Next Article in Journal
Estimating Crop LAI Using Spectral Feature Extraction and the Hybrid Inversion Method
Previous Article in Journal
FFAU—Framework for Fully Autonomous UAVs
Previous Article in Special Issue
An Improved Ground Control Point Configuration for Digital Surface Model Construction in a Coal Waste Dump Using an Unmanned Aerial Vehicle System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Remote Sensing of Mine Site Rehabilitation for Ecological Outcomes: A Global Systematic Review

1
Centre for Mined Land Rehabilitation, Sustainable Minerals Institute, The University of Queensland, Brisbane, QLD 4072, Australia
2
Lincoln Centre for Water and Planetary Health, School of Geography, The University of Lincoln, Lincoln LN6 7TS, UK
3
School of Earth and Environmental Sciences, The University of Queensland, Brisbane, QLD 4072, Australia
*
Author to whom correspondence should be addressed.
Remote Sens. 2020, 12(21), 3535; https://doi.org/10.3390/rs12213535
Submission received: 8 September 2020 / Revised: 21 October 2020 / Accepted: 23 October 2020 / Published: 28 October 2020

Abstract

:
The mining industry has been operating across the globe for millennia, but it is only in the last 50 years that remote sensing technology has enabled the visualization, mapping and assessment of mining impacts and landscape recovery. Our review of published literature (1970–2019) found that the number of ecologically focused remote sensing studies conducted on mine site rehabilitation increased gradually, with the greatest proportion of studies published in the 2010–2019 period. Early studies were driven exclusively by Landsat sensors at the regional and landscape scales while in the last decade, multiple earth observation and drone-based sensors across a diverse range of study locations contributed to our increased understanding of vegetation development post-mining. The Normalized Differenced Vegetation Index (NDVI) was the most common index, and was used in 45% of papers; while research that employed image classification techniques typically used supervised (48%) and manual interpretation methods (37%). Of the 37 publications that conducted error assessments, the average overall mapping accuracy was 84%. In the last decade, new classification methods such as Geographic Object-Based Image Analysis (GEOBIA) have emerged (10% of studies within the last ten years), along with new platforms and sensors such as drones (15% of studies within the last ten years) and high spatial and/or temporal resolution earth observation satellites. We used the monitoring standards recommended by the International Society for Ecological Restoration (SER) to determine the ecological attributes measured by each study. Most studies (63%) focused on land cover mapping (spatial mosaic); while comparatively fewer studies addressed complex topics such as ecosystem function and resilience, species composition, and absence of threats, which are commonly the focus of field-based rehabilitation monitoring. We propose a new research agenda based on identified knowledge gaps and the ecological monitoring tool recommended by SER, to ensure that future remote sensing approaches are conducted with a greater focus on ecological perspectives, i.e., in terms of final targets and end land-use goals. In particular, given the key rehabilitation requirement of self-sustainability, the demonstration of ecosystem resilience to disturbance and climate change should be a key area for future research.

1. Introduction

The practice of ecological restoration aims to facilitate a level of native ecosystem recovery following disturbance impacts on a scale, ranging from ecosystem degradation to complete ecosystem destruction [1,2]. Within this broad area of study, one particular sub-discipline has emerged. With a specific focus on the rehabilitation of landscapes following mining, post-mine rehabilitation is clearly distinct from non-mining restoration projects due to the high-levels of disturbance. As a result, recovery objectives are more likely aligned with new (novel) or hybrid ecosystems made up of native and exotic species that provide acceptable levels of stability and ecosystem functionality [3].
The importance of post-mine rehabilitation is identified by governments across multiple jurisdictions, who have introduced legislation that require mining companies to achieve specified standards for reconstructed landforms, topsoil, vegetation, and water quality. In the United States, the introduction of the Surface Mine Control and Reclamation Act of 1977 [4] represented one of the first federal government regulations to enforce rehabilitation standards, including the filling of remaining voids and pre-specified end land-uses [5]. Since then, numerous jurisdictions have implemented similar legislation administered by governments, aiming for best practice and including key features such as financial assurances, progressive rehabilitation, and targets that are safe, stable, non-polluting and ecologically sustainable [6,7,8,9,10,11].
The most common way for mining companies to demonstrate rehabilitation success and provide confidence for mine closure is through ecological monitoring [12]. This is typically achieved using ground-based transect and quadrat monitoring methods that are robust and repeatable, and subject to an appropriate level of scientific scrutiny [13,14,15,16]. Such an approach provides data on a range of conditions, from the initial stages of landform reshaping, to the demonstration of vegetation cover development and the evidence of complex ecosystem function and structure. Ground-based monitoring methods are often labor-intensive exercises and although technically precise, these programs are constrained spatially, temporally, and financially. While the aim for ground monitoring is a study design that is statistically representative of the sampled population, for the above reasons, these programs tend to sample a small proportion of the landscape, and their effectiveness might be limited due to this scale factor. The International Society for Ecological Restoration (SER) has developed an ‘ecological recovery wheel’ monitoring tool to guide restoration projects and quantify the trajectory of recovery in comparison to analogue native ecosystems [1]. This tool helps to define the attributes and sub-attributes that are necessary to determine successful ecosystem recovery, by focusing on the key areas of structural diversity, species composition, physical conditions, external exchanges, ecosystem function, and absence of threats. Although mined land rehabilitation that returns novel ecosystems might not be appropriate to compare directly with natural analogues [12,17,18], the tool developed by the SER offers an appropriate choice of metrics to measure and quantify the success of rehabilitation efforts.
In contrast to ground monitoring programs that are constrained by small areas, remote sensing offers the opportunity to monitor at a landscape scale to either compliment ground surveys, or replace them entirely. In the last two decades the science of remote sensing and spatial analysis underwent a fundamental shift, with new advances in earth observation and drone technology providing opportunities for rapid and cost effective assessments of multiscale landscape changes, and ecosystem development of restoration and rehabilitation projects [19,20]. However, due to the inherently heterogeneous nature of the rehabilitated landscapes, these areas can be challenging to monitor remotely. As a constructed landscape, rehabilitation is typically made up of discrete patches of vegetation in different stages of development, with underlying soil characteristics and landform slopes and topography that can vary significantly over short distances; particularly when compared to neighboring unmined land [12]. Additionally, rehabilitation managers often vary inputs, such as seed mixes, topsoil depth, and site preparation techniques, resulting in high levels of spectral heterogeneity within and between patches, across post-mining landscapes [21,22,23,24,25,26,27,28,29,30]. Nonetheless, the advantage of a remote sensing approach that uses discrete classification at fine spatial scales, is the potential to characterize the heterogeneity and the patterns associated with fine scale rehabilitation history, such as topsoil distribution, landform contour banks, and seeding patterns, using innovative classification techniques [31]. Moreover, a number of studies have shown the possibility of taking advantage of spectrally unique rehabilitation targets, using high temporal resolution provided by earth-observation satellites [32].
The capacity for remote sensing to convey the complexity of post-mine rehabilitation and the ecological drivers that lead to successful plant establishment and growth, and to mine closure is not comprehensively addressed across the literature. A number of review papers have examined the connection between remote sensing and mining, including fine scale land cover classifications [19], drone monitoring of restoration projects [20,33], and mine impacts [34,35]. Additionally, other review studies have assessed how restoration projects applied ecological attributes using ground monitoring [17,36]. However, there are no review papers that have assessed the role remote sensing has played in the ecological assessment of post-mine rehabilitation. Given the advances in remote sensing technology over the last 50 years and the future trajectories for the areas requiring rehabilitation globally, this paper represents an important and timely contribution.
We undertake a systematic review to assess the current status of remote sensing studies of vegetation dynamics on mine site rehabilitation between 1970 and 2019. We start by characterizing the sensors, platforms, and classification methods used, and the locations and types of mapped resources. We then discuss the key question of how closely remote sensing studies approximated the monitoring of ecological attributes and success of mine site rehabilitation. Using the International Society for Ecological Restoration’s ecological recovery wheel monitoring tool as a template [1], we categorize past studies to assess how researchers approached assessments of rehabilitated landscapes. In particular, what are the ecological attributes that remote sensing measures and where are the research gaps? Are there ecological attributes that remote sensing is not able to measure effectively? We conclude by providing a set of recommendations to guide future research.

2. Materials and Methods

2.1. Search Criteria

We conducted an online search on 23/01/2020 applying a systematic approach to our literature review, using Scopus to search for all papers published between 1970 and 2019. The fifty-year period was chosen to adequately capture the beginning of earth observation through to the current era, and thus provide insight into the progress of rehabilitation assessments from relatively coarse resolution satellites to modern, high spatial and temporal sensors. We used the following search query string:
(TITLE-ABS-KEY (“remote sensing”)) AND (TITLE-ABS-KEY (“mine” OR “mining” OR “mined”)) AND (TITLE-ABS-KEY (“restoration “ OR “reclamation “ OR “rehabilitat*” OR “revegetat*”)) AND (TITLE-ABS-KEY (“vegetat*” OR “eco*” OR “cover” OR “environment*”)) AND NOT (TITLE-ABS-KEY (“data mining”)).
For the purposes of this study, we used the definitions for rehabilitation, restoration, reclamation, and revegetation, derived from the International principles and standards for the practice of ecological restoration [1]. Many of these terms are used interchangeably throughout the literature. Herein, we use the term ‘rehabilitation’ to describe the process of landform reconstruction and reinstatement of a desired level of ecosystem function and vegetation structural development following mining disturbance.
Following the initial online Scopus search, papers were selected for the review based on a four-stage process. First, papers that met the above Scopus search criteria were selected and downloaded if they were relevant to the fields of remote sensing and mine site rehabilitation, published in peer review journals (conference proceedings, books, and non-peer reviewed articles were not included) and published in the English language. Second, papers were shortlisted according to their suitability to the review topic and based on the following criteria. Papers were required to:
  • include remote sensing as a core component of the study;
  • have mine rehabilitation as either a feature of the study or a discrete mapping class (regional or catchment scale studies that generate Land Use Land Cover (LULC) maps were not included);
  • make a novel contribution to the ecology of mine site rehabilitation and successional focus, rather than mapping studies that might have focused more broadly on LULC;
  • demonstrate a primary focus on vegetation rather than other biophysical attributes such as soil or water.
We eliminated a significant number of remote sensing papers that did not directly focus on rehabilitation. For example, many papers assessed the direct impacts of mining on land cover, using change detection at regional scales (e.g., [37,38]), while other studies showed a greater emphasis on soil assessments [39,40,41].
Third, we searched Google Scholar, Web of Science, and ScienceDirect, based on the following word combinations: “remote sensing mine rehabilitation”, “remote sensing mine reclamation”, “remote sensing mine closure”, “remote sensing mine restoration”, and “remote sensing mining land cover” to cross check the Scopus search and determine if any relevant papers were missed. We took the first entries until there were no relevant entries after five pages.
Fourth, during the reading and review process, any new and relevant literature discovered within the reference lists were included in the review.

