidate the causes of subsidence at the OT mine and provide data critical for ensuring the mine’s operational safety and environmental protection.

2. Materials

2.1. Study Area

The Oyu Tolgoi (OT) mining area, situated in the Khanbogd soum of Mongolia’s South Gobi province, is one of the world’s premier sites for copper and gold extraction. Positioned within the expansive Gobi Desert, the mine lies approximately 560 km south of Ulaanbaatar, the capital of Mongolia, and about 80 km north of the Chinese border. Its geographical coordinates span from 42°58′30″N to 43°03′00″N latitude and from 106°47′30″E to 106°55′00″E longitude (Figure 1).
Since Mongolia transitioned to a market economy in 1990, OT has emerged as a cornerstone for the nation’s economic development, showcasing its significant mineral wealth. The mining operations are concentrated within a mineralized corridor that extends about 8 km in length and 1 km in width, oriented north–northeast. This corridor encompasses several mining zones, with the main zones centered in the area. These zones vary in elevation from approximately 1140 m above sea level (masl) to 1215 masl, divided into four primary mineralized areas: southwest, south, central, and far north. The predominant extractions from this mine are copper and gold, which are critical to Mongolia’s mining output.

2.2. Datasets

The surface deformation in the Oyu Tolgoi (OT) mining area has been meticulously monitored using the Sentinel-1A SAR datasets, a key component of the European Space Agency’s (ESA) Copernicus Earth Observation Programme. Equipped with a C-band Synthetic Aperture Radar (SAR), Sentinel-1A boasts a revisit period of 12 days, enabling it to provide consistent, all-weather, day-and-night observations globally.
For this study, data were captured in Interferometric Wide Swath (IW) mode with VV polarization. This mode was chosen due to its suitability for capturing wide-area surface deformations with high resolution. Based on the vector files of the study area, Sentinel-1A datasets were collected from the Alaska Satellite Facility (ASF) (https://search.asf.alaska.edu/, accessed on 12 February 2023), selecting 120 coherent single-look complex (SLC) scenes from Path 84, Frame 135 recorded during an ascending orbit, covering the entire year from 2018 to 2022. The images offer a pixel spacing of 2.3 m in slant range and 13.9 m in azimuth, detailed further in Table 1.
Crucial to the SBAS-InSAR process is the use of precise orbit files, which ensure exact alignment between the SAR images over time, enhancing the inversion accuracy and yielding more precise deformation results [23]. Accordingly, precise orbit determination ephemeris data (https://scihub.copernicus.eu/gnss/#/home, accessed on 13 February 2023) corresponding to the Sentinel-1A images were employed. Additionally, the Shuttle Radar Topography Mission (SRTM) Digital Elevation Model (DEM) from NASA (http://gdex.cr.usgs.gov/gdex/, accessed on 16 February 2023), with a spatial resolution of 30 m, was utilized for orbit refining and topographic flattening.

3. Methods

This study harnesses Synthetic Aperture Radar (SAR) technology combined with Small Baseline Subset (SBAS) interferometric processing to analyze spatiotemporal deformation patterns in the Oyu Tolgoi mining area, situated in Mongolia’s Gobi Desert. Sentinel-1A orbit lifting data were utilized, with the reliability of the inversion results cross-validated using Persistent Scatterer InSAR (PS-InSAR) technology and prior studies on groundwater levels.

3.1. InSAR, DInSAR, and Time-Series InSAR

InSAR is a three-dimensional surface morphology information extraction technique that uses the phase information of SAR images for interferometric processing, thus enabling the extraction of surface elevation information [24]. DInSAR technology is an extension of InSAR technology. It mainly uses the phase information from SAR satellite images to measure surface deformation [25,26,27]. DInSAR has gained significant prominence as a widely utilized technique for deformation monitoring.
However, DInSAR processing is subject to various errors. Although deformation censuses can be performed, the limited accuracy of deformation does not allow for a detailed depiction of deformation over time-series and, thus, has certain shortcomings in the monitoring and warning of deformation. Several researchers have proposed different types of time-series InSAR analysis techniques to address the limitations of DInSAR [28,29,30,31]. Over two decades of progress and data accumulation have led to the gradual shift from traditional DInSAR to time-series InSAR as the primary method for acquiring surface deformation data. This approach can be categorized into two main types: PS-InSAR primarily for urban built-up areas and SBAS-InSAR mainly for countryside vegetation areas [32].

3.2. Principles of SBAS-InSAR

The SBAS-InSAR technology proposed by Berardino is a time-series InSAR method that can be used to monitor the temporal evolution of surface deformations [33]. Although it can be simply understood as a combination of multiple DInSAR, SBAS-InSAR effectively minimizes spatial and temporal correlation and atmospheric disturbances, enhancing measurement accuracy and enabling time-series analysis. Compared to labor-intensive traditional geodetic surveys like leveling and GPS, which struggle with large-area surface deformation monitoring, SBAS-InSAR offers efficient, all-weather, and wide-coverage deformation monitoring [34].
It obtains a series of short spatial baseline interferometric pairs by setting specific temporal and spatial baseline thresholds. Then, it performs differential interferometric processing on these interferometric pairs and utilize high-pass filtering of the temporal thresholds and low-pass filtering of the spatial thresholds to remove atmospheric interferences, so as to obtain the correct unwrapped phase information. Finally, the singular value decomposition (SVD) method was employed to invert the time-series deformation results based on the flexible combination of interferograms within a subset. The basic principles are as follows.
Assuming that there are N + 1 SAR images covering the same area acquired in a time-series ( t 0 , , t N ) and selecting one of them as the master image, the remaining images are aligned to it. Then, interferometric pairs are combined to produce M differential interferograms based on the spatiotemporal baseline threshold, where
( N + 1 ) 2 M N N + 1 2
Taking the moment t 0 as the reference moment, φ t i is the differential phase at the time t i i = 1 , , N . For the I-th differential interferogram generated at t A and t B , the interference phase value of any pixel r , x is
δ φ i r , x = φ t A , r , x φ t B , r , x   δ φ d e f , i + δ φ t o p , i + δ φ a t m , i + δ φ n o i , i
where δ φ d e f , i is the line-of-sight (LOS) deformation phase; δ φ t o p , i is the terrain phase error; δ φ a t m , i is the atmospheric phase error; and δ φ n o i , i is the noise phase error. Eliminating the terrain phase, atmosphere, and noise, the expression of δ φ i r , x can be simplified as
δ φ i r , x δ φ d e f , i = 4 π λ d t A , r , x d t B , r , x
where λ is the is the wavelength of the radar signal, and d t A , r , x and d t B , r , x are the LOS cumulative deformations at moments t A and t B , respectively. To derive a physically meaningful subsidence sequence, Equation (3) is formulated as the product of the average phase velocity and the time interval between two acquisition times:
v i = φ i φ i 1 t i t i 1
The value of the i-th differential interferogram can be reformulated as
δ φ i = k = t A , i + 1 t B , i t k t k 1 v k
In other words, it represents the integration of the velocity over each time period across the master and slave image time intervals. Equation (5) can be rewritten into the matrix form as
B v = δ φ
Equation (6) is a matrix with the dimension of M × N. If there are more than one small baseline sets, matrix B exhibits a rank deficiency. In such situations, the SVD method can be employed to perform the singular value decomposition of coefficient matrix B, allowing for the joint solution of multiple small baseline sets. Consequently, a least-squares solution was obtained, minimizing the norm of the cumulative shape variable and resulting in the estimation of the shape variable.

3.3. Data Processing

The processing of surface deformation data acquired through the SBAS-InSAR technique is comprehensive, encompassing several key phases: data preprocessing, SBAS-InSAR processing, and result visualization. Initially, the 120 collected SAR images were imported, aligned with precise orbit files, and clipped to precisely fit the study area. After that, SBAS-InSAR processing was performed. Finally, the results can be visualized. The primary steps of SBAS-InSAR processing include the following:
(1)
Connection Graph Generation: This step involves estimating the baseline and selecting the primary image for the dataset. The image taken on March 10 was designated as the super master image, and all other images were processed in a multi-view setup.
(2)
Interference Processing: This stage processes all interferometric pairs to generate coherence, perform flattening, filter data, and unwrap phases, resulting in unwrapped interferograms. External Digital Elevation Model (DEM) data were utilized here to simulate and correct the topographic phase, enhancing the accuracy of the results.
(3)
Orbit Refining and Re-flattening: Involves selecting control points to refine the orbit and correct the flattened data. The software automatically selects several Ground Control Points (GCPs) to estimate the residual phase in the initial unwrapping phase and remove any residual topographic phase.
(4)
SBAS Inversion: Comprises two distinct inversions. The first inversion estimates deformation rates and residual topography. The second inversion processes these estimates further to optimize the deformation results, refining the output for better accuracy. In this step, the SVD method was utilized to invert the time-series deformation results and the time-series surface subsidence maps were generated in the SAR coordinate system.
(5)
Geocoding: This final step sets a specific threshold to ensure the reliability of time-series deformation results. These results are then geocoded and converted from the SAR coordinate system to the geographic coordinate system, making them suitable for practical use and analysis.
Figure 2 illustrates the SBAS-InSAR technique, showing these five primary steps in detail, which effectively capture the complex process of transforming raw SAR data into actionable geospatial information.

4. Results

4.1. Characteristics of surface subsidence

As determined in Section 2.2, the SAR image from 10 March 2020, was used as the super master image. Figure 3 presents the spatiotemporal baseline diagram, which includes both the Time Baseline Map and the Spatial Baseline Map. In this figure, yellow dots represent the super master images, green dots denote the other SAR images, and the line segments between them depict the interferometric pairs used.
The interferometric processing of these pairs involved the use of the Goldstein filtering method and the Minimum Cost Flow (MCF) technique for untangling, which helped generate clear differential interferograms while eliminating topographic phase influences. Out of the 471 filtered interferograms, three were selected for their significant depiction of surface deformation. Figure 4 shows these interferograms with clear interference fringes, indicating deformation within the mining area.
After completing the SBAS inversion and geocoding steps, the final surface deformation data for the OT mining area were obtained and graphically represented. Figure 5 displays the annual average surface deformation rates, highlighting areas of subsidence and uplift.
The satellite line-of-sight (LOS) velocity map (Figure 5) illustrates that red and yellow areas, which indicate negative deformation rates, signify subsidence—i.e., the ground surface moving away from the satellite. Conversely, blue areas show positive deformation rates, indicating uplift—i.e., the ground surface moving towards the satellite. Most of the OT mine area is shown in yellow and red, with a few areas showing blue. This means the OT mine workings predominantly underwent significant subsidence with notably high rates, while a minor section of the region exhibited slight uplift during this time. The varying shades of red highlight different degrees of subsidence, with darker reds indicating more severe subsidence, which means that the subsidence velocity varied among the different areas. Moreover, the subsidence velocity decreases gradually from the center outwards. Notably, the darkest red square area on the right exhibits the highest subsidence velocity, with parts of the areas experiencing an annual average subsidence velocity of −158.11 mm/year. Such a high rate of subsidence annually may have serious implications for infrastructure stability, mining safety, and the environment. Additionally, a slight uplift is noted in a blue-colored area within the mine. Outside the primary working zones, the terrain of the OT mine remains relatively stable with no significant changes in elevation.