2.2. Data Compilation

The aims and objectives of each paper were summarized, in addition to the year of publication, Country, region, and location of the study. We compiled a database to summarize a number of variables including commodity, study type, sensors, spatial extent, classification method, temporal scale, and indices used (Table 1).
Mine ‘Commodity’ was recorded for each study to determine if remote sensing studies favored any particular commodities. Copper, zinc, gold, silver, and nickel mines were all grouped into the category ‘Metalliferous’ and small mines extracting hard-rock, calcareous sandstone, bentonite, and perlite were grouped into the category ‘Quarry’. The type of extraction method was also recorded and included open-cut, underground, small-scale artisanal mining, borehole, open pit, strip mining, and quarry. However, this was not included in the final assessment, since the majority of studies were open-cut (86%) and mine type can largely be inferred from commodity. Note that only one study was conducted on an underground coal mine assessing the subsidence impacts on vegetation [42].
Papers were categorized based on the ‘Study Type’ that was undertaken. Ecological studies were those that used remote sensing to gain a deeper understanding of the ecological processes that are occurring in rehabilitated areas. Often these studies are novel and combine remote sensing with ecological field data in an innovative way. This category was further separated into ecological studies that used field data (‘Ecol (Field Obs)’) and ecological studies that did not use field data (‘Ecol (No Field Obs)’). Other categories within the Study Type variable were ‘Land Cover’ assessments, ‘Theoretical’ studies (n = 2) that employed method development, ‘Review’ papers (n = 4) and papers that developed decision support system tools (‘DSS’), generally aimed at policy development (n = 2).
The ‘Sensors’ that were used in each study were recorded and then summarized into broad groups, based on the predominant sensor. These categories included: ‘EO Low’ Earth observation low resolution (>30 m); ‘EO Medium’ Earth observation medium resolution (5–30 m); ‘EO High’ Earth observation high resolution (<5 m); ‘Drone’ for drones using optical sensors (i.e., include various methods such as Structure from Motion (SfM) point clouds to produce Digital Elevation Models (DEMs) and the classification of multispectral imagery) and thermal sensors; ‘Terrestrial Light Detection and Ranging (LiDAR)’, ‘Aerial Optical’ (True-Colour/Multispectral), ‘Aerial Hyperspectral’, ‘Aerial LiDAR’, ‘SAR’ (Synthetic Aperture Radar), and ‘Handheld Hyperspectral’ sensors. Note that we found no studies that used drone LiDAR or drone hyperspectral sensors.
The ‘Spatial Extent’ of each study was categorized into three classes that best reflected the range of studies used in the project. This included the ‘Regional scale’, which were studies that included multiple mine sites (>50 km2), the ‘Mine Site Scale’ that were focused on the assessment of one entire mine site and surrounding areas (10–50 km2), and the ‘Block Scale’, which were studies that focused on 1–2 age-classes of rehabilitation or one rehabilitated block of land (1–10 km2). There were no studies that assessed rehabilitation at the continental or global scale.
‘Classification Method’ was summarized into four broad groups—‘GEOBIA’ for those studies employing Geographic Object-Based Image Analysis (GEOBIA) to segment and classify images based on image-objects rather than pixels; ‘Supervised’ for studies that used a variety of methods including machine learning algorithms such as Random Forests, Support Vector Machines, and boosted classification and regression trees (CART). Other supervised techniques included Maximum Likelihood, k-Nearest Neighbor (k-NN) and Artificial Neural Network (ANN) Classifiers. Studies that used training samples to classify images were also included in the Supervised class. ‘Unsupervised’ methods included Iterative Self-Organizing Data Analysis (ISODATA), self-organization cluster analysis (ISOCLUST—from IDRISI), and ISOSEG (using k-Means). Studies that used methods such as LandTrendr were classed separately as ‘Spectral Time-Series’. The studies that were classed as ‘Manual’ were those that did not use any of the above methods and manually created classes, either by digitizing polygons or using spectral indices to extract features such as vegetation cover, tree canopies, or bare areas.
‘Temporal Scale’ of the studies was summarized by the following classes—‘Uni-Temporal’ for those studies using a single image for the analysis; ‘Bi-Temporal’ when two images were used; ‘Tri-Temporal’ for those studies using 3 images; and ‘Multi-Temporal’ when multiple images were used. These studies spanned 1–2 decades (this can be a time-series [43] or 5+ images within the same decade [44]); and ‘Decadal’ was used to describe studies which used multiple images spanning multiple (>2) decades. Generally, these studies employed time-series analysis and change detection for LULC assessments [45].
The spectral indices (‘Index’) used were summarized for each study. We wanted to know which vegetation index (VI) was the most commonly used and if any particular index was more successfully employed, compared to others. Studies were grouped into ‘Single VI –NDVI’ for those that only used Normalized Differenced Vegetation Index (NDVI), ‘Single VI-Other’ for studies that used a vegetation index apart from NDVI, ‘Multi-VI’ when studies used more than one index (note that all of these included the use of NDVI) and ‘Sensor Bands Only’ for those studies that used spectral bands without applying band ratios.
To determine contribution to ecological outcomes, each study was categorized based on the attributes of the ecological recovery wheel designed for the monitoring of restoration projects by the International Society for Ecological Restoration (SER) [1]. In particular, we used the six attributes and eighteen sub-attributes developed in Principle 3 that are suited for the monitoring of recovering or rehabilitated ecosystems. Each study was categorized according to the most suitable attribute and sub-attribute that best represented the primary objective of the study. The six attributes included—Absence of Threats, Physical Conditions, Species Composition, Structural Diversity, Ecosystem Function, and External Exchanges. Note that of the total 99 papers in the review, those classed as review papers (n = 4) and decision support system tools (DSS) papers (n = 2) did not generate any ecological research outcomes and were, therefore, omitted from the ecological recovery wheel assessment, leaving a total of n = 93 papers. The total number of papers for each category were rescaled to the five-point scale featured in the ecological recovery wheel. This was based on the maximum number of studies for the spatial mosaic sub-attribute, which was the most common study type. In order to further explore the ecological recovery wheel as a remote sensing monitoring tool, we assessed each attribute and sub-attribute and ranked these on the 1–5 scale for the potential for a remote sensing monitoring approach on post-mine rehabilitation. Where remote sensing has a high potential to monitor a sub-attribute, these were ranked a maximum of 5. A moderate and low potential were ranked a 3 and 2, respectively, and scores were ranked zero when a remote sensing approach was unlikely to contribute to monitoring outcomes for mine site rehabilitation.
Finally, we recorded overall map accuracies (OMA %) which were grouped according to the Index used.

3. Results

The online search using Scopus, Google Scholar, Web of Science, and ScienceDirect, resulted in a total of 333 papers; of which 193 research and 4 review papers were assessed according to their suitability. Using our search criteria, we shortlisted a total of 99 papers for the final review comprising 93 research papers, 2 decision support system (DSS) and 4 review articles.
Over time, the number of papers published increased exponentially, with 5 (5%) papers published in the decade 1970–1979, 3 (3%) in 1980–1989, 6 (6%) in 1990–1999, 18 (18%) in 2000–2009, and 67 (68%) published in 2010–2019 (Figure 1). Early papers from the mid-1970s were all studies conducted on rehabilitated coal mines in the United States of America (USA) using Landsat sensors. Overall, the USA accounted for 21 (21%) studies, Australia—17, China—13, Canada—9, Germany—5, and Brazil, Bulgaria, Finland, Greece, India, Indonesia, Italy, Papua New Guinea, Laos, Poland, Russia, South Africa, Spain, and Turkey all had less than 5 studies each (Figure 1A). Remote sensing studies were mostly conducted on coal mines (62%), followed by metalliferous mines (17%), quarries (5%), and bauxite, iron-ore, mineral sand, oil sands, phosphate, sulfur, and uranium (less than 5%) (Figure 1B).
Figure 1C shows that Landsat sensors were the dominant choice of sensor in 50% of studies over the 50-year period. However, over time a greater number of sensors were used, especially between 2010–2019. This was particularly the case in 2017, 2018, and 2019, when drones were used in 25% of studies, aerial hyperspectral in 8%, aerial optical in 4%, and high-resolution earth observation Satellite Pour l’Observation de la Terre (SPOT) in 5%, and WorldView in 3% of studies (Figure 1C).
Interestingly, research conducted in the last decade covered a greater diversity of SER ecological sub-attributes, when compared to the entire 1970–2019 period. While spatial mosaic studies (represented by predominantly LULC studies) dominated over the 1970–2019 period with 63% of studies, the 2010–2019 decade saw an emergence of new studies with a focus across a broader range of sub-attributes. These included remote sensing papers that assessed topics such as productivity/cycling (14%), landscape flows (11%), desirable plants (3%), invasive species (3%), and habitat links (2%) (Figure 1D).
Over the five decades, the types of studies varied, with 36% of research papers attributed to ecological studies with field data, 37% land cover studies, and 18% ecological studies without field data (Figure 2A). In general, studies were dominated by earth observation (EO) low spatial-resolution sensors (55%), with drone-based sensors being the second highest (10%), followed by aerial hyperspectral sensors (8%), EO high (7%), and EO medium (6%) (Figure 2B). The spatial extent for more than half (52%) of the studies was at the mine site scale, with comparatively fewer studies at the regional scale (30%) and the block scale (13%) (Figure 2C). For the temporal scale, the studies were mostly uni-temporal (36%) or decadal (23%), and the classification methods tended to focus on supervised (43%) or manual methods (33%) (Figure 2D,E).
Figure 3A shows the SER ecological recovery wheel categorized for the 93 research papers. (note that of the total 99 papers, 4 review papers and 2 DSS were omitted from this assessment). The total number of papers for each category were rescaled, based on the maximum number of studies for a single sub-attribute, which was 59 (63%) for the sub-attribute spatial mosaic. As a result, all other sub-attributes were rescaled to 1 or below on the ecological recovery wheel scale (Figure 3A). This demonstrated that the literature was dominated by LULC studies, while a smaller proportion focused on ecosystem function (productivity/cycling) (11%) or external exchanges/landscape flows (7%). All other attributes were either under-represented or not represented at all.
It is our opinion that remote sensing has the potential to address 16 of 18 sub-attributes in the ecological recovery wheel, as shown in Figure 3B. Twelve sub-attributes scored a maximum of 5, suggesting that remote sensing is capable of assessing metrics including resilience/recruitment, invasive species, desirable plants, and all vegetation strata. Desirable animals scored 3 indicating a moderate capacity for assessment. All three physical conditions sub-attributes scored 2, suggesting a low capacity, while gene flows and all trophic levels scored zero, suggesting remote sensing is not suitable (Figure 3B).
Figure 4 shows the frequency of spectral indices used as a total of the reviewed publications. Normalized Differenced Vegetation Index (NDVI) was the most common index used, with 44 studies (44%) employing the index. A total of 43 studies (43%) did not employ an index and used sensor bands, most commonly within a supervised classification (47%). The orthogonal indices tessellated cap and principal component analysis (PCA) were used in 9 studies (9%) and 4 studies (4%), respectively; while 7 studies (7%) used soil adjusted vegetation index (SAVI). Six (6) studies used enhanced vegetation index (EVI), normalized moisture difference index (NMDI), and simple ratio (SR), while 5 studies used fractional vegetation cover (Frac Veg). A total of 6 narrow band indices were used with hyperspectral imagery, while 7 indices were used with standard color (RGB) imagery from drone studies.
Of the total studies, 37 quantified overall map accuracy using error matrices, with an average overall accuracy of 84% and a range from 41% to 99% (Table 2). Fourteen publications used NDVI as the only index with an average overall map accuracy of 83% and a range between 52% and 99%. The majority of these studies used GEOBIA classification, along with high-resolution earth observation imagery and did not use field data to validate classifications. Thirteen of the studies used sensor bands with an average overall map accuracy of 85% and a range from 41% to 98%. Comparatively fewer studies used multiple vegetation indices (including NDVI) (6) and single vegetation indices (not NDVI), with an average overall map accuracy of 86% and 74%, respectively (Table 2).
Figure 5 shows an alluvial chart that represents the proportions of data within each variable and the interconnectedness between variables (columns) and categories (rows). The black lines (nodes) represent each category, and the length of the node represents the proportion of publications. For example, the decade 2010–2019 produced the highest number of studies, and most of these publications were produced by the USA, Australia, Canada, and China. A large proportion of the studies produced in the USA were on coal mine rehabilitation, mostly conducted at the site and regional scale. On the right hand side of the alluvial chart, the dominance of the ecological attribute structural diversity and corresponding sub-attribute spatial mosaic is apparent (63%), while important metrics such as resilience/recruitment represent only 1% of studies (Figure 5).

4. Discussion

4.1. Overview

The use of remote sensing to monitor mine site rehabilitation is well established, with studies conducted across 19 different countries on a diverse range of commodity types and using a variety of sensors and platforms. Since the early 1970s, scientists have looked to remote sensing technology to gauge land cover change and quantify rehabilitation outcomes associated with mining impacts [34,46,47,48]. However, the vast body of research for the last five decades shows little evidence that remote sensing has provided practical outputs that capture the key ecological characteristics required for ecological monitoring, and the application of rehabilitation science. Although there was a greater diversity of studies in the 2010–2019 decade, we found that overall, studies paid little attention to the importance of the rehabilitation process, ecosystem development and functioning, and the final goals and objectives of rehabilitation, including mine closure. This is most clearly shown when studies were categorized using the Society for Ecological Restoration’s (SER) ecological recovery wheel monitoring approach [1]. From the total 18 sub-attributes, 63% of studies focused on the spatial mosaic sub-attribute; while comparatively fewer studies addressed more complex topics such as ecosystem function, species composition, and absence of threats. Whereas the underrepresentation of studies on attributes associated with physical conditions was expected (since our review primarily focused on vegetation and actively filtered studies with a focus on substrate), the concentration of studies on spatial mosaic demonstrates the need for a greater understanding of the capacity for remote sensing to monitor rehabilitation success. This involves moving beyond historical approaches for remote sensing, which have commonly been on land-use and land-cover mapping to monitor change.
Interestingly, a review of 68 studies that assessed restoration success through ground monitoring campaigns also found that not all ecological monitoring attributes defined by SER were measured [17]. The most commonly assessed attributes included diversity, vegetation structure and ecological process attributes; while attributes that were poorly measured included those requiring long-time scales such as recruitment, integration with the landscape and self-sustainability. However, in a follow-up review nearly a decade later, Wortley et al. (2013) found that many studies had since addressed those attributes, indicating that the understanding of long-term processes was highly valuable to determining the success of restoration [36]. Coincidently, attributes that require long-term data sets and can indicate the achievement of key goals of mine site rehabilitation such as sustainability (e.g., second-generation recruitment/resilience to disturbance events such as fire) are well placed for remote sensing assessments given spatial and temporal descriptive capabilities. Other attributes that are well placed for remote sensing assessments, and were under represented in the literature include attributes within ecosystem functioning, absence of threats, species composition, external exchanges, and structural diversity. Recent technological advances, coupled with improvements in our understanding of rehabilitation science, suggests that new approaches to remote sensing of rehabilitation are timely, needed, and achievable. Sensors such as hyperspectral and LiDAR have the potential to understand structural and developmental changes that are occurring across rehabilitated areas, while automated processing techniques and software are increasingly available for land managers and stakeholders. Furthermore, historical remote sensing archives can be used to “go back in time”, even if field data is unavailable to fill this important data gap commonly absent in field-based studies [49].

4.2. Chronological Development of Remote Sensing for Mapping Rehabilitation

The choice of sensors for rehabilitation monitoring projects always requires a trade-off between the available spatial and temporal resolution. The potential for remote sensing for quantifying rehabilitation extent and vegetation cover was initially demonstrated by a number of studies using early Landsat Multi Spectral Sensor (MSS) in comparative studies with aerial photography [47,48,50,51]. These pioneering early studies concluded that the new series of satellite earth observation sensors were well suited to land cover mapping, including the inference of vegetation development and rehabilitation vigor [34]. Although now considered a low spatial and temporal resolution earth observation satellite, the Landsat series sensors were used consistently to characterize a range of rehabilitation features including land cover change [52,53,54,55,56,57,58], vegetation health [42,59], species level maps, including invasive species [60,61], vegetation biomass [62], successional development and age class differentiation [22,27,63], carbon sink calculations [64], disturbance and recovery change patterns [65], and surface temperature changes, as a function of vegetation cover and soil moisture properties [66]. Additionally, moderate to high spatial resolution earth observation sensors such as Satellite Pour l’Observation de la Terre (SPOT), IKONOS, Quickbird, RapidEye, GeoEye-1, and WorldView were typically employed in land cover studies, demonstrating the potential of classification methods using single or bi-temporal imagery [31,44,67,68,69,70], while other studies assessed rehabilitation for resilience and recovery from secondary disturbances [29,71].
Beyond the use of passive optical sensors, active sensors demonstrate the capacity for ecological monitoring of rehabilitation but are still relatively under-utilized. Bao et al. (2019) found that fusing WorldView-3 optical bands with Sentinel-1 SAR bands increased the correlation with on-ground measurements of above ground biomass and resulted in a model with a higher prediction accuracy (r2 = 0.79, RMSE = 22.82gm−2) compared to WorldView-3 and Sentinel SAR, separately in a rehabilitated semi-arid grassland in North China [24]. Other studies showed improvement in accuracy by including auxiliary non-spectral light detection and ranging (LiDAR) data. For example, Maxwell et al. (2015) showed that the inclusion of LiDAR had greater classification accuracy than using multispectral data which resulted in confusion between similar land cover classes such as unmined forest versus rehabilitated woodlands [72,73].
More recently, developments in high spatial and temporal resolution using drone-based sensors have contributed significantly to rehabilitation assessments. These studies showed the capacity for measuring vegetation canopies and woody species [74,75], vegetation health [76], impacts from fire disturbance [30], and cover changes over time, along with landform stability to demonstrate rehabilitation success at the block scale [77]. Vegetation structure was also assessed using Structure from Motion (SfM) derived 3D point clouds to measure canopy height and vegetation cover [25]. However, the accuracy of drone-derived products can be significantly impacted by a number of factors, such as season of capture, canopy shadows, camera settings, cloud shadows, and windy conditions that move foliage between captures [25,30,74,75]. Additionally, while drones are becoming increasingly popular, the cost of hardware and image capture, when compared with online free imagery from Sentinel and Landsat, might reduce their popularity in the longer term, and a lack of historic archival imagery is a key limitation.
The assessment of indices used showed the popularity of the Normalized Vegetation Difference Index (NDVI) as a tool for monitoring vegetation development on rehabilitated landscapes. In rehabilitation, as in other remote sensing studies, the NDVI is one of the most widely used indices for vegetation assessment [78]. The NDVI was used in 44% of studies and when employed as the only index (n = 14 studies) produced an average overall map accuracy of 83%. When NDVI was used in combination with one or more indices (n = 6 studies) overall average map accuracy increased marginally to 86% with a range of 45% to 97%. Interestingly when no vegetation indices were used (n = 13 studies), in predominantly supervised classification, overall map accuracy average was a high 85%, suggesting that while indices continue to remain popular, their use in classifying imagery is not essential. The diversity of indices used in papers assessing rehabilitation was notable, with 44 different spectral, orthogonal, SAR, and thermal indices used over the fifty-year period. Spectral indices derived from multispectral imagery (26 indices) were most common, followed by visible true color imagery (red green blue) (7 indices) and hyperspectral imagery (7 indices) (Table A1).

4.3. Recommendations for Operationalising Remote Sensing for Mine Site Rehabilitation

A number of studies demonstrated the increased scrutiny on rehabilitation outcomes following legislative changes, due to increased incentives for mining companies to report progressive rehabilitation and demonstrate compliance to Government agencies. In the United States, the introduction of the Surface Mine Reclamation Act (SMCRA) of 1977 led to an initial surge in remote sensing studies developing methods to inventory rehabilitation efforts [22,47,48,50,51,79,80]. Similar legislation in other countries led to comparable studies in Canada [81,82], Poland [83], South Africa [58], Italy [84], Spain [85], Greece [53], China [86], and Australia [65,74,87,88].
Due to its popularity and availability, scientists from a broad range of disciplines are increasingly using remote sensing as a tool for environmental monitoring. However, remote sensing assessments of mine site rehabilitation require considerable knowledge and skill to interpret outputs of analysis and derive meaningful results. This is highlighted by a number of studies emphasizing the heterogeneous nature of mine site rehabilitation and the unique spatial and temporal challenges for remote sensing assessments [21,22,23,24,25,26,27,28,29,30]. Alternatively, the distinctive spectral, temporal and spatial patterns of rehabilitated landscapes provides opportunities for remote sensing to derive valuable ecological information. For example, progressive rehabilitation over a long period can result in a patchwork of different age-classes that potentially can be assessed with a single image to build change trajectories of ecological metrics such as vegetation cover, woody density, and species richness. Such assessments might be possible, but only with field-based knowledge such as establishment age, site preparation methods, seeding mixes, and reflectance values for endemic and exotic species. One key limitation we found was that less than half of the studies (43%) integrated ground-based field data with remote sensing data. Ground data are essential for calibrating and validating remotely sensed models and for testing correlations between ecological attributes such as vegetation cover. Another common limitation we identified was the assumption that increases in vegetation cover, identified by mapping vegetation cover, are equivalent to a final measure of success for mine site rehabilitation. While a measured increase in vegetation cover might indicate stability and show early stages of plant establishment, this represents just one metric in a suite of criteria that is needed to inform rehabilitation success. Importantly, this assumption ignores the value of desired plant species, absence of weeds, and the structural components of developing ecosystems that are essential to meet target community attributes required for end land-use and mine closure.
Several knowledge gaps can be addressed given the current state of remote sensing technology and image availability. Based on our knowledge of rehabilitation and remote sensing science, we make the following suggestions for future work:
  • The monitoring of ecosystem function in the remote sensing literature is scant and is a key metric for rehabilitation success. In particular, the SER sub-attribute of ecological resilience requires further research. Given that one of the key aims for mine site rehabilitation is an ecosystem that is self-sustainable, the measurement and demonstration of resilience is one key area for future work [12,29]. This is particularly salient given that disturbances such as fire, drought, disease, floods, and storms are inevitable and predicted to increase, given climate change.
  • Comparing baselines and reference sites to rehabilitated sites is a common approach used when monitoring for restoration success [1], and is often a requirement of ecological monitoring programs. However, the comparison between mined and unmined ecosystems was only addressed by a few notable exceptions [44,89]. Comparisons between unmined natural ecosystems versus rehabilitation will provide stakeholders with confidence in rehabilitation success and is important to support mine closure, and can be directly addressed using remote sensing.
  • Studies showing long-term (decadal) development and achievement of ecological processes, such as structural canopy development [65], are required to confirm that a post-mining ecosystem is self-sustainable and resilient. A single snapshot or multiple snapshots in time provide little evidence of how rehabilitation is responding to long-term environmental fluctuations and changes over time. For example, many studies used NDVI as a proxy for vegetation vigor and rehabilitation success, however, this index is highly responsive to seasonal fluctuations. A time-series of NDVI provides more insight into rehabilitation performance rather than a uni-temporal assessment.
  • Large-scale studies using long-term ground monitoring data integrated with remote sensing metrics to assess ecosystem development should be explored, given the availability of comprehensive ground monitoring data-sets that are often required by regulators to demonstrate rehabilitation progression.
  • Regional or continental scale inventories of rehabilitation estates can provide industry-wide (and company-wide) perspectives on completed and still-to-be-completed rehabilitation projects.
  • There are many examples of remote sensing approaches for measuring vegetation health, which could be used for rehabilitation monitoring that have been demonstrated in other disciplines, such as forestry [90].
  • Field-based monitoring is still the most common approach used globally for assessing rehabilitation success, yet it is time-consuming, labor intensive, and expensive. In order for remote sensing to be further operationalized by mining companies and government regulators, researchers need to demonstrate which field-based metrics can be confidently derived using remote sensing techniques.
Throughout 50 years of remote sensing monitoring, there is yet to be a clear scientific consensus on the best approach for determining rehabilitation success through remote sensing methods. There are great opportunities for moving beyond the conventional land cover assessments and developing standardized approaches to measure a number of ecological attributes as defined by SER International standards. These opportunities could arise through the development of tools to assist rehabilitation practitioners and regulators, to measure short-term achievement of success criteria, or methods to quantify long-term success, including resilience to climate change and stochastic events.

5. Conclusions

Our review of the available literature indicates that over the past five decades, the number of ecologically focused remote sensing studies conducted on mine site rehabilitation increased gradually from the early 1970s to reach a peak in the 2010–2019 decade. Initial research was dominated by the Landsat series of satellites, through broad-scale land cover assessments, while the introduction of new earth observation and drone sensors enabled studies to explore more complex ecological metrics. However, the SER ecological monitoring attributes remain significantly underrepresented, and future research should focus on demonstrating the potential to measure sub-attributes associated with ecosystem function, absence of threats, species composition, and structural diversity. The average overall mapping accuracy for 37 studies that conducted error assessments was 84% and the NDVI was the most common spectral index, used in 44% of studies, with an average overall map accuracy of 83% when used as the solitary index. It is clear that the discipline is in the early stages of growth and there are significant opportunities for the assessment of progressive rehabilitation and achievement of end land-use targets. Given the comparatively recent rapid advances in sensor technology, processing software, and spatial analysis techniques, along with changing legal and social environments, the importance of developing standardized monitoring methods for gauging rehabilitation success is evident. Indeed, improvements in the characterization of ecological recovery through remotely sensed methods will lead to a number of positive outcomes; from increased efficacy of on-ground management, to increased stakeholder confidence throughout the progressive rehabilitation period, and at the time of mine closure.

Author Contributions

Conceptualization, P.B.M., A.M.L., P.D.E., and S.P.; methodology, P.B.M., A.M.L., and P.D.E.; software, P.B.M.; validation, A.M.L. and P.D.E.; formal analysis, P.B.M.; investigation, P.B.M.; resources, P.B.M.; data curation, P.B.M.; writing—original draft preparation, P.B.M., writing—review and editing, A.M.L., P.D.E.; visualization, P.B.M.; supervision, A.M.L., P.D.E.; and S.P., project administration, P.B.M.; funding acquisition, P.B.M. All authors have read and agreed to the published version of the manuscript.

Funding

Phillip McKenna is paid a scholarship by the Queensland Resources Council.