4.2. Time-Series Analysis of Surface Subsidence

To analyze the temporal variability of surface subsidence at the OT mine site over this period, we mapped and compared accumulated surface deformation data from 11 January 2018, to 28 December 2022 (Figure 6). This figure highlights nine cumulative surface deformation datasets across various time intervals, offering a nuanced view of the temporal and spatial dynamics within the OT mining regions. Over the study period, specific areas of the mine exhibited a progressive increase in uplift, denoted by an intensifying blue hue. In contrast, conspicuous yellow and red discolorations emerged, signaling the gradual development of subsidence funnels.
These findings indicate ongoing surface subsidence from 2018 to 2022, with the affected area and the degree of subsidence both expanding over time. Notably, the central area showed higher cumulative subsidence compared to the peripheral edges. The water storage areas on the mine’s right side recorded the highest cumulative subsidence, with parts of the areas experiencing a cumulative subsidence of −741.01 mm from 2018 to 2022.
Figure 5 presents the meticulous selection of ten characteristic subsidence points within the primary subsidence zones, chosen for comprehensive subsidence analysis. The distribution of these points varies within the study area. Specifically, P1 to P7 are situated near the mining area, while P8 to P10 are positioned in the water storage area, all integral to the primary operational zones. Simultaneously, these points exhibit varying degrees of subsidence. To investigate the temporal evolution of subsidence at specific feature points, a time-series subsidence map was generated for each one. Figure 7 illustrates that P2 and P6 display minor subsidence, followed by P1, P3, P5, and P7. P4, P8, and P10 exhibit more substantial subsidence, with P9 demonstrating the most significant subsidence, with the overall cumulative subsidence in the range [0, −400] mm over time. Furthermore, we note that the variation in cumulative subsidence at each point remains relatively minimal until 2020. However, starting from that point, these differences progressively intensify.

4.3. Deformation Velocity Analysis

To facilitate the analysis of deformation rates, the data were categorized and visually represented through a color grading system. This stratification helps in understanding the varying velocities of surface deformation and its spatial distribution patterns across different regions of the OT mine. As depicted in Figure 8, the mine area has been segmented into zones labeled A through F, reflecting different subsidence characteristics.
In this figure, subsidence is indicated by a spectrum from red to light blue, with red representing the most severe subsidence rates. Uplift is shown in shades of blue, with darker blues indicating higher uplift rates. To aid detailed analysis, subsidence rates were categorized into five classes:
  • Class I: [−160, −80] mm/year;
  • Class II: [−80, −50] mm/year;
  • Class III: [−50, −30] mm/year;
  • Class IV: [−30, −15] mm/year;
  • Class V: [−15, 0] mm/year.
Zone A shows the most pronounced subsidence, with concentric color bands progressing from red at the center to light blue at the outer edges, representing a decrease in subsidence severity from Class I through Class V. Zones B, C, D, and E display similar patterns, with subsidence decreasing from the center outward. Specifically, Zone C transitions from orange at the center through yellow, green, and light blue, corresponding to Class II through Class V subsidence rates, respectively. Zones B, D, and E exhibit gradients from yellow to light blue, indicating Classes III to V.
Notably, Zone F illustrates a unique pattern with subsidence shown in yellow on the left side, while uplift is depicted by dark blue on the right. Blank areas in the map indicate regions where monitoring data are unavailable.

4.4. Precision Analysis of SBAS-InSAR Surface Deformation Results

Conducting field-level surveys for data validation at the OT site and obtaining usable GNSS data is extremely challenging due to stringent policies and regulations as well as lack of instrumentation deployment such as GNSS. In the absence of direct empirical data, cross-validation methods such as PS-InSAR are frequently utilized to confirm the reliability of monitoring outcomes [35,36]. Correlation analysis was conducted on the deformation using both SBAS-InSAR and PS-InSAR techniques, which showed great correlation, as depicted in Figure 9. In addition, numerous studies have consistently indicated a strong link between surface subsidence and groundwater levels [37,38]. Specifically, in the OT mine area, Okadera et al. [39] noted significant reductions in groundwater depths, ranging from 0.11 to 1.64 m, thereby highlighting the substantial environmental impacts resulting from mining activities.

5. Discussion

This section explores the potential factors influencing surface subsidence at the OT mine site, focusing on changes in surface stress, mining activities, and groundwater levels. As shown in Figure 8, the mine area has been categorized into six distinct zones (A to F), each exhibiting unique subsidence characteristics based on their geological and operational contexts.
Zone A, which includes the mine’s water storage and adjacent wetland areas, experiences more pronounced subsidence compared to other regions. This is likely due to the accumulation of silt, suspended sediments, and organic matter that consolidate over time, resulting in substantial ground settlement [40]. Zones B through E, characterized by the presence of mineral sands, experience consolidation settlement as loose soils are compacted and dehydrated under external pressures. These areas are inherently prone to subsidence due to the geological processes involved [41]. Zone F, marked by its open-pit mining operations, displays varied subsidence patterns: substantial subsidence in the western and southern sectors, and slight elevation in the north, possibly from the accumulation of slag from ongoing mining activities. The above factors may contribute to the spatial variations of subsidence.
Table 2 and Table 3 detail the mineral extraction data from the OT mine’s annual reports [42,43] and surface cumulative subsidence records between 2018 and 2022, respectively. Notably, except for Point 5, subsidence at the mine has progressively increased alongside the volume of extracted minerals (Table 2). Figure 10a illustrates a strong positive correlation between cumulative subsidence and mining activities at selected sites, with correlation coefficients all surpassing 0.89 (p < 0.05), suggesting that ongoing mining activities significantly contribute to the observed surface subsidence.
Furthermore, Figure 10b examines the correlation between annual subsidence levels and mineral extraction across different mining areas. Points 1, 3, 5, and 6, which are closer to active mining zones, show higher correlation with annual mining outputs compared to Points 2 and 4. This differential suggests that proximity to mining activities plays a crucial role in the extent of surface deformation experienced, as mining operations likely alter surface stress and induce rock mass deformations, such as collapses and fractures [44,45].
The impact of groundwater extraction on surface subsidence at the mine site was further explored. The OT mine, located in the arid South Gobi region of Mongolia, where water resources are scarce, heavily relies on groundwater to meet its operational needs due to the insufficiency of surface water. Figure 11 depicts the significant volumes of groundwater extracted annually from 2016 to 2020 [46], which have led to alterations in groundwater levels, potentially triggering terrain subsidence above the extraction sites.
Subsequent well monitoring near the mine site has shown a notable decrease in water levels, with reductions ranging from 16 to 46 m below established benchmarks across 28 wells, as reported by the Human Pipeline Groundwater Palace. These findings are consistent with observations by Okadera et al. [39], who noted declines in groundwater levels from 0.11 to 1.64 m in the OT mine area, and Nakayama et al. [47], who documented significant decreases in the Galba River’s groundwater levels due to mining activities starting in 2010. These consistent studies and reports indicate the decline in groundwater levels is also a contributing factor to surface subsidence at the mine site, as it causes subsurface rocks to expand and contract [48].
SBAS-InSAR-based measurements have been instrumental in providing essential insights for the sustainable development of mining sites. However, despite these advances, challenges persist, including unmeasured deformation at specific mining locations and extensive blank areas within the study zone. These issues may arise from data decoherence during differential interference, a limitation of InSAR noted by Ferretti et al. [49], or from rapid and extensive changes over brief intervals. Consequently, data decoherence obscures part of the measured region, leading to their appearance as blanks on the deformation result map. Notably, the blank regions in Zones A and F are primarily due to significant daily water level fluctuations and variations in the vertical heights of stockpiles, which induce decoherence and complicate the monitoring process.
This study leveraged SBAS-InSAR technology to explore surface subsidence at the OT mine site, yet several limitations are notable:
(1)
Spatial Coverage: SBAS-InSAR’s reliability depends on the coherence of satellite images over time. Rapid landscape changes due to mining or natural processes can disrupt this coherence, limiting the accuracy of subsidence measurements.
(2)
Temporal Resolution: the 12-day revisit cycle of the Sentinel-1A satellite may miss rapid subsidence events, potentially leading to underestimations of subsidence rates and extents.
(3)
Vertical Accuracy: despite corrections for atmospheric effects, inherent limitations due to atmospheric disturbances and the satellite’s viewing angle may still impact the vertical accuracy of measurements.
(4)
Supplementary Data: The study’s validation and analysis of subsidence is limited by the availability and accuracy of measured data, GNSS data, groundwater data, and mining volumes. Correlating subsidence with groundwater levels is also challenged by the sporadic availability and variable accuracy of groundwater data, complicating the establishment of a direct causative relationship.
(5)
External Factors and Generalizability: the analysis does not extensively cover other potential subsidence drivers, like geological processes or external human activities, and findings may not apply to other regions with different conditions.
Future research should address these limitations by integrating more frequent imaging, refined groundwater data, refined mining volumes, measured data, available GNSS data, and broader environmental factors to enhance the robustness and applicability of the findings. To mitigate the potential threats of subsidence to infrastructure stability, mining safety, water resources, and the environment, it is necessary for OT mines to adopt the following measures to reduce ongoing subsidence: continuous on-site monitoring, reinforcement of infrastructure, improved management of water resources, and a reduction in groundwater extraction [50,51,52].

6. Conclusions

This study leveraged Sentinel-1A data from 2018 to 2022 and SBAS-InSAR technology to effectively monitor and analyze surface subsidence at the OT mine in Mongolia. It successfully delineated the spatiotemporal characteristics of subsidence and pinpointed key contributing factors. The significant findings are as follows:
Substantial Subsidence: the OT mine has experienced notable surface subsidence, with the most severe cumulative decrease reaching −742.01 mm and the highest annual subsidence rate at −158.11 mm/year.
Efficacy of SBAS-InSAR: This technology has proven extremely effective in arid, sparsely vegetated areas like the OT mine, showcasing its enhanced capability for accurate subsidence detection and making it a valuable tool for monitoring similar environments.
Influential Factors: the analysis confirmed that significant fluctuations in groundwater levels, changes in surface stress, and intensive mining activities are the primary drivers behind the observed subsidence.
These insights are valuable both practically and academically; they inform sustainable mining practices and contribute to geological hazard prevention worldwide, while also enriching the understanding of subsidence monitoring in Mongolian mining regions and similar settings.
Despite its contributions, this study acknowledges limitations related to the temporal and spatial resolution of SBAS-InSAR technology, which may overlook rapid subsidence events and experience data coherence issues in dynamic environments. Additionally, the variability in groundwater data accuracy and the challenges in generalizing findings underscore the need for caution.
Future research should focus on mitigating subsidence in mining areas by integrating a broader spectrum of environmental data, including high-frequency InSAR images and detailed groundwater measurements. Such an approach promises a more comprehensive evaluation of long-term impacts and the refinement of monitoring technologies applicable to diverse geological settings. By addressing the current limitations, further studies can improve prediction accuracy and broaden the scope of these findings to more global applications.