Acknowledgments

David Doley for proof reading the manuscript and the anonymous reviewers for greatly improving the paper. Danny McKenna for assistance with graphic design.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Summary of indices used in the review. Hyper = Hyperspectral, Multi = Multispectral, SAR = Synthetic Aperture Radar, VIS = Visible. * Note that many studies used multiple indices.
Table A1. Summary of indices used in the review. Hyper = Hyperspectral, Multi = Multispectral, SAR = Synthetic Aperture Radar, VIS = Visible. * Note that many studies used multiple indices.
Imagery TypeIndexDescriptionFormulaNo. Studies *% of Total Studies
HyperHyper (CARC)Carotenoid ContentPSRI = (R680 − R500)/R75011.0
HyperHyper (CC1)Chlorophyll ContentR695/R67011.0
HyperHyper (CC2)Chlorophyll Content 2(R850-R710)/(R850-R680)11.0
HyperHyper (LAC)Leaf Anthocyanin ContentARTI1 = (1/R700) − (1/R550)11.0
HyperHyper (NDGI)Normalised Differenced Green IndexNDGI = (G − R)/(G + R)11.0
HyperHyper (Slope)SlopeSlope = (NIR1125−NIR935)/(NIR1125−NIR935)11.0
HyperHyper (SU)Spectral Unmixing Endmember-33.0
MultiARVIAtmospherically Resistent Vegetation IndexARVI = (NIR − (R − (B − R)))/(NIR + (R − (B − R)))11.0
MultiCVIChlorophyll Vegetation IndexNIR×R/G211.0
MultiDVIDifference Veg IndexDVI = NIR − R22.0
MultiEVIEnhanced Vegetation IndexEVI = G × (NIR − R)/(NIR + C1 ×R − C2 ×B + L)66.1
MultiEVI2Enhanced Vegetation Index 2EVI2 = 2.5 × NIR-R/NIR + 2.4 ×R + L11.0
MultiFrac VegFractional Vegetation CoverFVC = (NDVI − NDVImin)/(NDVImax − NDVImin)55.1
MultiMSAVIModified Soil Adjusted Vegetation IndexMSAVI = (NIR − R)/(NIR + R+L) × (1 + L)11.0
MultiMSAVI2Modified Soil Adjusted Vegetation Index 2MSAVI2 = (2 ×NIR + 1-SQRT((2 × NIR + 1)2 − 8 × (NIR-R)) ×0.522.0
MultiNBR1Normalised Burn Ratio 1NBR = (NIR-SWIR2)/(NIR + SWIR2) (note SWIR2 =2.11-2.29µm on Landsat 8)44.0
MultiNBR2Normalised Burn Ratio 2NBR2 = (SWIR1 − SWIR2)/(SWIR1 + SWIR2) (note SWIR1 =1.57−1.65 µm SWIR2 = 2.11−2.29 µm on Landsat 8)11.0
MultiNDVINormalised Differenced Vegetation IndexNDVI = (NIR − R)/(NIR + R)4444.4
MultiNDVI(B)Normalised Differenced Vegetation Index (Blue)BNVI = (NIR − B)/(NIR + B)11.0
MultiNDVI(G)Normalised Differenced Vegetation Index (Green)GNVI = (NIR − G)/(NIR + G)22.0
MultiNDVI(RE)Normalised Differenced Vegetation Index (Red Edge)NDVI(RE) = (NIR − RE)/(NIR + RE)11.0
MultiNDWI1Normalised Differenced Wetness Index 1NDWI1 = (G − NIR)/(G + NIR)11.0
MultiNMDINormalised Moisture Differenced IndexNMDI = (NIR-SWIR1)/(NIR + SWIR1) (note SWIR1 =1.57−1.65 µm on L8)66.1
MultiPVIPerpendicular Vegetation IndexPVI = NIR × sinx − R × cosx11.0
MultiRatio (Unspecified)Vegetation RatioUnspecified (suspected NDVI)11.0
MultiRIRegrowth indexRI = NDVI (inner patch) − NDVI (outer patch)11.0
MultiRSRReduced Simple RatioRSR = SR×1 − SWIR − SWIRmin × SWIRmax − SWIRmin22.0
MultiRVIRatio Vegetation IndexRVI = R/NIR33.0
MultiSAVISoil Adjusted Vegetation IndexSAVI = (NIR-R)/(NIR + R) +L × (1 + L)77.1
MultiSRSimple RatioNIR/R66.1
MultiSR2Simple Ratio 2NIR/SWIR11.0
MultiTSAVITransformed Soil Adjusted Vegetation IndexTSAVI = s × (NIR-s × R-a)/(a × NIR + R-a × s + X ×(1 + s2)22.0
MultiWDVIWeighted difference veg IndexWDVI = NIR-s × R11.0
No_IndexSensor Bands Only-4343.4
OrthagonalPCAPrincipal Components Analysis-44.0
OrthagonalTass CapTasselated Cap-99.1
SARSAR (SNR)Signal to Noise Ratio (not a spectral index)SNR = M/LSD11.0
ThermalLELatent Energy Heat FluxLE = Rn × (0.114 + 0.78 ×EVI + 0.004 × LST)22.0
VISRBG (GRI)Green Ratio indexGRI = NIR/G11.0
VISRGB (EGI)Excess Green indexEGI = 2 × G-R-B11.0
VISRGB (EGIR)Excess Green Index RatioEGIR = R/2 × G × 100011.0
VISRGB (MEGI)Modified Excess Green IndexMEGI = 2 × G-R11.0
VISRGB (TGI)Triangular Greenness IndexTGI = -0.5[(670-480)(R-G)-(670-550)(R-B)11.0
VISRGB (VARI)Visible Atmospheric Resistant IndexVARI = (G-R)/(G+R-B)11.0
VISRGB (VI)Vegetation IndexVI = (2 × G-R-B) − (1.4 × R-G)11.0