Author Contributions

Conceptualization, Y.X.; Methodology, L.T.; Data processing, L.T.; Validation, Y.X.; Investigation, Y.X.; Data curation, A.D.; Writing—original draft, Y.X.; Writing—review & editing, Q.W.; Visualization, T.T.; Supervision, A.D. and Q.W.; Project administration, H.B.; Funding acquisition, H.B. and Q.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Key R&D Program of China (No. 2022YFE0119500) and the Research on Advancement of Climate Change Impact Assessment Methodology (No. 2125AA130).

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Acknowledgments

The authors wish to express their gratitude to the colleagues and staff at the School of Environmental and Geographical Sciences, Shanghai Normal University for their support and invaluable contributions throughout the course of this research. We are also grateful for the funding support received from the National Key R&D Program of China (No. 2022YFE0119500) and the Research on Advancement of Climate Change Impact Assessment Methodology (No. 2125AA130), National Institute for Environmental Studies of Japan, which made this research possible. Lastly, our gratitude extends to all the peer reviewers and editors for their constructive comments that significantly improved the manuscript.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Schoderer, M.; Karthe, D.; Dombrowsky, I.; Dell’Angelo, J. Hydro-social dynamics of miningscapes: Obstacles to implementing water protection legislation in Mongolia. J. Environ. Manag. 2021, 292, 112767. [Google Scholar] [CrossRef] [PubMed]
  2. Zandariya, B. Improving the policy framework for financial assurance for mine closure in Mongolia. Resour. Policy 2022, 77, 102628. [Google Scholar] [CrossRef]
  3. Batdelger, T.; Zagdbazar, M. Does mining improve rural livelihood?: Evidence from Mongolia. Resour. Policy 2022, 78, 102794. [Google Scholar] [CrossRef]
  4. Fraser, J.; Kunz, N.C.; Batdorj, B. Can mineral exploration projects create and share value with communities? A case study from Mongolia. Resour. Policy 2019, 63, 101455. [Google Scholar] [CrossRef]
  5. Prior, T.; Giurco, D.; Mudd, G.; Mason, L.; Behrisch, J. Resource depletion, peak minerals and the implications for sustainable resource management. Glob. Environ. Chang. 2012, 22, 577–587. [Google Scholar] [CrossRef]
  6. Wantzen, K.M.; Mol, J.H. Soil erosion from agriculture and mining: A threat to tropical stream ecosystems. Agriculture 2013, 3, 660–683. [Google Scholar] [CrossRef]
  7. Zhengfu, B.; Inyang, H.I.; Daniels, J.L.; Frank, O.T.T.O.; Struthers, S. Environmental issues from coal mining and their solutions. Min. Sci. Technol. 2010, 20, 215–223. [Google Scholar]
  8. Zhang, L.; Zhai, Z.; Zhou, Y.; Liu, S.; Wang, L. The Landscape Pattern Evolution of Typical Open-Pit Coal Mines Based on Land Use in Inner Mongolia of China during 20 Years. Sustainability 2022, 14, 9590. [Google Scholar] [CrossRef]
  9. Conde, M.; Le Billon, P. Why do some communities resist mining projects while others do not? Extr. Ind. Soc. 2017, 4, 681–697. [Google Scholar] [CrossRef]
  10. Cui, X.; Wang, J.; Liu, Y. Prediction of progressive surface subsidence above longwall coal mining using a time function. Int. J. Rock Mech. Min. Sci. 2001, 38, 1057–1063. [Google Scholar] [CrossRef]
  11. Ding, R.; Liu, H.; Yuan, G.; Dong, J. Deformation and subsidence prediction on surface of Yuzhou mined-out areas along middle route project of south-to-north water diversion, China. Open Geosci. 2018, 10, 834–843. [Google Scholar] [CrossRef]
  12. Siriwardane, H.J.; Amanat, J. Analysis of subsidence caused by underground mining. Int. J. Min. Sci. Technol. 1984, 2, 271–290. [Google Scholar] [CrossRef]
  13. Simons, M.; Fialko, Y.; Rivera, L. Coseismic deformation from the 1999 M w 7.1 Hector Mine, California, earthquake as inferred from InSAR and GPS observations. Bull. Seismol. Soc. Am. 2002, 92, 1390–1402. [Google Scholar] [CrossRef]
  14. Nie, L.; Wang, H.; Xu, Y.; Li, Z. A new prediction model for mining subsidence deformation: The arc tangent function model. Nat. Hazards 2018, 75, 2185–2198. [Google Scholar] [CrossRef]
  15. Wempen, J.M.; McCarter, M.K. Comparison of L-band and X-band differential interferometric synthetic aperture radar for mine subsidence monitoring in central Utah. Int. J. Min. Sci. Technol. 2017, 27, 159–163. [Google Scholar] [CrossRef]
  16. Hui, X.; Ma, F.; Zhao, H.; Xu, J. Monitoring and statistical analysis of mine subsidence at three metal mines in China. Bull. Seismol. Soc. Am. 2019, 78, 3983–4001. [Google Scholar] [CrossRef]
  17. Casu, F.; Manzo, M.; Lanari, R. A quantitative assessment of the SBAS algorithm performance for surface deformation retrieval from DInSAR data. Remote Sens. Environ. 2006, 102, 195–210. [Google Scholar] [CrossRef]
  18. Kim, J.-R.; Lin, C.-W.; Lin, S.-Y. The use of InSAR phase coherence analyses for the monitoring of aeolian erosion. Remote Sens. 2021, 13, 2240. [Google Scholar] [CrossRef]
  19. Ghorbani, Z.; Khosravi, A.; Maghsoudi, Y.; Mojtahedi, F.F.; Javadnia, E.; Nazari, A. Use of InSAR data for measuring land subsidence induced by groundwater withdrawal and climate change in Ardabil Plain, Iran. Sci. Rep. 2022, 12, 13998. [Google Scholar] [CrossRef] [PubMed]
  20. Hongming, W.; Ruren, L.; Yiting, Q.; Zhuqing, L.I.; Jun, G. Application of time series InSAR technology in monitoring ground deformation of mining area: A case study at Huolinhe open pit mining area in Inner Mongolia. Chin. J. Geol. Hazard Control 2022, 33, 71–78. [Google Scholar] [CrossRef]
  21. Zhou, D.; Wang, L.; An, S.; Wang, X.; An, Y. Integration of unmanned aerial vehicle (UAV)-based photogrammetry and InSAR for mining subsidence and parameters inversion: A case study of the Wangjiata Mine, China. Bull. Seismol. Soc. Am. 2022, 81, 343. [Google Scholar] [CrossRef]
  22. Pawluszek-Filipiak, K.; Borkowski, A. Monitoring mining-induced subsidence by integrating differential radar interferometry and persistent scatterer techniques. Eur. J. Remote Sens. 2021, 54, 18–30. [Google Scholar] [CrossRef]
  23. Yang, W.; He, Y.; Wang, W.H.; Chen, Y.D.; Chen, Y. InSAR monitoring of 3-D surface deformation in Jinchuan Mining area, Gansu Province. Remote Sens. Nat. Resour. 2022, 34, 177–188. [Google Scholar]
  24. Wegmttller, U.; Strozzi, T.; Werner, C.; Wiesmann, A.; AG, G.R.S. Sar interferometry for topographic mapping and surface deformation monitoring. FJP 2002, 18, 24–32. [Google Scholar]
  25. Massonnet, D.; Feigl, K.L. Radar interferometry and its application to changes in the Earth’s surface. Rev. Geophys. 1998, 36, 441–500. [Google Scholar] [CrossRef]
  26. Rosen, P.A.; Hensley, S.; Joughin, I.R.; Li, F.K.; Madsen, S.N.; Rodriguez, E.; Goldstein, R.M. Synthetic aperture radar interferometry. Proc. IEEE 2000, 88, 333–382. [Google Scholar] [CrossRef]
  27. Hanssen, R.F. Radar Interferometry: Data Interpretation and Error Analysis; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2001; pp. 1–307. [Google Scholar]
  28. Ferretti, A.; Prati, C.; Rocca, F. Permanent scatterers in SAR interferometry. IEEE Trans. Geosci. Remote Sens. 2001, 39, 8–20. [Google Scholar] [CrossRef]
  29. Colesanti, C.; Ferretti, A.; Novali, F.; Prati, C.; Rocca, F. SAR monitoring of progressive and seasonal ground deformation using the permanent scatterers technique. IEEE Trans. Geosci. Remote Sens. 2003, 41, 1685–1701. [Google Scholar] [CrossRef]
  30. Hooper, A.; Zebker, H.A. Phase unwrapping in three dimensions with application to InSAR time series. JOSA A 2007, 24, 2737–2747. [Google Scholar] [CrossRef] [PubMed]
  31. Rocca, F. Modeling interferogram stacks. IEEE Trans. Geosci. Remote Sens. 2007, 45, 3289–3299. [Google Scholar] [CrossRef]
  32. Zhang, P.; Guo, Z.; Guo, S.; Xia, J. Land subsidence monitoring method in regions of variable radar reflection characteristics by integrating PS-InSAR and SBAS-InSAR techniques. Remote Sens. 2022, 14, 3265. [Google Scholar] [CrossRef]
  33. Berardino, P.; Fornaro, G.; Lanari, R.; Sansosti, E. A new algorithm for surface deformation monitoring based on small baseline differential SAR interferograms. IEEE Trans. Geosci. Remote Sens. 2002, 40, 2375–2383. [Google Scholar] [CrossRef]
  34. Li, S.; Xu, W.; Li, Z. Review of the SBAS InSAR Time-series algorithms, applications, and challenges. Geod. Geodyn. 2022, 13, 114–126. [Google Scholar] [CrossRef]
  35. He, Y.; Yan, H.; Yang, W.; Yao, S.; Zhang, L.; Chen, Y.; Liu, T. Time-series analysis and prediction of surface deformation in the Jinchuan mining area, Gansu Province, by using InSAR and CNN–PhLSTM network. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2022, 15, 6732–6751. [Google Scholar] [CrossRef]
  36. Li, Z.; Wang, Z.; Liu, W.; Li, X.; Zhou, M.; Zhang, B. Detecting, Monitoring, and Analyzing the Surface Subsidence in the Yellow River Delta (China) Combined with CenterNet Network and SBAS-InSAR. J. Spectrosc. 2022, 2022, 2672876. [Google Scholar] [CrossRef]
  37. Poland, J.; Davis, G. Subsidence of the land surface in the Tulare-Wasco (Delano) and Los Banos-Kettleman City area, San Joaquin Valley, California. EOS Trans. Am. Geophys. Union 1956, 37, 287–296. [Google Scholar] [CrossRef]
  38. Motagh, M.; Shamshiri, R.; Haghighi, M.H.; Wetzel, H.U.; Akbari, B.; Nahavandchi, H.; Roessner, S.; Arabi, S. Quantifying groundwater exploitation induced subsidence in the Rafsanjan plain, southeastern Iran, using InSAR time-series and in situ measurements. Eng. Geol. 2017, 218, 134–151. [Google Scholar] [CrossRef]
  39. Okadera, T.; Wang, Q.X.; Deni, E.; Nakayama, T. Groundwater monitoring for evaluating the pasture carrying capacity and its vulnerability in arid and semi-arid regions: A case study of urban and mining areas in Mongolia. IOP Conf. Ser. Earth Environ. Sci. 2019, 266, 012013. [Google Scholar] [CrossRef]
  40. Shi, C.; Zhang, D.; You, L.; Li, B.; Zhang, Z.; Zhang, O. Land subsidence as a result of sediment consolidation in the Yellow River Delta. J. Coast. Res. 2007, 23, 173–181. [Google Scholar] [CrossRef]