Appendix B

Table A2. A summary of the 99 papers used in the study, sorted by year of publication.
Table A2. A summary of the 99 papers used in the study, sorted by year of publication.
No.AuthorRefYearCountryCommodity SummarySensor SummaryCraft PrimarySpatial ExtentTemporal ScaleClassification TypeField ObsStudy Type 1Study Type 2 (Attributes)Study Type 2 (Sub-Attributes)
1Wobber etal[47]1975USACoalEO LowLandsatRegionalUni-TemporalManualYLand CoverStructural DiversitySpatial Mosaic
2Anderson & Schubert[91]1976USACoalEO LowLandsatSite ScaleUni-TemporalSupervisedYLand CoverStructural DiversitySpatial Mosaic
3Anderson etal[50]1977USACoalEO LowLandsatRegionalTri-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
4Mamula[48]1978USACoalEO LowLandsatRegionalUni-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
5Brumbaugh[51]1979USACoalEO LowLandsatRegionalUni-TemporalManualNLand CoverStructural DiversitySpatial Mosaic
6Game etal[92]1982USACoalAerial OpticalAerial_OpticalSite ScaleMulti-TemporalSupervisedYEcol (Field Obs)Structural DiversitySpatial Mosaic
7Irons & Kennard[93]1986USACoalEO LowLandsatRegionalUni-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
8Parks etal[22]1987USACoalEO LowLandsatRegionalBi-TemporalSupervisedYLand CoverStructural DiversitySpatial Mosaic
9Phinn etal[27]1991AustraliaMineral SandEO LowLandsatSite ScaleUni-TemporalSupervisedYEcol (Field Obs)Structural DiversitySpatial Mosaic
10Hill & Phinn[63]1993AustraliaMineral SandEO LowLandsatSite ScaleUni-TemporalSupervisedYEcol (Field Obs)Species CompositionDesirable Animals
11Rathore & Wright[34]1993NANANANANANANANAReviewNANA
12Felinks etal[26]1998GermanyCoalEO LowLandsatSite ScaleBi-TemporalManualYEcol (Field Obs)Structural DiversitySpatial Mosaic
13Schmid etal[21]1998GermanyCoalEO LowLandsatSite ScaleMulti-TemporalSupervisedYEcol (Field Obs)Structural DiversitySpatial Mosaic
14Staenz etal[94]1999CanadaMetalliferousAerial HyperAerial_HyperspecSite ScaleUni-TemporalSupervisedYEcol (Field Obs)Structural DiversitySpatial Mosaic
15Lévesque etal[95]2000CanadaMetalliferousAerial HyperAerial_HyperspecSite ScaleUni-TemporalSupervisedNEcol (No Field Obs)Structural DiversitySpatial Mosaic
16Almeida-filho[45]2002BrazilMetalliferousEO LowLandsatSite ScaleDecadalUnsupervisedNLand CoverStructural DiversitySpatial Mosaic
17Bonifazi etal[84]2003ItalyQuarryEO LowLandsatRegionalUni-TemporalManualYEcol (Field Obs)Structural DiversitySpatial Mosaic
18Cutaia etal[96]2004ItalyQuarryEO LowLandsatRegionalTri-TemporalManualYEcol (Field Obs)Structural DiversitySpatial Mosaic
19Ganas etal[53]2004GreeceMetalliferousEO LowLandsatSite ScaleTri-TemporalSupervisedNDSSNANA
20Pfitzner etal[97]2006AustraliaUraniumField HyperField_HyperspecNANANANTheoreticalSpecies CompositionDesirable Plants
21Trisasongko etal[98]2006IndonesiaMetalliferousSARSARRegionalUni-TemporalUnsupervisedNLand CoverStructural DiversitySpatial Mosaic
22Weiersbye etal[99]2006South AfricaMetalliferousAerial HyperAerial_HyperspecBlock ScaleUni-TemporalSupervisedYEcol (Field Obs)Absence of ThreatsContamination
23Antwi etal[52]2008GermanyCoalEO LowLandsatSite ScaleBi-TemporalManualNEcol (No Field Obs)Structural DiversitySpatial Mosaic
24Gillanders etal[55]2008CanadaOil SandsEO LowLandsatRegionalDecadalUnsupervisedNLand CoverStructural DiversitySpatial Mosaic
25Halounová[54]2008Czech RepublicCoalEO LowLandsatRegionalTri-TemporalManualNEcol (No Field Obs)Ecosystem FunctionProductivity/Cycling
26Lau etal[100]2008AustraliaBauxiteAerial HyperAerial_HyperspecSite ScaleUni-TemporalNAYEcol (Field Obs)Physical ConditionsWater Chemo-Physical
27Lévesque & Staenz[82]2008CanadaMetalliferousAerial HyperAerial_HyperspecSite ScaleTri-TemporalUnsupervisedYEcol (Field Obs)Structural DiversitySpatial Mosaic
28Richter etal[101]2008CanadaMetalliferousAerial HyperAerial_HyperspecSite ScaleUni-TemporalSupervisedYLand CoverStructural DiversitySpatial Mosaic
29Yang[102]2008USACoalEO LowLandsatSite ScaleTri-TemporalSupervisedYLand CoverStructural DiversitySpatial Mosaic
30Lu etal[103]2009ChinaCoalEO LowEO-1_HyperionSite ScaleUni-TemporalNAYEcol (Field Obs)Ecosystem FunctionProductivity/Cycling
31Townsend etal[79]2009USACoalEO LowLandsatRegionalDecadalUnsupervisedNLand CoverStructural DiversitySpatial Mosaic
32Wei etal[104]2009ChinaCoalEO LowLandsatSite ScaleDecadalManualNLand CoverStructural DiversitySpatial Mosaic
33Sonwalkar etal[105]2010USAMetalliferousEO LowMODISSite ScaleMulti-TemporalManualNLand CoverStructural DiversitySpatial Mosaic
34Spyropoulos etal[106]2010GreeceMetalliferousEO LowLandsatSite ScaleTri-TemporalSupervisedNDSSNANA
35Demirel etal[67]2011TurkeyCoalEO HighIKONOS_&_QuickbirdSite ScaleBi-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
36Demirel etal[68]2011TurkeyCoalEO HighIKONOS_&_QuickbirdSite ScaleBi-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
37Erener[59]2011TurkeyCoalEO LowLandsatRegionalTri-TemporalManualNEcol (No Field Obs)Ecosystem FunctionProductivity/Cycling
38Procházka etal[107]2011Czech RepublicCoalEO LowLandsatSite ScaleUni-TemporalUnsupervisedYEcol (Field Obs)External ExchangesLandscape Flows
39Sun etal[108]2011ChinaCoalField HyperField_HyperspecBlock ScaleUni-TemporalSupervisedYEcol (Field Obs)Species CompositionDesirable Plants
40Bao etal[69]2012ChinaCoalEO HighQuickbirdSite ScaleUni-TemporalManualNEcol (No Field Obs)Species CompositionDesirable Plants
41Bodlak etal[28]2012Czech RepublicCoalEO LowLandsatSite ScaleBi-TemporalManualYEcol (Field Obs)External ExchangesLandscape Flows
42Brom etal[66]2012Czech RepublicCoalEO LowLandsatSite ScaleDecadalManualNEcol (No Field Obs)External ExchangesLandscape Flows
43Sen etal[32]2012USACoalEO LowLandsatRegionalDecadalSupervisedNEcol (No Field Obs)Ecosystem FunctionProductivity/Cycling
44Fletcher & Erskine[74]2013AustraliaCoalDroneDroneBlock ScaleUni-TemporalManualNEcol (No Field Obs)Structural DiversityAll vegetation strata
45Lemke etal[60]2013USACoalEO LowLandsatRegionalDecadalManualYEcol (Field Obs)Absence of ThreatsInvasive Species
46Oparin etal[62]2013RussiaCoalEO LowLandsatSite ScaleBi-TemporalManualNLand CoverStructural DiversitySpatial Mosaic
47Petropoulos etal[56]2013GreeceQuarryEO LowLandsatRegionalTri-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
48Raval etal[70]2013AustraliaCoalEO HighWorldViewSite ScaleUni-TemporalManualNEcol (No Field Obs)Structural DiversitySpatial Mosaic
49Antwi etal[109]2014GermanyCoalEO LowLandsatRegionalDecadalSupervisedYEcol (Field Obs)Structural DiversitySpatial Mosaic
50Bao etal[31]2014AustraliaMetalliferousEO MedSPOTSite ScaleUni-TemporalGEOBIAYEcol (Field Obs)Structural DiversitySpatial Mosaic
51Bao etal[44]2014AustraliaMetalliferousEO MedSPOTSite ScaleMulti-TemporalManualNEcol (No Field Obs)Ecosystem FunctionProductivity/Cycling
52Maxwell etal[23]2014USACoalAerial LiDARAerial_LiDARSite ScaleUni-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
53Maxwell etal[110]2014USACoalAerial OpticalAerial_OpticalSite ScaleUni-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
54Wang etal[64]2014ChinaCoalEO LowLandsatRegionalDecadalGEOBIAYEcol (Field Obs)Ecosystem FunctionProductivity/Cycling
55Zhang etal[111]2014CanadaOil SandsEO MedSPOTSite ScaleMulti-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
56Badreldin & Sanchez-Azofeifa[112]2015CanadaCoalTerrest LiDARTerrest LiDARSite ScaleDecadalManualYEcol (Field Obs)Ecosystem FunctionProductivity/Cycling
57Li etal[113]2015USACoalEO LowLandsatRegionalDecadalSupervisedNLand CoverStructural DiversitySpatial Mosaic
58Li etal[114]2015USACoalEO LowLandsatRegionalDecadalSupervisedNLand CoverStructural DiversitySpatial Mosaic
59Maxwell & Warner[72]2015USACoalAerial OpticalAerial_OpticalRegionalUni-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
60Maxwell etal[73]2015USACoalEO MedRapidEyeSite ScaleUni-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
61Szostak etal[83]2015PolandSulfurEO LowLandsatSite ScaleBi-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
62Tong etal[115]2015ChinaPhosphateDroneDroneSite ScaleUni-TemporalGEOBIANLand CoverStructural DiversitySpatial Mosaic
63Bao etal[116]2016ChinaCoalEO HighWorldViewSite ScaleUni-TemporalGEOBIANEcol (No Field Obs)Structural DiversitySpatial Mosaic
64Götze etal[117]2016Czech RepublicCoalAerial HyperAerial_HyperspecSite ScaleUni-TemporalSupervisedYEcol (Field Obs)Physical ConditionsSubstrate Chemical
65Karan etal[118]2016IndiaCoalEO LowLandsatSite ScaleBi-TemporalSupervisedYLand CoverStructural DiversitySpatial Mosaic
66Lechner, Kassulke & Unger[87]2016AustraliaCoalAerial OpticalAerial_OpticalRegionalUni-TemporalManualNLand CoverStructural DiversitySpatial Mosaic
67Liu[119]2016ChinaCoalEO LowLandsatSite ScaleDecadalManualNLand CoverStructural DiversitySpatial Mosaic
68Chen etal[43]2017ChinaCoalEO LowMODISRegionalMulti-TemporalManualYEcol (Field Obs)External ExchangesLandscape Flows
69Esposito[120]2017ItalyQuarryDroneDroneBlock ScaleBi-TemporalNANLand CoverStructural DiversitySpatial Mosaic
70LeClerc & Wiersma[81]2017CanadaMetalliferousEO LowLandsatRegionalDecadalSupervisedYEcol (Field Obs)External ExchangesHabitat links
71Macfarlane etal[121]2017AustraliaBauxiteEO LowLandsatRegionalDecadalManualYTheoreticalStructural DiversitySpatial Mosaic
72McKenna etal[30]2017AustraliaCoalDroneDroneBlock ScaleBi-TemporalSupervisedNEcol (No Field Obs)External ExchangesLandscape Flows
73Oliphant etal[61]2017USACoalEO LowLandsatSite ScaleMulti-TemporalSupervisedYEcol (Field Obs)Absence of ThreatsInvasive Species
74Padmanaban etal[42]2017GermanyCoalEO LowLandsatSite ScaleMulti-TemporalSupervisedNLand CoverStructural DiversitySpatial Mosaic
75Yang etal[71]2017ChinaCoalEO HighGeoEye-1Site ScaleUni-TemporalManualYEcol (Field Obs)Ecosystem FunctionHabitat & Interactions
76Zenkov etal[57]2017BulgariaCoalEO LowLandsatSite ScaleDecadalManualNLand CoverStructural DiversitySpatial Mosaic
77Bujalsky etal[122]2018Czech RepublicCoalAerial HyperAerial_HyperspecSite ScaleUni-TemporalManualNEcol (No Field Obs)External ExchangesLandscape Flows
78Chasmer etal[123]2018CanadaOil SandsEO MedSPOTSite ScaleMulti-TemporalManualYEcol (Field Obs)Ecosystem FunctionProductivity/Cycling
79Chen etal[19]2018NANANANANANANANAReviewNANA
80Correa etal[124]2018BrazilMetalliferousEO LowMODISSite ScaleMulti-TemporalNAYEcol (Field Obs)Ecosystem FunctionProductivity/Cycling
81Gastauer etal[125]2018NANANANANANANANAReviewNANA
82McKenna etal[29]2018AustraliaCoalEO HighWorldViewBlock ScaleTri-TemporalSupervisedYEcol (Field Obs)Ecosystem FunctionResilience/recruitment
83Whiteside & Bartolo[75]2018AustraliaUraniumDroneDroneBlock ScaleTri-TemporalGEOBIANEcol (No Field Obs)Structural DiversitySpatial Mosaic
84Yang etal[126]2018USACoalEO LowLandsatRegionalDecadalManualNEcol (No Field Obs)Ecosystem FunctionHabitat & Interactions
85Yang etal[65]2018AustraliaCoalEO LowLandsatRegionalDecadalSpectral Time-SeriesYEcol (Field Obs)Ecosystem FunctionProductivity/Cycling
86Zenkov etal[127]2018RussiaIron OreEO LowLandsatSite ScaleBi-TemporalManualNLand CoverStructural DiversitySpatial Mosaic
87Bao etal[24]2019ChinaCoalSARSARBlock ScaleUni-TemporalManualYEcol (Field Obs)Ecosystem FunctionProductivity/Cycling
88Buters etal[20]2019NANANANANANANANAReviewNANA
89Dlamini etal[58]2019South AfricaMineral SandEO LowLandsatSite ScaleDecadalSpectral Time-SeriesNEcol (No Field Obs)Structural DiversitySpatial Mosaic
90Erskine etal[89]2019AustraliaUraniumEO LowLandsatSite ScaleDecadalNAYEcol (Field Obs)External ExchangesLandscape Flows
91Isokangas etal[76]2019FinlandMetalliferousDroneDroneBlock ScaleUni-TemporalManualYEcol (Field Obs)Absence of ThreatsContamination
92Johansen, Erskine & McCabe[77]2019AustraliaCoalDroneDroneBlock ScaleTri-TemporalGEOBIAYEcol (Field Obs)Structural DiversitySpatial Mosaic
93Kun[128]2019TurkeyCoalDroneDroneBlock ScaleUni-TemporalManualYEcol (Field Obs)Structural DiversitySpatial Mosaic
94Lechner etal[49]2019PNG_&_LaosMetalliferousEO LowLandsatRegionalDecadalSupervisedNLand CoverStructural DiversitySpatial Mosaic
95Moudry etal[25]2019Czech RepublicCoalDroneDroneBlock ScaleUni-TemporalManualYEcol (Field Obs)Structural DiversityAll vegetation strata
96Padro etal[85]2019SpainQuarryDroneDroneBlock ScaleUni-TemporalSupervisedYEcol (Field Obs)Structural DiversitySpatial Mosaic
97Vasuki etal[129]2019AustraliaBauxiteEO LowLandsatRegionalDecadalSupervisedNLand CoverStructural DiversitySpatial Mosaic
98Xu etal[130]2019ChinaCoalEO MedSPOTRegionalTri-TemporalSupervisedNEcol (No Field Obs)Structural DiversitySpatial Mosaic
99Zhang etal[86]2019ChinaCoalEO LowLandsatSite ScaleDecadalSupervisedNEcol (No Field Obs)Structural DiversitySpatial Mosaic