  41. Steckler, M.S.; Oryan, B.; Wilson, C.A.; Grall, C.; Nooner, S.L.; Mondal, D.R.; Akhter, S.H.; DeWolf, S.; Goodbred, S.L. Synthesis of the distribution of subsidence of the lower Ganges-Brahmaputra Delta, Bangladesh. Earth-Sci. Rev. 2022, 224, 103887. [Google Scholar] [CrossRef]
  42. Oyu Tolgoi LLC. Year in Review 2020. 2021. Available online: https://www.ot.mn/media/otnew/content/reports/2020_YiR/20210428_Annual_report_2020-English.pdf (accessed on 7 March 2024).
  43. Oyu Tolgoi LLC. Year in Review 2022. 2023. Available online: https://www.ot.mn/media/otnew/content/Year_in_review_2022.pdf (accessed on 7 March 2024).
  44. Parsons, T.; Wu, P.C.; Wei, M.; D’Hondt, S. The weight of New York City: Possible contributions to subsidence from anthropogenic sources. Earth’s Future 2023, 11, e2022EF003465. [Google Scholar] [CrossRef]
  45. Kratzsch, H. Mining Subsidence Engineering; Springer: Berlin/Heidelberg, Germany, 1983; ISBN 978-3-642-81925-4. [Google Scholar]
  46. Oyu Tolgoi LLC. Audit Report on Water Use for the Bear Head Project 2016–2020. 2022. Available online: https://www.ot.mn/media/otnew/content/Water_Audit_report_2021_Final_MNG.pdf (accessed on 7 March 2024).
  47. Nakayama, T.; Wang, Q.; Okadera, T. Evaluation of spatio-temporal variations in water availability using a process-based eco-hydrology model in arid and semi-arid regions of Mongolia. Ecol. Model. 2021, 440, 109404. [Google Scholar] [CrossRef]
  48. Donnelly, L.J. A review of international cases of fault reactivation during mining subsidence and fluid abstraction. Q. J. Eng. Geol. Hydrogeol. 2009, 42, 73–94. [Google Scholar] [CrossRef]
  49. Ferretti, A.; Monti-Guarnieri, A.; Prati, C.; Rocca, F.; Massonet, D. InSAR Principles: Guideline for SAR Interferometry Processing and Interpretation; ESA Publication: Noordwijk, The Netherlands, 2007. [Google Scholar]
  50. Sahu, P.; Lokhande, R.D. An Investigation of Sinkhole Subsidence and its Preventive Measures in Underground Coal Mining. Procedia Earth Planet. Sci. 2015, 11, 63–75. [Google Scholar] [CrossRef]
  51. Kondolf, G.M.; Schmitt, R.J.P.; Carling, P.; Darby, S.; Arias, M.; Bizzi, S.; Castelletti, A.; Cochrane, T.A.; Gibson, S.; Kummu, M.; et al. Changing sediment budget of the Mekong: Cumulative threats and management strategies for a large river basin. Sci. Total Environ. 2018, 625, 114–134. [Google Scholar] [CrossRef] [PubMed]
  52. Tang, W.; Zhao, X.; Motagh, M.; Bi, G.; Li, J.; Chen, M.; Chen, H.; Liao, M. Land Subsidence and Rebound in the Taiyuan Basin, Northern China, in the Context of Inter-Basin Water Transfer and Groundwater Management. Remote Sens. Environ. 2022, 269, 112792. [Google Scholar] [CrossRef]
Figure 1. Geographical location of the study area—(a) Digital elevation model (DEM) of Mongolia; (b) Sentinel-2 image of OT mine on 30 December 2022.
Figure 1. Geographical location of the study area—(a) Digital elevation model (DEM) of Mongolia; (b) Sentinel-2 image of OT mine on 30 December 2022.
Remotesensing 16 02166 g001
Figure 2. Workflow of SBAS-InSAR technique.
Figure 2. Workflow of SBAS-InSAR technique.
Remotesensing 16 02166 g002
Figure 3. Spatiotemporal baseline diagram—(a) Time Baseline Map; (b) Spatial Baseline Map.
Figure 3. Spatiotemporal baseline diagram—(a) Time Baseline Map; (b) Spatial Baseline Map.
Remotesensing 16 02166 g003
Figure 4. Filtered differential interferograms—(a) Filtered differential interferogram generated from two images acquired on 19 April 2018 and 3 May 2018; (b) Filtered differential interferogram generated from two images acquired on 8 March 2020 and 2 January 2020; (c) Filtered differential interferogram generated from two images acquired on 14 January 2022 and 7 February 2022.
Figure 4. Filtered differential interferograms—(a) Filtered differential interferogram generated from two images acquired on 19 April 2018 and 3 May 2018; (b) Filtered differential interferogram generated from two images acquired on 8 March 2020 and 2 January 2020; (c) Filtered differential interferogram generated from two images acquired on 14 January 2022 and 7 February 2022.
Remotesensing 16 02166 g004
Figure 5. Annual average surface deformation rate (unit: mm/year) with P1 to P10. P1 to P10 indicate ten characteristic subsidence points used for comprehensive time-series subsidence analysis in Section 4.2.
Figure 5. Annual average surface deformation rate (unit: mm/year) with P1 to P10. P1 to P10 indicate ten characteristic subsidence points used for comprehensive time-series subsidence analysis in Section 4.2.
Remotesensing 16 02166 g005
Figure 6. Spatiotemporal distribution of accumulated surface deformation in the study area.
Figure 6. Spatiotemporal distribution of accumulated surface deformation in the study area.
Remotesensing 16 02166 g006
Figure 7. Time-series subsidence map of feature points.
Figure 7. Time-series subsidence map of feature points.
Remotesensing 16 02166 g007