References

  1. Gann, G.D.; McDonald, T.; Walder, B.; Aronson, J.; Nelson, C.R.; Jonson, J.; Hallett, J.G.; Eisenberg, C.; Guariguata, M.R.; Liu, J.; et al. International principles and standards for the practice of ecological restoration. Second edition. Restor. Ecol. 2019, 27, S1–S46. [Google Scholar] [CrossRef] [Green Version]
  2. Hobbs, R.J.; Norton, D.A. Towards a conceptual framework for Restoration Ecology. Restor. Ecol. 1996, 4, 93–110. [Google Scholar] [CrossRef]
  3. Doley, D.; Audet, P. Adopting novel ecosystems as suitable rehabilitation alternatives for former mine sites. Ecol. Process. 2013, 2, 1–11. [Google Scholar] [CrossRef] [Green Version]
  4. US Government. Surface Mining Control and Reclamation Act of 1977, Title 30, Chapter 25. In Title 30, Chapter 25, Office of Surface Mining and Reclamation. Washington, United States of America; 1977. Available online: https://www.govinfo.gov/content/pkg/STATUTE-91/pdf/STATUTE-91-Pg445.pdf (accessed on 27 October 2020).
  5. Skousen, J.; Zipper, C.E. Post-mining policies and practices in the Eastern USA coal region. Int. J. Coal Sci. Technol. 2014, 1, 135–151. [Google Scholar] [CrossRef] [Green Version]
  6. Queensland Government. Mineral and Energy Resources (Financial Provisioning) Bill 2018; Department of Environment and Science: Brisbane, Australia, 2018; p. 204. Available online: https://www.parliament.qld.gov.au/Documents/TableOffice/TabledPapers/2018/5618T173.pdf (accessed on 1 February 2020).
  7. Western Australia Government. Mining Rehabilitation Fund Act 2012; Department of Mines, Industry Regulation and Safety: Perth, Australia, 2014; p. 31. Available online: www.legislation.wa.gov.au (accessed on 1 October 2020).
  8. Province of Alberta. Environmental Protection and Enhancement Act. Revised Statutes of Alberta 2000 Chapter E-12; Alberta, Canada. 2020, p. 168. Available online: https://www.qp.alberta.ca/ (accessed on 1 October 2020).
  9. People’s Republic of China. Mineral Resources Law; Ministry of Commerce: Beijing, China, 2009. Available online: http://www.china.org.cn/environment/2007-08/20/content_1034342.htm (accessed on 1 October 2020).
  10. Republic of South Africa. Mineral and Petroleum Resources Development Act; South African Government: Cape Town, Republic of South Africa, 2002; p. 122. Available online: https://www.gov.za/documents/ (accessed on 1 October 2020).
  11. The European Parliament. Directive 2006/21/EC of the European Parliament and of the Council of 15 March 2006 on the Management of Waste from Extractive Industries and Amending Directive 2004/35/EC. Strasbourg, The European Parliament. 2006. Available online: https://eur-lex.europa.eu/eli/dir/2006/21/2009-08-07 (accessed on 1 October 2020).
  12. Erskine, P.D.; Fletcher, A.T. Novel Ecosystems created by coal mines in Central Queensland’s Bowen Basin. Ecol. Process. 2013, 2, 33. [Google Scholar] [CrossRef] [Green Version]
  13. Gould, S.F. Comparison of Post-mining Rehabilitation with Reference Ecosystems in Monsoonal Eucalypt Woodlands, Northern Australia. Restor. Ecol. 2012, 20, 250–259. [Google Scholar] [CrossRef]
  14. Vickers, H.; Gillespie, M.; Gravina, A. Assessing the development of rehabilitated grasslands on post-mined landforms in north west Queensland, Australia. Agric. Ecosyst. Environ. 2012, 163, 72–84. [Google Scholar] [CrossRef]
  15. Audet, P.; Gravina, A.; Glenn, V.; McKenna, P.B.; Vickers, H.; Gillespie, M.; Mulligan, D. Structural development of vegetation on rehabilitated North Stradbroke Island: Above/belowground feedback may facilitate alternative ecological outcomes. Ecol. Process. 2013, 2, 20. [Google Scholar] [CrossRef] [Green Version]
  16. Gravina, A.A.; McKenna, P.B.; Glenn, V. Evaluating the success of mineral sand mine rehabilitation on North Stradbroke Island, Queensland: Comparisons with reference eucalypt communities. Proc. R. Soc. Queensl. 2011, 117, 419–436. [Google Scholar]
  17. Ruiz-Jaen, M.C.; Aide, T.M. Restoration success: How is it being measured? Restor. Ecol. 2005, 13, 569–577. [Google Scholar] [CrossRef]
  18. Doley, D.; Audet, P.; Mulligan, D.R. Examining the Australian context for post-mined land rehabilitation: Reconciling a paradigm for the development of natural and novel ecosystems among post-disturbance landscapes. Agric. Ecosyst. Environ. 2012, 163, 85–93. [Google Scholar] [CrossRef]
  19. Chen, W.; Li, X.; He, H.; Wang, L. A review of fine-scale land use and land cover classification in open-pit mining areas by remote sensing techniques. Remote Sens. 2018, 10, 15. [Google Scholar] [CrossRef] [Green Version]
  20. Buters, T.M.; Bateman, P.W.; Robinson, T.; Belton, D.; Dixon, K.; Cross, A.T. Methodological Ambiguity and Inconsistency Constrain Unmanned Aerial Vehicles as A Silver Bullet for Monitoring Ecological Restoration. Remote Sens. 2019, 11, 1180. [Google Scholar] [CrossRef] [Green Version]
  21. Schmidt, H.; Glaesser, C. Multitemporal analysis of satellite data and their use in the monitoring of the environmental impacts of open cast lignite mining areas in eastern Germany. Int. J. Remote Sens. 1998, 19, 2245–2260. [Google Scholar] [CrossRef]
  22. Parks, N.F.; Petersen, G.W.; Baumer, G.M. High resolution remote sensing of spatially and spectrally complex coal surface mines of central Pennsylvania: A comparison between simulated SPOT MSS and Landsat-5 Thematic Mapper. Photogramm. Eng. Remote Sens. 1987, 53, 415–420. [Google Scholar]
  23. Maxwell, A.E.; Warner, T.A.; Strager, M.P.; Pal, M. Combining RapidEye Satellite Imagery and Lidar for Mapping of Mining and Mine Reclamation. Photogramm. Eng. Remote Sens. 2014, 80, 179–189. [Google Scholar] [CrossRef]
  24. Bao, N.; Li, W.; Gu, X.; Liu, Y. Biomass estimation for semiarid vegetation and mine rehabilitation using worldview-3 and sentinel-1 SAR imagery. Remote Sens. 2019, 11, 2855. [Google Scholar] [CrossRef] [Green Version]
  25. Moudrý, V.; Gdulová, K.; Fogl, M.; Klápště, P.; Urban, R.; Komárek, J.; Moudrá, L.; Štroner, M.; Barták, V.; Solský, M. Comparison of leaf-off and leaf-on combined UAV imagery and airborne LiDAR for assessment of a post-mining site terrain and vegetation structure: Prospects for monitoring hazards and restoration success. Appl. Geogr. 2019, 104, 32–41. [Google Scholar] [CrossRef]
  26. Felinks, B.; Pilarski, M.; Wiegleb, G. Vegetation survey in the former brown coal mining area of eastern Germany by integrating remote sensing and ground-based methods. Appl. Veg. Sci. 1998, 1, 233–240. [Google Scholar] [CrossRef]
  27. Phinn, S.R.; Hill, G.J.E.; Johnston, S.; Kirwood, C. Landsat Thematic Mapper Imagery for Mapping Disturbed and Rehabilitated Vegetation, North Stradbroke Island, Queensland. Queensl. Geogr. J. 1991, 6, 47–58. [Google Scholar]
  28. Bodlák, L.; Křováková, K.; Nedbal, V.; Pechar, L. Assessment of landscape functionality changes as one aspect of reclamation quality—The case of Velká podkrušnohorská dump, Czech Republic. Ecol. Eng. 2012, 43, 19–25. [Google Scholar] [CrossRef]
  29. McKenna, P.B.; Phinn, S.; Erskine, P. Fire Severity and Vegetation Recovery on Mine Site Rehabilitation Using WorldView-3 Imagery. Fire 2018, 1, 22. [Google Scholar] [CrossRef] [Green Version]
  30. McKenna, P.B.; Erskine, P.D.; Lechner, A.M.; Phinn, S. Measuring fire severity using UAV imagery in semi-arid central Queensland, Australia. Int. J. Remote Sens. 2017, 38, 4244–4264. [Google Scholar] [CrossRef]
  31. Bao, N.; Lechner, A.M.; Johansen, K.; Ye, B. Object-based classification of semi-arid vegetation to support mine rehabilitation and monitoring. J. Appl. Remote Sens. 2014, 8, 1–19. [Google Scholar] [CrossRef]
  32. Sen, S.; Zipper, C.E.; Wynne, R.H.; Donovan, P.F. Identifying Revegetated Mines as Disturbance/Recovery Trajectories Using an Interannual Landsat Chronosequence. Photogramm. Eng. Remote Sens. 2012, 78, 223–235. [Google Scholar] [CrossRef]
  33. Park, S. Applications of Unmanned Aerial Vehicles in Mining from Exploration to Reclamation: A Review. Minerals 2020, 10, 663. [Google Scholar] [CrossRef]
  34. Rathore, C.S.; Wright, R. Monitoring environmental impacts of surface coal mining. Int. J. Remote Sens. 1993, 14, 1021–1042. [Google Scholar] [CrossRef]
  35. Werner, T.T.; Bebbington, A.; Gregory, G. Assessing impacts of mining: Recent contributions from GIS and remote sensing. Extr. Ind. Soc. 2019, 6, 993–1012. [Google Scholar] [CrossRef]
  36. Wortley, L.; Hero, J.M.; Howes, M. Evaluating ecological restoration success: A review of the literature. Restor. Ecol. 2013, 21, 537–543. [Google Scholar] [CrossRef]
  37. McIntyre, N.; Bulovic, N.; Cane, I.; McKenna, P.B. A multi-disciplinary approach to understanding the impacts of mines on traditional uses of water in Northern Mongolia. Sci. Total Environ. 2016, 557, 404–414. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, Z.; He, G.; Wang, M.; Wang, Z.; Long, T.; Peng, Y. Detecting Decadal Land Cover Changes in Mining Regions based on Satellite Remotely Sensed Imagery: A Case Study of the Stone Mining Area in Luoyuan County, SE China. Photogramm. Eng. Remote Sens. 2015, 81, 745–751. [Google Scholar] [CrossRef]
  39. Bao, N.; Wu, L.; Ye, B.; Yang, K.; Zhou, W. Assessing soil organic matter of reclaimed soil from a large surface coal mine using a field spectroradiometer in laboratory. Geoderma 2017, 288, 47–55. [Google Scholar] [CrossRef]
  40. Bollhöfer, A.; Pfitzner, K.; Ryan, B.; Martin, P.; Fawcett, M.; Jones, D.R. Airborne gamma survey of the historic Sleisbeck mine area in the Northern Territory, Australia, and its use for site rehabilitation planning. J. Environ. Radioact. 2008, 99, 1770–1774. [Google Scholar] [CrossRef] [PubMed]
  41. Zhang, S.; Shen, Q.; Nie, C.; Huang, Y.; Wang, J.; Hu, Q.; Ding, X.; Zhou, Y.; Chen, Y. Hyperspectral inversion of heavy metal content in reclaimed soil from a mining wasteland based on different spectral transformation and modeling methods. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 211, 393–400. [Google Scholar] [CrossRef]
  42. Padmanaban, R.; Bhowmik, A.; Cabral, P. A Remote Sensing Approach to Environmental Monitoring in a Reclaimed Mine Area. ISPRS Int. J. Geo-Inf. 2017, 6, 401. [Google Scholar] [CrossRef] [Green Version]
  43. Chen, G.; Wang, M.; Liu, Z.; Chi, W. The biogeophysical effects of revegetation around mining areas: A case study of Dongsheng mining areas in Inner Mongolia. Sustainability 2017, 9, 628. [Google Scholar] [CrossRef] [Green Version]
  44. Bao, N.; Lechner, A.M.; Fletcher, A.; Erskine, P.; Mulligan, D.; Bai, Z. SPOTing long-term changes in vegetation over short-term variability. Int. J. Min. Reclam. Environ. 2014, 28, 2–24. [Google Scholar] [CrossRef]
  45. Almeida-filho, R.; Shimabukuro, Y.E. Digital processing of a Landsat-TM time series for mapping and monitoring degraded areas caused by independent gold miners, Roraima State, Brazilian Amazon. Remote Sens. Environ. 2002, 79, 42–50. [Google Scholar] [CrossRef]
  46. Alexander, S.S.; Dein, J.; Gold, D.P. The use of ERTS-1 MSS data for mapping strip mines and acid mine drainage in Pennsyvania. NASA Spec. Publ. 1973, 372, 569–575. [Google Scholar]
  47. Wobber, F.J.; Russell, O.R.; Deely, D.J. Multiscale aerial and orbital techniques for management of coal-mined lands. Photogrammetria 1975, 31, 117–133. [Google Scholar] [CrossRef]
  48. Mamula, N. Remote-sensing methods for monitoring surface coal mining in the northern great plains. J. Res. U.S. Geol. Surv. 1978, 6, 149–160. [Google Scholar]
  49. Lechner, A.M.; Owen, J.; Ang, M.L.E.; Edraki, M.; Che Awang, N.A.; Kemp, D. Historical socio-environmental assessment of resource development footprints using remote sensing. Remote Sens. Appl. Soc. Environ. 2019, 15, 1–12. [Google Scholar] [CrossRef]
  50. Anderson, A.T.; Schultz, D.; Buchman, N.; Nock, H.M. Landsat imagery for surface-mine inventory. A band-ratio method for measuring disturbed surface areas proved to be both temporally and geographically extendable. Photogramm. Eng. Remote Sens. 1977, 43, 1027–1036. [Google Scholar]
  51. Brumbaugh, F.R. Strip mine reclamation and remote sensing technology. Min. Congr. J. 1979, 65, 57–61. [Google Scholar]
  52. Antwi, E.K.; Krawczynski, R.; Wiegleb, G. Detecting the effect of disturbance on habitat diversity and land cover change in a post-mining area using GIS. Landsc. Urban Plan. 2008, 87, 22–32. [Google Scholar] [CrossRef]
  53. Ganas, A.; Aerts, J.; Astaras, T.; De Vente, J.; Frogoudakis, E.; Lambrinos, N.; Riskakis, C.; Oikonomidis, D.; Filippidis, A.; Kassoli-Fournaraki, A. The use of earth observation and decision support systems in the restoration of open-cast nickel mines Evia, central Greece. Int. J. Remote Sens. 2004, 25, 3261–3274. [Google Scholar] [CrossRef]
  54. Halounová, L. Reclamation areas and their development studied by vegetation indices. Int. J. Digit. Earth 2008, 1, 155–164. [Google Scholar] [CrossRef]
  55. Gillanders, S.N.; Coops, N.C.; Wulder, M.A.; Goodwin, N.R. Application of Landsat satellite imagery to monitor land-cover changes at the Athabasca Oil Sands, Alberta, Canada. Can. Geogr. 2008, 52, 466–485. [Google Scholar] [CrossRef]
  56. Petropoulos, G.P.; Partsinevelos, P.; Mitraka, Z. Change detection of surface mining activity and reclamation based on a machine learning approach of multi-temporal Landsat TM imagery. Geocarto Int. 2013, 28, 323–342. [Google Scholar] [CrossRef]
  57. Zenkov, V.; Yuronen, Y.P.; Nefedov, B.N.; Zayats, V.V. Remote monitoring of ecological state of disturbed lands in the area of trojanovo open pit coal mine in Bulgaria. Eurasian Min. 2017, 2017, 38–41. [Google Scholar] [CrossRef]
  58. Dlamini, L.Z.D.; Xulu, S. Monitoring mining disturbance and restoration over RBM site in South Africa using landtrendr algorithm and landsat data. Sustainability 2019, 11, 6916. [Google Scholar] [CrossRef] [Green Version]
  59. Erener, A. Remote sensing of vegetation health for reclaimed areas of Seyitömer open cast coal mine. Int. J. Coal Geol. 2011, 86, 20–26. [Google Scholar] [CrossRef]
  60. Lemke, D.; Schweitzer, C.J.; Tadesse, W.; Wang, Y.; Brown, J.A. Geospatial Assessment of Invasive Plants on Reclaimed Mines in Alabama. Invasive Plant Sci. Manag. 2013, 6, 401–410. [Google Scholar] [CrossRef]
  61. Oliphant, A.J.; Wynne, R.H.; Zipper, C.E.; Ford, W.M.; Donovan, P.F.; Li, J. Autumn olive (Elaeagnus umbellata) presence and proliferation on former surface coal mines in Eastern USA. Biol. Invasions 2017, 19, 179–195. [Google Scholar] [CrossRef]
  62. Oparin, V.N.; Potapov, V.P.; Giniyatullina, O.L.; Shchastlivtsev, E.L. Studies into the process of mine waste dump filling up by vegetation using remote sensing data. J. Min. Sci. 2013, 49, 976–982. [Google Scholar] [CrossRef]
  63. Hill, G.J.E.; Phinn, S.R. Revegetated sand mining areas, swamp wallabies and remote sensing: North Stradbroke Island, Queensland. Aust. Geogr. Stud. 1993, 31, 3–13. [Google Scholar] [CrossRef]
  64. Wang, J.; Jiao, Z.; Bai, Z. Changes in carbon sink value based on RS and GIS in the Heidaigou opencast coal mine. Environ. Earth Sci. 2014, 71, 863–871. [Google Scholar] [CrossRef]
  65. Yang, Y.; Erskine, P.D.; Lechner, A.M.; Mulligan, D.; Zhang, S.; Wang, Z. Detecting the dynamics of vegetation disturbance and recovery in surface mining area via Landsat imagery and LandTrendr algorithm. J. Clean. Prod. 2018, 178, 353–362. [Google Scholar] [CrossRef]
  66. Brom, J.; Nedbal, V.; Procházka, J.; Pecharová, E. Changes in vegetation cover, moisture properties and surface temperature of a brown coal dump from 1984 to 2009 using satellite data analysis. Ecol. Eng. 2012, 43, 45–52. [Google Scholar] [CrossRef]
  67. Demirel, N.; Emil, M.K.; Duzgun, H.S. Surface coal mine area monitoring using multi-temporal high-resolution satellite imagery. Int. J. Coal Geol. 2011, 86, 3–11. [Google Scholar] [CrossRef]
  68. Demirel, N.; Düzgün, Ş.; Emil, M.K. Landuse change detection in a surface coal mine area using multi-temporal high-resolution satellite images. Int. J. Min. Reclam. Environ. 2011, 25, 342–349. [Google Scholar] [CrossRef]
  69. Bao, N.; Ye, B.; Bai, Z. Rehabilitation of Vegetation Mapping of ATB Opencast Coal-Mine Based on GIS and RS. Sens. Lett. 2012, 10, 387–393. [Google Scholar] [CrossRef]
  70. Raval, S.; Merton, R.N.; Laurence, D. Satellite based mine rehabilitation monitoring using WorldView-2 imagery. Min. Technol. 2013, 122, 200–207. [Google Scholar] [CrossRef]
  71. Yang, Y.; Ren, X.; Zhang, S.; Chen, F.; Hou, H. Incorporating ecological vulnerability assessment into rehabilitation planning for a post-mining area. Environ. Earth Sci. 2017, 76, 1–16. [Google Scholar] [CrossRef]
  72. Maxwell, A.E.; Warner, T.A. Differentiating mine-reclaimed grasslands from spectrally similar land cover using terrain variables and object-based machine learning classification. Int. J. Remote Sens. 2015, 36, 4384–4410. [Google Scholar] [CrossRef] [Green Version]
  73. Maxwell, A.E.; Warner, T.A.; Strager, M.P.; Conley, J.F.; Sharp, A.L. Assessing machine-learning algorithms and image- and lidar-derived variables for GEOBIA classification of mining and mine reclamation. Int. J. Remote Sens. 2015, 36, 954–978. [Google Scholar] [CrossRef]
  74. Fletcher, A.T.; Erskine, P.D. Rehabilitation closure criteria assessment using high resolution photogrammetrically derived surface models. In Proceedings of the International Archives of the Photogrammetry, Remote Sensing and Spatial Information Sciences, Rostock, Germany, 4–6 September 2013; pp. 1–4. [Google Scholar]
  75. Whiteside, T.G.; Bartolo, R.E. A robust object-based woody cover extraction technique for monitoring mine site revegetation at scale in the monsoonal tropics using multispectral RPAS imagery from different sensors. Int. J. Appl. Earth Obs. Geoinf. 2018, 73, 300–312. [Google Scholar] [CrossRef]
  76. Isokangas, E.; Davids, C.; Kujala, K.; Rauhala, A.; Ronkanen, A.K.; Rossi, P.M. Combining unmanned aerial vehicle-based remote sensing and stable water isotope analysis to monitor treatment peatlands of mining areas. Ecol. Eng. 2019, 133, 137–147. [Google Scholar] [CrossRef]
  77. Johansen, K.; Erskine, P.D.; McCabe, M.F. Using Unmanned Aerial Vehicles to assess the rehabilitation performance of open cut coal mines. J. Clean. Prod. 2019, 209, 819–833. [Google Scholar] [CrossRef]
  78. Huang, S.; Tang, L.; Hupy, J.P.; Wang, Y.; Shao, G. A commentary review on the use of normalized difference vegetation index (NDVI) in the era of popular remote sensing. J. For. Res. 2020, 1–6. [Google Scholar]
  79. Townsend, P.A.; Helmers, D.P.; Kingdon, C.C.; McNeil, B.E.; de Beurs, K.M.; Eshleman, K.N. Changes in the extent of surface mining and reclamation in the Central Appalachians detected using a 1976–2006 Landsat time series. Remote Sens. Environ. 2009, 113, 62–72. [Google Scholar] [CrossRef]
  80. Carr, J.R.; Glass, C.E.; Schowengerdt, R.A. Signature extension versus retraining for multispectral classification of surface mines in arid regions (Arizona). Photogramm. Eng. Remote Sens. 1983, 49, 1193–1199. [Google Scholar]
  81. LeClerc, E.; Wiersma, Y.F. Assessing post-industrial land cover change at the Pine Point Mine, NWT, Canada using multi-temporal Landsat analysis and landscape metrics. Environ. Monit. Assess. 2017, 189, 185. [Google Scholar] [CrossRef] [PubMed]
  82. Lévesque, J.; Staenz, K. Monitoring mine tailings revegetation using multitemporal hyperspectral image data. Can. J. Remote Sens. 2008, 34, S172–S186. [Google Scholar] [CrossRef]
  83. Szostak, M.; Wężyk, P.; Hawryło, P.; Pietrzykowski, M. The analysis of spatial and temporal changes of land cover and land use in the reclaimed areas with the application of airborne orthophotomaps and LANDSAT images. Geod. Cartogr. 2015, 64, 75–86. [Google Scholar] [CrossRef] [Green Version]
  84. Bonifazi, G.; Cutaia, L.; Massacci, P.; Roselli, I. Monitoring of abandoned quarries by remote sensing and in situ surveying. Ecol. Modell. 2003, 170, 213–218. [Google Scholar] [CrossRef]
  85. Padró, J.C.; Carabassa, V.; Balagué, J.; Brotons, L.; Alcañiz, J.M.; Pons, X. Monitoring opencast mine restorations using Unmanned Aerial System (UAS) imagery. Sci. Total Environ. 2019, 657, 1602–1614. [Google Scholar] [CrossRef]
  86. Zhang, M.; Wang, J.; Li, S. Tempo-spatial changes and main anthropogenic influence factors of vegetation fractional coverage in a large-scale opencast coal mine area from 1992 to 2015. J. Clean. Prod. 2019, 232, 940–952. [Google Scholar] [CrossRef]
  87. Lechner, A.M.; Kassulke, O.; Unger, C. Spatial assessment of open cut coal mining progressive rehabilitation to support the monitoring of rehabilitation liabilities. Resour. Policy 2016, 50, 234–243. [Google Scholar] [CrossRef]
  88. Unger, C.; Lechner, A.M.; Glenn, V.; Edraki, M.; Mulligan, D. Mapping and prioritising rehabilitation of abandoned mines in Australia. Proc. Life Mine 2012, 259–266. [Google Scholar]
  89. Erskine, P.D.; Bartolo, R.; McKenna, P.B.; Humphrey, C. Using reference sites to guide ecological engineering and restoration of an internationally significant uranium mine in the Northern Territory, Australia. Ecol. Eng. 2019, 129, 61–70. [Google Scholar] [CrossRef]
  90. Lechner, A.M.; Foody, G.M.; Boyd, D.S. Applications in Remote Sensing to Forest Ecology and Management. One Earth 2020, 2, 405–412. [Google Scholar] [CrossRef]
  91. Anderson, A.T.; Schubert, J. ERTS-1 Data Applied to Strip Mining. Photogramm. Eng. Remote Sens. 1976, 42, 211–219. [Google Scholar]
  92. Game, M.; Carrel, J.E.; Hotrabhavandra, T. Patch Dynamics of Plant Succession on Abandoned Surface Coal Mines: A Case History Approach. Br. Ecol. Soc. 1982, 70, 707–720. [Google Scholar] [CrossRef]
  93. Irons, J.R.; Kennard, R.L. The utility of Thematic Mapper sensor characteristics for surface mine monitoring. Photogramm. Eng. Remote Sens. 1986, 52, 389–396. [Google Scholar]
  94. Staenz, K.; Neville, R.A.; Lévesque, J.; Szeredi, T.; Singhroy, V.; Borstad, G.A.; Hauff, P. Evaluation of casi and SFSI hyperspectral data for environmental and geological applications - Two case studies. Can. J. Remote Sens. 1999, 25, 311–322. [Google Scholar] [CrossRef]
  95. Lévesque, J.; Staenz, K.; Szeredi, T. The impact of spectral band characteristics on unmixing of hyperspectral data for monitoring mine tailings site rehabilitation. Can. J. Remote Sens. 2000, 26, 231–240. [Google Scholar] [CrossRef]
  96. Cutaia, L.; Massacci, P.; Roselli, I. Analysis of Landsat 5 TM images for monitoring the state of restoration of abandoned quarries. Int. J. Surf. Mining, Reclam. Environ. 2004, 18, 122–134. [Google Scholar] [CrossRef]
  97. Pfitzner, K.; Bollhöfer, A.; Carr, G. A standard design for collecting vegetation reference spectra: Implementation and implications for data sharing. J. Spat. Sci. 2006, 51, 79–92. [Google Scholar] [CrossRef]
  98. Trisasongko, B.; Lees, B.; Paull, D. Polarimetric classification in a tailings deposition area at the Timika Mine Site, Indonesia. Mine Water Environ. 2006, 25, 246–250. [Google Scholar] [CrossRef]
  99. Weiersbye, I.M.; Margalit, N.; Feingersh, T.; Revivo, G.; Stark, R.; Zur, Y.; Heller, D.; Braun, O.; Cukrowska, E.M. Use of airborne hyper-spectral remote sensing (HSRS) to focus remediation and monitor vegetation processes on gold mining landscapes in South Africa. Mine Clos. 2006, 2006, 601–611. [Google Scholar]
  100. Lau, I.; Hewson, R.; Ong, C.; Tongway, D. Remote Mine Site Rehabilitation Monitoring Using Airborne Hyperspectral Imaging and Landscape Function Analysis (LFA). In Proceedings of the International Archives of the Photogrammetry, Remote Sensing and Spatial Information Sciences, Beijing, China, 3–11 July 2008; Volume 37, pp. 325–330. [Google Scholar]
  101. Richter, N.; Staenz, K.; Kaufmann, H. Spectral unmixing of airborne hyperspectral data for baseline mapping of mine tailings areas. Int. J. Remote Sens. 2008, 29, 3937–3956. [Google Scholar] [CrossRef]
  102. Yang, J.J. Detecting landscape changes pre-and post surface coal mining in Indiana, USA. Geogr. Inf. Sci. 2008, 14, 36–43. [Google Scholar] [CrossRef]
  103. Lu, X.; Hu, Z.Q.; Guo, L. Quantitative inverse modeling of nitrogen content from hyperion data under stress of exhausted coal mining sites. Min. Sci. Technol. 2009, 19, 31–35. [Google Scholar] [CrossRef]
  104. Wei, S.; Zhong-Ping, S.; Dao-Liang, L.; Raaj, R.; Xiang, Z.; Xiang-Yun, G.; Su, W.; Sun, Z.-P.; Li, D.-L.; Ramsankaran, R.; et al. Vegetation recovery monitoring over the waste dump in Haizhou opencast coalmine area, China, during 1975-2000 using NDVI and VF index. J. Indian Soc. Remote Sens. 2009, 37, 631–645. [Google Scholar]
  105. Sonwalkar, M.; Fang, L.; Sun, D. Use of NDVI dataset for a GIS based analysis: A sample study of TAR Creek superfund site. Ecol. Inform. 2010, 5, 484–491. [Google Scholar] [CrossRef]
  106. Spyropoulos, N.; Valmis, S.; Ganas, A.; Metaxas, G. Land reclamation of surface mining based on imagery intelligence and spatial decision support systems. World Trans. Eng. Technol. Educ. 2010, 8, 115–119. [Google Scholar]
  107. Procházka, J.; Brom, J.; Št’astný, J.; Pecharová, E. The impact of vegetation cover on temperature and humidity properties in the reclaimed area of a brown coal dump. Int. J. Min. Reclam. Environ. 2011, 25, 350–366. [Google Scholar] [CrossRef]
  108. Sun, H.; Li, M.; Li, D. The vegetation classification in coal mine overburden dump using canopy spectral reflectance. Comput. Electron. Agric. 2011, 75, 176–180. [Google Scholar] [CrossRef]
  109. Antwi, E.K.; Boakye-Danquah, J.; Asabere, S.B.; Takeuchi, K.; Wiegleb, G. Land cover transformation in two post-mining landscapes subjected to different ages of reclamation since dumping of spoils. SpringerPlus 2014, 3, 1–22. [Google Scholar] [CrossRef] [Green Version]
  110. Maxwell, A.E.; Strager, M.P.; Warner, T.A.; Zégre, N.P.; Yuill, C.B. Comparison of NAIP orthophotography and rapideye satellite imagery for mapping of mining and mine reclamation. GIScience Remote Sens. 2014, 51, 301–320. [Google Scholar] [CrossRef]
  111. Zhang, Y.; Guindon, B.; Lantz, N.; Shipman, T.; Chao, D.; Raymond, D. Quantification of anthropogenic and natural changes in oil sands mining infrastructure land based on RapidEye and SPOT5. Int. J. Appl. Earth Obs. Geoinf. 2014, 29, 31–43. [Google Scholar] [CrossRef]
  112. Badreldin, N.; Sanchez-Azofeifa, A. Estimating forest biomass dynamics by integrating multi-temporal Landsat satellite images with ground and airborne LiDAR data in the Coal Valley Mine, Alberta, Canada. Remote Sens. 2015, 7, 2832–2849. [Google Scholar] [CrossRef] [Green Version]
  113. Li, J.; Xia, Q.; Zipper, C.E.; Li, S.; Donovan, P.F.; Wynne, R.H.; Oliphant, A.J. Character analysis of mining disturbance and reclamation trajectory in surface coal-mine area by time-series NDVI. Trans. Chin. Soc. Agric. Eng. 2015, 31, 251–257. [Google Scholar]
  114. Li, J.; Zipper, C.E.; Donovan, P.F.; Wynne, R.H.; Oliphant, A.J. Reconstructing disturbance history for an intensively mined region by time-series analysis of Landsat imagery. Environ. Monit. Assess. 2015, 187, 557. [Google Scholar] [CrossRef]
  115. Tong, X.; Liu, X.; Chen, P.; Liu, S.; Luan, K.; Li, L.; Liu, S.; Liu, X.; Xie, H.; Jin, Y.; et al. Integration of UAV-based photogrammetry and terrestrial laser scanning for the three-dimensional mapping and monitoring of open-pit mine areas. Remote Sens. 2015, 7, 6635–6662. [Google Scholar] [CrossRef] [Green Version]
  116. Bao, N.; Wu, L.X.; Liu, S.J.; Li, N. Scale parameter optimization through high-resolution imagery to support mine rehabilitated vegetation classification. Ecol. Eng. 2016, 97, 130–137. [Google Scholar] [CrossRef]
  117. Götze, C.; Beyer, F.; Gläßer, C. Pioneer vegetation as an indicator of the geochemical parameters in abandoned mine sites using hyperspectral airborne data. Environ. Earth Sci. 2016, 75, 1–14. [Google Scholar] [CrossRef]
  118. Karan, S.K.; Samadder, S.R.; Maiti, S.K. Assessment of the capability of remote sensing and GIS techniques for monitoring reclamation success in coal mine degraded lands. J. Environ. Manag. 2016, 182, 272–283. [Google Scholar] [CrossRef]
  119. Liu, X.; Zhou, W.; Bai, Z. Vegetation coverage change and stability in large open-pit coal mine dumps in China during 1990–2015. Ecol. Eng. 2016, 95, 447–451. [Google Scholar] [CrossRef]
  120. Esposito, G.; Mastrorocco, G.; Salvini, R.; Oliveti, M.; Starita, P. Application of UAV photogrammetry for the multi-temporal estimation of surface extent and volumetric excavation in the Sa Pigada Bianca open-pit mine, Sardinia, Italy. Environ. Earth Sci. 2017, 76, 1–16. [Google Scholar] [CrossRef]
  121. Macfarlane, C.; Grigg, A.H.; Daws, M.I. A standardised Landsat time series (1973–2016) of forest leaf area index using pseudoinvariant features and spectral vegetation index isolines and a catchment hydrology application. Remote Sens. Appl. Soc. Environ. 2017, 6, 1–14. [Google Scholar] [CrossRef]
  122. Bujalský, L.; Jirka, V.; Zemek, F.; Frouz, J. Relationships between the normalised difference vegetation index and temperature fluctuations in post-mining sites. Int. J. Min. Reclam. Environ. 2018, 32, 254–263. [Google Scholar] [CrossRef]
  123. Chasmer, L.; Baker, T.; Carey, S.K.; Straker, J.; Strilesky, S.; Petrone, R. Monitoring ecosystem reclamation recovery using optical remote sensing: Comparison with field measurements and eddy covariance. Sci. Total Environ. 2018, 642, 436–446. [Google Scholar] [CrossRef] [PubMed]
  124. Corrêa, R.S.; Balduíno, A.P.D.C.; Teza, C.T.V.; Baptista, G.M.D.M. Vegetation cover development resulting from different restoration approaches of exploited mines. Floresta Ambient. 2018, 25, 1–9. [Google Scholar] [CrossRef]
  125. Gastauer, M.; Silva, J.R.; Caldeira Junior, C.F.; Ramos, S.J.; Souza Filho, P.W.M.; Furtini Neto, A.E.; Siqueira, J.O. Mine land rehabilitation: Modern ecological approaches for more sustainable mining. J. Clean. Prod. 2018, 172, 1409–1422. [Google Scholar] [CrossRef]
  126. Yang, Z.; Li, J.; Zipper, C.E.; Shen, Y.; Miao, H.; Donovan, P.F. Identification of the disturbance and trajectory types in mining areas using multitemporal remote sensing images. Sci. Total Environ. 2018, 644, 916–927. [Google Scholar] [CrossRef] [PubMed]
  127. Zenkov, I.V.; Vokin, V.N.; Kiryushina, E.V.; Raevich, K.V. Remote monitoring data on opencast mining and disturbed land ecology in the bakal iron ore field. Eurasian Min. 2018, 2018, 29–33. [Google Scholar] [CrossRef]
  128. Kun, M. Assessment and monitoring of rehabilitation studies on coal mine dump site with UAV’S. Appl. Ecol. Environ. Res. 2019, 17, 7381–7393. [Google Scholar] [CrossRef]
  129. Vasuki, Y.; Yu, L.; Holden, E.J.; Kovesi, P.; Wedge, D.; Grigg, A.H. The spatial-temporal patterns of land cover changes due to mining activities in the Darling Range, Western Australia: A Visual Analytics Approach. Ore Geol. Rev. 2019, 108, 23–32. [Google Scholar] [CrossRef]
  130. Xu, J.; Zhao, H.; Yin, P.; Wu, L.; Li, G. Landscape ecological quality assessment and its dynamic change in coal mining area: A case study of Peixian. Environ. Earth Sci. 2019, 78, 1–13. [Google Scholar] [CrossRef]
Figure 1. Stacked bar charts showing the number of publications for each year, according to the country of study (A), the commodity (B), the primary craft used for the assessment (C), and the sub-attributes taken from the SER ecological recovery wheel (D).
Figure 1. Stacked bar charts showing the number of publications for each year, according to the country of study (A), the commodity (B), the primary craft used for the assessment (C), and the sub-attributes taken from the SER ecological recovery wheel (D).
Remotesensing 12 03535 g001
Figure 2. Summary statistics for the 99 publications reviewed showing the type of study (A), type of sensor (B), spatial extent of study (C), temporal scale (D), and classification methods employed (E).
Figure 2. Summary statistics for the 99 publications reviewed showing the type of study (A), type of sensor (B), spatial extent of study (C), temporal scale (D), and classification methods employed (E).
Remotesensing 12 03535 g002
Figure 3. Remote sensing studies plotted on the ecological recovery wheel for monitoring projects (International Society for Ecological Restoration (SER)) based on their primary objective and rescaled by the sub-attribute spatial mosaic which had the highest number of studies at 59 (A). Potential application of remote sensing opportunities for monitoring in mine site rehabilitation (B). This figure was adapted from the ecological recovery wheel, accessed at http://seraustralasia.com/wheel/wheel.html.
Figure 3. Remote sensing studies plotted on the ecological recovery wheel for monitoring projects (International Society for Ecological Restoration (SER)) based on their primary objective and rescaled by the sub-attribute spatial mosaic which had the highest number of studies at 59 (A). Potential application of remote sensing opportunities for monitoring in mine site rehabilitation (B). This figure was adapted from the ecological recovery wheel, accessed at http://seraustralasia.com/wheel/wheel.html.
Remotesensing 12 03535 g003
Figure 4. The frequency of indices used. Note that a number of studies used multiple indices within one publication. Hyper = Hyperspectral imagery narrow band indices, and RGB = derived from color red green blue imagery. For a list of index names see Table A1.
Figure 4. The frequency of indices used. Note that a number of studies used multiple indices within one publication. Hyper = Hyperspectral imagery narrow band indices, and RGB = derived from color red green blue imagery. For a list of index names see Table A1.
Remotesensing 12 03535 g004
Figure 5. Alluvial chart demonstrating the proportions within variables and the correlations between variables. From Left to Right: Decade of publication, Country of study, Commodity, Sensor used in study, Spatial Extent, Temporal Scale, Classification method, ecological attribute, and ecological sub-attribute. NA = review papers, DSS, and theoretical papers.
Figure 5. Alluvial chart demonstrating the proportions within variables and the correlations between variables. From Left to Right: Decade of publication, Country of study, Commodity, Sensor used in study, Spatial Extent, Temporal Scale, Classification method, ecological attribute, and ecological sub-attribute. NA = review papers, DSS, and theoretical papers.
Remotesensing 12 03535 g005
Table 1. Metadata used to categorize each study in the review. Each study was given a class according to the variables below. * Study Type was also classified according to the International Society for Ecological Restoration (SER) ecological recovery wheel attributes (not listed here).
Table 1. Metadata used to categorize each study in the review. Each study was given a class according to the variables below. * Study Type was also classified according to the International Society for Ecological Restoration (SER) ecological recovery wheel attributes (not listed here).
Variables
CommodityStudy Type *SensorsSpatial ExtentClassification MethodTemporal ScaleIndex
Categories/Class
-
Coal
-
Metalliferous
-
Bauxite
-
Iron-Ore
-
Mineral Sand
-
Oil Sands
-
Phosphate
-
Uranium
-
Quarry
-
Sulphur
-
Ecological (with Field Observations)
-
Ecological (without Field Observations)
-
Land Cover
-
Review
-
Theoretical
-
Decision Support System
-
EO Low (>30 m)
-
EO Medium (5–30 m)
-
EO High (<5 m)
-
Drone Optical (RGB/MS/Thermal)
-
Terrestrial LiDAR
-
Aerial Optical (RBG/MS)
-
Aerial Hyperspectral sensors
-
Aerial LiDAR
-
SAR
-
Handheld Hyperspectral sensors
-
Regional scale (multiple Mine sites, > 50 km2)
-
Mine site scale (1 Mine site, 10–50 km2)
-
Block scale (1–2 age-classes of rehabilitation, 1–10 km2)
-
GEOBIA
-
Supervised
-
Unsupervised
-
Spectral Time Series
-
Manual
-
Uni-Temporal
-
Bi-Temporal
-
Tri-Temporal
-
Multi-Temporal
-
Decadal
-
Single VI-NDVI
-
Single VI-Other
-
Multi VI
-
Sensor Bands Only
Table 2. Summary of overall map accuracies indices taken from studies in the review that quantified overall map accuracies.
Table 2. Summary of overall map accuracies indices taken from studies in the review that quantified overall map accuracies.
Index Categoryn (Values)n (Publications)Average (%)Min (%)Max (%)SD
Multi VI (NDVI+)19686459712.9
Sensor Bands Only971385419812.3
Single VI NDVI331483529912.0
Single VI Other9474529115.0
Total/Overall1583784419913.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

McKenna, P.B.; Lechner, A.M.; Phinn, S.; Erskine, P.D. Remote Sensing of Mine Site Rehabilitation for Ecological Outcomes: A Global Systematic Review. Remote Sens. 2020, 12, 3535. https://doi.org/10.3390/rs12213535

AMA Style

McKenna PB, Lechner AM, Phinn S, Erskine PD. Remote Sensing of Mine Site Rehabilitation for Ecological Outcomes: A Global Systematic Review. Remote Sensing. 2020; 12(21):3535. https://doi.org/10.3390/rs12213535

Chicago/Turabian Style

McKenna, Phillip B., Alex M. Lechner, Stuart Phinn, and Peter D. Erskine. 2020. "Remote Sensing of Mine Site Rehabilitation for Ecological Outcomes: A Global Systematic Review" Remote Sensing 12, no. 21: 3535. https://doi.org/10.3390/rs12213535

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