Figure 8. Annual average surface deformation classified rate. Maps to the right are from Google Earth for December 2018. The Points 1 to 6 are six example points used for correlation analysis in Section 5.
Figure 8. Annual average surface deformation classified rate. Maps to the right are from Google Earth for December 2018. The Points 1 to 6 are six example points used for correlation analysis in Section 5.
Remotesensing 16 02166 g008
Figure 9. Accuracy verification.
Figure 9. Accuracy verification.
Remotesensing 16 02166 g009
Figure 10. Correlation matrix—(a) Correlation matrix of cumulative subsidence and cumulative mineral mined (CMM refers to cumulative mineral mined); (b) Correlation matrix of annual subsidence and annual minerals mined (‘Mineral’ refers to mineral mined).
Figure 10. Correlation matrix—(a) Correlation matrix of cumulative subsidence and cumulative mineral mined (CMM refers to cumulative mineral mined); (b) Correlation matrix of annual subsidence and annual minerals mined (‘Mineral’ refers to mineral mined).
Remotesensing 16 02166 g010
Figure 11. Total water withdrawn from wells from 2016 to 2020 (unit: 1000 M3).
Figure 11. Total water withdrawn from wells from 2016 to 2020 (unit: 1000 M3).
Remotesensing 16 02166 g011
Table 1. Summary of Sentinel-1A images used in this study.
Table 1. Summary of Sentinel-1A images used in this study.
Sentinel-1A Data Parameters
No.Acq. DateNo.Acq. DateNo.Acq. Date
111 January 20184130 May 2019817 January 2021
223 January 20184211 June 20198219 January 2021
34 February 20184323 June 20198331 January 2021
428 February 2018445 July 20198412 February 2021
512 March 20184517 July 20198524 February 2021
624 March 20184629 July 2019868 March 2021
75 April 20184710 August 20198720 March 2021
817 April 20184822 August 2019881 April 2021
929 April 2018493 September 20198913 April 2021
1011 May 20185015 September 20199025 April 2021
1123 May 20185127 September 2019917 May 2021
124 June 20185221 October 20199219 May 2021
1316 June 2018532 November 2019936 July 2021
1428 June 20185414 November 20199411 August 2021
1510 July 20185526 November 20199523 August 2021
1622 July 2018568 December 2019964 September 2021
173 August 20185720 December 20199716 September 2021
1815 August 20185813 January 20209828 September 2021
1927 August 2018596 February 20209910 October 2021
208 September 20186018 February 202010022 October 2021
2120 September 2018611 March 20201013 November 2021
222 October 20186213 March 202010215 November 2021
2314 October 20186325 March 202010327 November 2021
2426 October 2018646 April 20201049 December 2021
257 November 20186518 April 202010514 January 2022
2619 November 20186630 April 202010626 January 2022
271 December 20186712 May 20201077 February 2022
2813 December 20186824 May 202010819 February 2022
2925 December 2018695 June 20201093 March 2022
3018 January 20197017 June 202011015 March 2022
3130 January 20197129 June 202011127 March 2022
3211 February 20197211 July 20201128 April 2022
3323 February 20197323 July 202011320 April 2022
347 March 2019744 August 20201142 May 2022
3519 March 20197516 August 202011526 May 2022
3631 March 20197621 September 202011625 July 2022
3712 April 2019773 October 20201176 August 2022
3824 April 20197815 October 202011822 November 2022
396 May 20197920 November 20201194 December 2022
4018 May 2019802 December 202012028 December 2022
‘No.’ is the scene number of the 120 scenes and ‘Acq. Date’ is the acquisition date of Sentinel-1A.
Table 2. Annual mineral mined.
Table 2. Annual mineral mined.
Year20182019202020212022
Mineral Mined (Million tons)91101988597
Cumulative Mineral Mined (Million tons)91192290375473
Table 3. Cumulative subsidence of six example points (see Figure 8 for location).
Table 3. Cumulative subsidence of six example points (see Figure 8 for location).
Year20182019202020212022
Point 161.02 132.61 192.87 245.06 299.24
Point 249.76 105.44 147.36 181.81 203.41
Point 324.13 71.65 103.79 127.75 147.97
Point 414.58 46.70 112.67 143.71 149.14
Point 523.47 85.88 107.60 122.97 119.72
Point 611.27 53.12 69.79 93.26 109.05
unit: mm
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xie, Y.; Bagan, H.; Tan, L.; Te, T.; Damdinsuren, A.; Wang, Q. Time-Series Analysis of Mining-Induced Subsidence in the Arid Region of Mongolia Based on SBAS-InSAR. Remote Sens. 2024, 16, 2166. https://doi.org/10.3390/rs16122166

AMA Style

Xie Y, Bagan H, Tan L, Te T, Damdinsuren A, Wang Q. Time-Series Analysis of Mining-Induced Subsidence in the Arid Region of Mongolia Based on SBAS-InSAR. Remote Sensing. 2024; 16(12):2166. https://doi.org/10.3390/rs16122166

Chicago/Turabian Style

Xie, Yuxin, Hasi Bagan, Luwen Tan, Terigelehu Te, Amarsaikhan Damdinsuren, and Qinxue Wang. 2024. "Time-Series Analysis of Mining-Induced Subsidence in the Arid Region of Mongolia Based on SBAS-InSAR" Remote Sensing 16, no. 12: 2166. https://doi.org/10.3390/rs16122166

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop