The International Journal of Coal Science & Technology is a peer-reviewed open access journal. It focuses on key topics of coal scientific research and mining development, serving as a forum for scientists to present research findings and discuss challenging issues.
Coverage includes original research articles, new developments, case studies and critical reviews in all aspects of scientific and engineering research on coal, coal utilizations and coal mining. Among the broad topics receiving attention are coal geology, geochemistry, geophysics, mineralogy, and petrology; coal mining theory, technology and engineering; coal processing, utilization and conversion; coal mining environment and reclamation and related aspects.
The International Journal of Coal Science & Technology is published with China Coal Society, who also cover the publication costs so authors do not need to pay an article-processing charge.
The journal operates a single-blind peer-review system, where the reviewers are aware of the names and affiliations of the authors, but the reviewer reports provided to authors are anonymous.
A forum for new research findings, case studies and discussion of important challenges in coal science and mining development
Offers an international perspective on coal geology, coal mining, technology and engineering, coal processing, utilization and conversion, coal mining environment and reclamation and more
Published with the China Coal Society
Research Article
Open Access
Published: 10 July 2024
0 Accesses
International Journal of Coal Science & Technology Volume 11, article number 58, (2024)
1.
School of Minerals and Energy Resources Engineering, University of New South Wales, Sydney, Australia
2.
School of Mines, China University of Mining & Technology, Xuzhou, China
3.
State Key Laboratory of Coal Resources and Mine Safety, Xuzhou, China
Assessment of mining impact on groundwater is one of critical considerations for longwall extension and sustainability, however usually constrained by limited data availability, hydrogeological variation, and the complex coupled hydro-mechanical behaviour. This paper aims to determine the factors and mechanism of groundwater depressurisation and identify knowledge gaps and methodological limitations for improving groundwater impact assessment. Analysis of dewatering cases in Australian, Chinese, and US coalfields demonstrates that piezometric drawdown can further lead to surface hydrology degradation, while the hydraulic responses vary with longwall parameters and geological conditions. Statistical interpretation of 422 height of fracturing datasets indicates that the groundwater impact positively correlates to panel geometry and depth of cover, and more pronounced in panel interaction and top coal caving cases. In situ stress, rock competency, clay mineral infillings, fault, valley topography, and surface–subsurface water interaction are geological and hydrogeological factors influencing groundwater hydraulics and long-term recovery. The dewatering mechanism involves permeability enhancement and extensive flow through fracture networks, where interconnected fractures provide steep hydraulic gradients and smooth flow pathways draining the overlying water to goaf of lower heads. Future research should improve fracture network identification and interconnectivity quantification, accompanied by description of fluid flow dynamics in the high fracture frequency and large fracture aperture context. The paper recommends a research framework to address the knowledge gaps with continuous data collection and field-scale numerical modelling as key technical support. The paper consolidates the understanding of longwall mining impacting mine hydrology and provides viewpoints that facilitate an improved assessment of groundwater depressurisation.
Longwall mining is a productive method for underground coal recovery, however leaving extensive and long-term dewatering impacts on the regional hydrology (Booth 1986, 2006; Luan et al. 2020; Sidle et al. 2000; Zhang et al. 2011) and surface features that are susceptible to groundwater drawdown (Hebblewhite 2009). In Australia, concerns over past and potential future impacts have been challenging longwall approvals for years due to environmental, social, and governance (ESG) considerations. Evaluating groundwater depressurisation and its further influence on the regional water system has been a component of Australian longwall practice.
Assessment of the groundwater impact involves a parameter often referred to as “height of fracturing” (HoF), which measures the upper limit of continuous fracturing with respect to the mining horizon and can be determined via Lugeon style packer testing coupled with borehole acoustic or optical scanning. A number of conceptual models derived from longwall practice and field measurements summarise the fracturing mechanics and groundwater hydraulics above a mined panel (Adhikary and Guo 2015; Adhikary et al. 2017; Coe and Stowe 1984; Galvin 2016; Guo et al. 2007; Hebblewhite 2020; Hebblewhite et al. 2008; Kendorski 1993, 2006; Kendorski et al. 1979; Minns et al. 1995; Seedsman 2019; Tammetta 2013, 2015, 2016). Empirical descriptions indicate a positive correlation between HoF and mining height, while the coefficient varies a lot, from 15 to 60, between mines. Rock units below the HoF horizon exhibit fracturing and caving into the mined-out area, with pronounced increases in porosity and hydraulic conductivity (Heritage et al. 2022a, b). Strata above the HoF horizon coherently sag and remain uncracked, acting as barriers separating upper aquifers from the drainage effect underneath (Booth 1999). The ground surface has temporary fractures in tensile areas lateral of the subsidence trough (Kendorski 1993, 2006), featuring water table decline due to vertical permeability enhancement (Minns et al. 1995; Tieman and Rauch 1986) and potential groundwater regime change in the long term (Seedsman 2019). These identifications have formed a fundamental understanding of longwall mining impacting the regional hydraulics.
To current knowledge, the groundwater depressurisation is already easy to understand, but still challenging to predict due to multiple influencing factors, complex fluid flow behaviour in fractured rock masses, and geological and hydrogeological variations between mines. In addition, field data are often unavailable to underpin an improved understanding of these problems. The groundwater impact prediction is no longer simply about investigating rock fractures in terms of its intensity and distribution above mined panels, but also about assessing the short- to long-term drainage effect of fracture networks on groundwater regime. In this context, this paper takes a step towards determining the factors and mechanism of mine dewatering, identifying knowledge gaps and methodological limitations, and establishing a research framework to facilitate achieving further understanding and assessment of groundwater depressurisation at mine sites.
A direct result of groundwater depressurisation is an increase in mine inflow, which has been recognised from a flooding event at Heaton Colliery, England, in the early 1800s (Zhou 2017). By the late nineteenth century, engineers at an Ohio mine were conscious that the water stored in sandstones and gravels could flow into roadways and lead to convergence, and that coal pillars were significant for stabilising overlying waters (Haseltine 1892).
Hebblewhite et al. reported that underground mining in the Southern Coalfield, NSW, Australia, caused water leakage from deeper consolidated bedrocks and shallow unconsolidated sediments, along with loss of surface flows recharging subsurface voids through stream bed cracks (Hebblewhite 2009; Hebblewhite et al. 2008). Groundwater table declines because the recharge from surface waters is insufficient to offset the rapid leakage as longwall panel retreats underneath. Subsurface water leakage was reported at some Illinois mines (Booth 2006, 2007; Booth et al. 1999, 1998) and coalfields in Pennsylvania (Wahler 1979), West Virginia (Sgambat et al. 1980), and Kentucky (Hutcheson et al. 2000; Minns et al. 1995).
The Australian experience suggests that the dewatering effect can extend to surface flows and ecology. Significant natural features including watercourses, upland swamps, and important flora and fauna in aquatic ecosystems can be disrupted (Darmody et al. 2014; Hebblewhite et al. 2008; Jankowski and Knights 2010; Kay et al. 2006; Mills 2007; Zhang et al. 2018a). Such surface dewatering was also reported in coalfields in China (Fan and Zhang 2015; Fan and Ma 2018; Sun et al. 2021; Zhang et al. 2011), US (Dixon and Rauch 1990; Gill 2000; Sidle et al. 2000), and India (Chabukdhara and Singh 2016; Dhakate et al. 2019; PanIgrahy et al. 2015), and more pronounced in arid and semi-arid mine sites with scarce rainfall recharge (Luo et al. 2018). For example, the Kuye River in northwest China used to be a perennial watercourse through Ordos Basin coalfields. 30 years of underground mining has significantly shrunk the river, and the reduced drainage area, number of tributaries, and annual streamflow are less likely to recover (Guo et al. 2016, 2019a; Li et al. 2016b; Luan et al. 2020). The water level decline can cover a far broader area than the longwall domain (Frimpter and Maevsky 1979), which gives rise to extensive ecological impacts since the dropped water table is no longer as serviceable as before. Case studies show that the topsoil moisture can be impacted (Bian et al. 2009; Yang et al. 2016), followed by degradations of wetland, agricultural, and forest ecosystems whose survival is dependent on high moisture contents (Bai et al. 2022; Hebblewhite 2009; Lechner et al. 2016; Vishwakarma et al. 2020).
These cases demonstrate that the dewatering effect of longwall mining initiates from the seam horizon and propagates to bedrock and regolith aquifers and then to surface aquatic systems. Such features can be considered indicators of mine dewatering occurring, although to some extent different between mines and coalfields due to variations in mining parameters and regional hydrogeology. As summarised in Table 1, the indicators are categorised into five levels according to how closely they relate to and to what extent they represent the dewatering.
Position | Objects | Indicators | Monitoring approaches or tools | Level | References |
---|---|---|---|---|---|
Surface | Surface water | Water level | Water level sensor Stream gauge | DS | Bay (1968), Booth (2006), Dhakate et al. (2019), Ji et al. (2014) |
Streamflow | Remote sensing Stream gauge | DS | Ackman and Jones (1991), Booth (2006), III et al. (2014), Li et al. (2016b) | ||
Drainage area | Remote sensing | DS | |||
Tributary number | Remote sensing | DS | |||
Water quality | In situ testing Sampling and lab testing GIS contouring | IS | Ali et al. (2017), Chatterjee et al. (2010) Dhakate et al. (2019) | ||
Topsoil | Moisture content | Tensiometer Remote sensing | IS | ||
Fertility | Sampling and lab testing Trace element detection | IS | |||
Plant | Vegetation index | Field sampling Remote sensing | IS | Heras et al. (2008), Lei et al. (2010), Trotter and Frazier (2007), Yang et al. (2018) | |
Sub-surface | Groundwater | Water level | Piezometer Well sensor | DA | |
Flow rate and direction | Spontaneous potential Isotope tracing | DM | Artugyan and Urdea (2014), Frost et al. (2002), Mather et al. (1969) | ||
Water quality | Sampling and lab testing | IM | Booth and Bertsch (1999), Chatterjee et al. (2010), Tiwari et al. (2016) | ||
Overburden strata | Water head and permeability | Piezometer Packer testing Geoelectrical measurement | DA | Butler (2019), Niwas et al. (2011), Smerdon et al. (2014), Walsh et al. (2022) | |
Fractures | Geophysical logging Packer testing | DA | Bureau (2000), Corbett (2022), Heritage et al. (2022a), Wei et al. (2017) | ||
Mine water | Roof seepage | Visual check | DM | Zhang et al. (2013c) | |
Water inflow | Sensor network Underground upholes | DA | |||
Water ponding | Visual check Electromagnetic detection | DS |
This section identifies factors influencing the groundwater depressurisation at longwall mines. An HoF database is established sourcing extensive field data from 168 mines in tens of coalfields in China, as summarised in Fig. 1. The 422 HoF datasets are interpreted from both the mining activity and geology and hydrogeology aspects. Case analyses and comparisons are also conducted to consolidate the depressurisation contribution of some factors, or to serve as supplements for factor identification.
Mining height and panel width are geometric parameters of a panel, widely accepted as dominant factors influencing fracture propagation and groundwater drainage. Here an arch model is introduced to provide an insight into the HoF correlation to panel geometry, Fig. 2. In the panel domain, continuous fracturing propagates upwards and terminates at a unit where fragmented blocks are hinged, above which the layers sag but remain uncracked. An equilibrium arch structure forms with the upper limit of continuous fracturing as the top, Fig. 2a. The arch is composed of the suspended end of each layer, bearing vertical loading and non-uniform horizontal compression.
The arch is symmetric along the vertical centreline. The left half is taken for structural mechanics analysis, Fig. 2b. The bending moment at the top vertex O obeys:
where \({h}_{f}\) denotes the arch rise, pyhsically the height of fracturing; \(w\) is the arch span, i.e., panel width; \({F}_{vi}\) and \({F}_{hi}\) are vertical and horizontal forces acting on the ith straum; \({F}_{sv}\) and \({F}_{sh}\) are vertical and horizontal reactions of the abutment. The bending moment at any point along the arch trajectory should be zero because of the static equilibirum:
Combining Eqs. (1) and (2) obtains:
Equation (3) suggests a positive correlation between HoF and panel width in the ideal context of rigid abutment. In realistic scenarios, abutment (pillar) compression and failure cause mining height and rock strength factors to be involved via the mechanical term \({F}_{sv}\), for example the Rezaei semi-empirical equation (Rezaei et al. 2015a):
where \(t\) is mining height; \(\gamma\) is the unit weight of overlying rocks; \(D\) is the depth of cover; \({E}_{p}\) and \({E}_{g}\) are elastic modulus of pillar and goaf materials. Substituting Eq. 4 into Eq. 3 indicates that the height of fracturing is positively correlated to mining height and panel width, as demonstrated by many case studies (Fan and Zhang 2015; Fan et al. 2020; Hou et al. 2020; Karacan et al. 2005; Mills and O'Grady 1998; Zhang et al. 2011) and summaries (Fan et al. 2019c; Khanal et al. 2019b; Mondal et al. 2020).
Based on this mechanical analysis, the paper compares the contribution of mining height and panel width to HoF by regression analysis on the 422 datasets. Hypothesis testing has been performed in advance to ensure the statistical comparability required for data regression. Table 2 lists the regression equation (Eq. 5) and associated information. The coefficients indicate that mining height has a greater contribution than panel width. The R-square of 0.66 suggests that the linear model captures a moderate to substantial portion of the underlying relationship between HoF and panel geometry factors. The residual 34% of the variability is attributed to other factors that the model does not account for, implying that HoF predictions only based on panel geometry factors can be inaccurate or randomly accurate.
Regression equation | \(13.43t+0.212w-29.92\)(Eq. 5) | |
---|---|---|
Weight coefficient | t | 13.43 |
w | 0.212 | |
R-square | 0.66 | |
Upper bound (Confidence level of 95%) | \(14.42t+0.26w-19.73\) | |
Lower bound (Confidence level of 95%) | \(12.34t+0.17w-39.41\) |
Figure 3 compares predictions and measurements, including an HoF-t plane projection in Fig. 3b and HoF-w plane projection in Fig. 3c. Comparing Figs. 3b, c shows that the height of fracturing is more dependent on mining height because of a better consistency between predictions and measurements in the HoF-t plane. The HoF-t ratio can be almost the same between panels with similar longwall parameters at a mine, for example, approximately 18.3 for three panels (w = 300 m, D = 215 m) at Daping Mine, 20.9 for five panels (w = 198 m, D = 117 m) at Jinjitan Mine, and 27.1 for four panels (w = 193 m, D = 146 m) at Ningtiaota Mine, as shown in Fig. 3b.
Continuous fracturing is more likely to approach the shallow aquifer and cause subsurface dewatering in shallow mining conditions. Reportedly, the height of fracturing over panel LW9 at Dendrobium Mine, Australia, was almost 100% of the depth of cover at topographically low points (Brown and Minchin 2018), partially because the panel extraction caused the low surface topography to converge towards the seam horizon. Such extensive fracturing was also captured in Ordos Basin coalfields, China, for example, panel LW22407 at 131 m depth at Halagou Mine (He et al. 2020) and LW1203 at 50–65 m depth at Daliuta Mine (Wang et al. 2018a). In these cases, all rock units above mined panels are included into caved and fractured zones, and the conventional four-zone or five-zone structure becomes no longer applicable (Meng et al. 2016; Zhao et al. 2019).
Another consideration of the depth factor is about its correlation to mine subsidence and continuous fracturing, which involves a concept termed “critical extraction”. This terminology evaluates whether full ground subsidence is allowed to occur in the longwall domain, measured using the ratio of extraction width to depth, i.e., \(w/D\). Subcritical extraction refers to where the \(w/D\) ratio is insufficient to allow the full potential subsidence, and contrast is supercritical extraction (Li et al. 2022). The Australian longwall experience suggests that,
If individual panels have \(w/D\) ratios ≥ 1.4, the longwall domain is critical or supercritical, while,
If individual panels have \(w/D\) ratios < 1.4, the longwall domain can be critical or supercritical when the total void width is greater than 2.5D.
Figure 4 compares several cases of incremental subsidence under sub- to super-critical conditions. The Hunter Coalfield case demonstrates a supercritical extraction, featuring minor panel interactions since the caving and fracturing within a panel is encapsulated by chain pillars and thus not significantly impacted by the extraction of subsequent panels. The subsidence profile gradually fattens and panel interaction enhances with \(w/D\) decreasing, as indicated by the Daliuta case (Xu et al. 2021) and East Ordos case. The 450 m deep longwall domain in the Southern Coalfield has a \(w/D\) ratio of 0.32, characterising a “bath-tub” concave profile with substantial chain pillar compression and panel interaction (Li et al. 2022). The impact of sub- or super-critical extraction on groundwater depressurisation can be complicated. Supercritical extraction leads to insignificant panel interactions. The dewatering is constrained within each panel, while fractures can be more interconnected in areas with high tilts and strains. In contrast, subcritical extraction leads to significant panel interactions and compression of chain pillars and surrounding strata. The groundwater depressurisation of sequential panels is mutually connected to cover the entire longwall domain, while the post-mining groundwater recovery can be more satisfactory, as indicated by the reported Jefferson County case (Booth 1999, 2002, 2006).
A data regression is conducted to evaluate the contribution of the depth of cover, obtaining the following equation:
Equation (6) suggests a positive HoF-D correlation, while the cover depth has a comparatively smaller contribution than mining height and panel width. Equation (6) has an R2 of 0.77, meaning that the HoF prediction is improved after the depth of cover is taken into account. Figure 5 exhibits the prediction interval with a confidence level of 95% and 422 HoF data points. The model underestimates several top coal caving cases with large mining height (9–13.5 m) and moderate depth of cover (420–695 m). The number of cases beyond prediction bounds is decreased from the previous 134 in Eq. (5) to 87 in Eq. (6), demonstrating the influence of the depth of cover on HoF and the advantage of incorporating it as one of independent variables for HoF prediction.
Retreat rate is considered a flexible factor that can be adjusted to change the perturbance induced by coal removal (Luo et al. 2001; Zhang et al. 2009a, 2011). A potential explanation is that increasing the retreat rate helps reduce the frequency of lower, advance, and set (LAS) cycles, mitigate the roof failure induced by excessive confinement from canopy, prolong periodic weighting intervals, and shorten the propagation of frontal abutment pressurisation (Cui et al. 2020). Another explanation involves the voussoir beam theory, where an increased retreat rate can enhance inter-block friction and decrease sliding displacement of key voussoir blocks (Wang et al. 2020b). Longwall experience in Australia, China, and the US suggests a negative correlation of dynamic deformation, hydraulic conductivity, and fracture propagation versus retreat rate (He et al. 2020; Karacan and Goodman 2009). Comparatively, some German experience holds a view that lower retreat rates can mitigate the subsidence impact on surface structures (Luo et al. 2001).
Zhang et al. investigated the retreat rate impact on groundwater drainage (Zhang et al. 2010), as shown in Fig. 6. The water table decline decreased from greater than 1.5 m at BH11 and BH17 to about 1 m as the retreat rate was gradually increased from 5 m/d near BH17 to more than 15 m/d after BH04, Figs. 6b, c, accompanied by mine inflow decreasing from 20 to 6.7 m3/h. The results indicate that an increased retreat rate helps mitigate the regional groundwater drawdown. Another case is about the direct impact of different retreat rates on pillar rib convexity, measured by the authors at a longwall panel with depth of cover of 660 to 750 m. Figure 7 compares the rib deformation at three monitoring points MP1 to MP3. The results show that MP2 where the face passes with the highest retreat rate features the most pronounced rib deformation rate. The comparison suggests that a rapid retreat causes rib convexity to be accelerated, while the total deformation reduces. Such impacts need further investigation via more field data, accompanied by numerical parametric studies to quantify the pillar compression, continuous fracturing, and incremental subsidence responses to varying retreat rates.
Groundwater drawdown responses to retreat rates at Bulianta Mine (Zhang et al. 2010). a Exhibits the distribution of monitoring boreholes b Summaries the water table variation at 11 boreholes c Compares the average water table decline before and after the face acceleration
Pillar rib convexity with face retreating at different rates. The MP1 position is taken x-axis 0, and MP2 and MP3 are respectively x = 48 and 88. The 40 m long zone centred around each monitoring point is considered the critical period of pillar pressurisation and compression due to face pass, labelled in different colours for ease of comparison
Chain pillar is a key component of a longwall system. It has been identified that the pillar geometry influences rock fracturing and piezometric responses (Booth 1986; Tammetta 2016), and that increasing the aspect ratio helps enhance the pillar strength and thus stabilise overlying waters (Galvin et al. 1999; Unlu et al. 2013). The pillar contribution can be understood from the arch structure in Fig. 2. Assuming the pillar rib has a yielding depth of \({w}_{py}\), the extraction width (\(w\)) should sum the primary mined-out span (\({w}_{m}\)) and rib yielding depth (\({w}_{py}\)):
Substituting Eq. (7) into Eqs. (3) and (4) obtains that the height of fracturing increases with the yielding depth of pillar rib. Such impact is pronounced in narrow pillar configurations where the yielding depth increases due to low pillar strength and hence panel interaction intensifies in sequential panel extractions (Xie et al. 2020). For example, the Kemira Colliery experience indicates significant pillar failure and ground subsidence enhancement by more than 1 m above three sequential panels with chain pillars about 20 m (Mills 2022). Comparatively, as indicated by the Bulianta Mine experience, Fig. 6, the 50 m wide pillar constrained the horizontal expansion of groundwater drainage to BH18, to some extent contributing to the mitigation of aquifer dewatering and mine inflow (Zhang et al. 2010).
The HoF database collects several cases regarding panel interaction and HoF increment due to narrow pillars between panels. Figure 8a exhibits HoF increments with sequential panel extractions at four mines. Comparing the two histograms of each mine suggests that the second panel has a greater HoF than the first. Xiagou Mine adopted the smallest pillar width under a large mining height, featuring the most pronounced panel interaction and HoF increment. Xinglongzhuang Mine shows a slight HoF increment due to the largest pillar aspect ratio \({w}_{p}/{h}_{p}\) among the four cases. The panel interaction degree is negatively correlated to \({w}_{p}/{h}_{p}\). The comparison consolidates that panel interaction can lead to fracture network expansion and dewatering intensification (Booth 1986).
Vertical interaction can occur in multi-seam mining where the upper goaf is reactivated due to lacking competent interlayers to buffer the upward depressurisation induced by the panel extraction underneath (Yang et al. 2019). Figure 8b compares two scenarios of multi-seam mining and induced continuous fracturing above two seams. At Hongliulin Mine, the lower panel extraction causes the new fracture network to connect to the upper one because the distance between two seams is smaller than the HoF of the upper panel. The two networks are hydraulically connected and superposed onto the lower seam, resulting in an extreme HoF measured by packer testing for the lower panel. In contrast, the Kuangou case has no vertical interactions since the interlayers are thick enough to prevent the continuous fracturing from reaching the upper goaf. In longwall practice, a stacked panel layout tends to result in more concentrated depressurisation, while a staggered layout increases the draw angle and depressurisation area (Ghabraie et al. 2017). Therefore, optimising the extension distance of the lower panel may help control the interaction impact on overlying rock units and groundwater hydraulics when vertical panel interaction is inevitable in multi-seam mining.
Longwall mining usually ultises single pass, longwall top coal caving (LTCC), or slice mining method according to the coal seam thickness and mine geology. The three methods have different impacts on groundwater depressurisation, as compared in Fig. 9. Figure 9a shows that the LTCC method features the greatest HoF/t ratio with respect to the single pass and slice mining methods. The Lu’an Coalfield datasets capture some details of the HoF responses to slice mining and LTCC mining, Fig. 9b, showing that, for the 6 m thick seam, slice mining has lower HoF/t ratios (11.8–15.8) than LTCC (18.8–20.2). The difference between single pass and LTCC methods is also significant as indicated by the Zhangji Mine and Kuangou Mine datasets. Du and Gao investigated such difference between LTCC and slice mining using physical and numerical simulation techniques (Du and Gao 2017), Fig. 9c. They attributed the smaller impact of slice mining to an increase in the bulking factors of caved materials. Slice mining causes a secondary depressurisation environment where caved rocks further crack and hence refill the mined-out area to a greater extent, such that the overburden bending and fracturing can be mitigated.
Comparison of single pass, slice, and LTCC mining methods in terms of induced continuous fracturing. a Compares the mining methods at three mines b Details the Lu’an Coalfield data c Compares the height of fracturing in LTCC and slice mining scenarios (Du and Gao 2017)
Slice mining can be employed as the primary method and combined with the single pass or top coal caving method within each slice to mine ultra-thick coal seams greater than 20 m (Fan et al. 2011; Wang et al. 2019). For example, Yaojie No. 3 Mine combines slice mining and sublevel caving for a 34–85 m thick seam (Dai et al. 2022). Such combination is also used at some mines in Junggar coalfields, where the average seam thickness is 90 m (Huang et al. 2018; Qin et al. 2019; Wang et al. 2019). Underground mining of such ultra-thick coal seams inevitably causes severe groundwater drainage and flow system dysfunction (Ren and Wang 2020; Zhang et al. 2018b). Therefore, further understanding the impact of panel geometry within each slice and interaction between adjacent slices is significant for drainage mitigation and post-mining groundwater recovery.
The stress environment in coal measure formations can be very different between rock units. In longwall practice, the horizontal stress is measured by maximum (\({\sigma }_{H}\)) and minimum (\({\sigma }_{h}\)) horizontal principal stresses. Figure 10 shows 58 stress measurements at 49 mines in China, from which the stress variation with depth is obtained as follows:
Figure 10a and Eq. (8) reflect that the vertical stress increases by approximately 25 kPa per metre with depth, while the increase in \({\sigma }_{H}\) and \({\sigma }_{h}\) is comparatively irregular. Figure 10b compares \({\sigma }_{H}/{\sigma }_{v}\) and \({\sigma }_{h}/{\sigma }_{v}\), indicating that (1) Horizontal stresses are greater than the vertical in shallow depth (D < 200 m) conditions, and (2) Horizontal stresses increase with depth but show a decreasing trend relative to the vertical stress. The increase in \({\sigma }_{v}\) and decrease in \({\sigma }_{H,h}/{\sigma }_{v}\) ratios may contribute to the positive correlation between HoF and depth of cover because the horizontal confinement on overburden is equivalently reduced. In longwall systems, underground workings are generally orientated parallel to \({\sigma }_{H}\) to minimise potential rock failures (Galvin 2016). In this context, the fracture propagation, viewed from the panel width section, is mainly influenced by the \({\sigma }_{h}\) magnitude, with a negative correlation (Vakili and Hebblewhite 2010). Shear stress also exists in coal measure formations, enhancing bedding plane shearing and parting as parallel to formations, while contributing to high-angle fracturing as perpendicular to formations. Quantification of the impact of horizontal stress on continuous fracturing and dewatering remains unanswered at present. Further research based on field data coupled with numerical parametric studies is therefore recommended.
Overburden of longwall panels is composed of varying stratum units, among which the competent ones provide mechanical support and soft ones help narrow fracture openings due to intrinsic dilatability (Ren and Wang 2020; Zhang and Peng 2005). It has been proposed that the HoF/t ratio has a positive correlation to the number of bedding planes and average porosity of overlying strata (Palchik 2002, 2003). A mechanical understanding involves the rotation and bulking behaviour of caved rock blocks. Competent and massive rocks such as sandstone and siltstone tend to cave at a gentler angle and bulk greater than weak laminated rocks such as shale and mudstone (Galvin 2016), thus refilling the goaf and mitigating fracture propagation to a greater extent (Rezaei et al. 2015b).
Mining experience of eastern coalfields in China suggests that the height of fracturing can be smaller in formations characterised by soft and weathered soft mudstones and claystones, exhibited in Fig. 11a. A potential explanation derives from the low permeability nature and limited permeability enhancement of soft geomaterials. Figure 11b compares several coal measure rocks in terms of their permeability variation in response to external loading. The bar chart indicates that the mudstone features (1) A lower initial permeability, approximately two to three orders of magnitude, than brittle rocks, and (2) Minor permeability changes from the elastic to failure stages. Scanning electron microscope images in Fig. 11b indicate that the mudstone is rich in clay minerals on the surface of voids and fissures. Such soft infillings can swell and narrow flow pathways (Gale 2008; Seedsman 2019), potentially contributing to the halt of aquifer dewatering and hence post-mining groundwater recovery (Fan et al. 2018; Sgambat et al. 1980). Further investigations are needed to better understand the mechanism underlying the disintegration and swelling behaviour of soft infillings in the fluid flow context.
Potential contribution of soft geomaterials to mitigating rock fracturing. a Presents HoF increase with mining height in different rock strength and lithology conditions, sourcing the mining experience of eastern coalfields in China b Compares the permeability of typical coal measure rocks and their microstructures (Fan et al. 2018; Zhang et al. 2013b)
Fault is a frequent geological structure at coal mines. Such large discontinuities in the vicinity of longwall panels can be activated due to mining perturbance, causing loose geomaterials to be more fragmented and natural fissures to be extensively connected as excellent fluid flow pathways. Massive groundwater migration can occur along faults that extend from panels to upper aquifers (Zhang and Peng 2005). A recent study details hydraulic features of mine faults via investigations on the Elouera Fault, a geological structure connecting the Wongawilli Seam at Dendrobium Mine and upper Avon Reservoir in NSW, Australia (Walsh et al. 2022). Borehole loggings indicate that (1) The mine fault is characterised by fault cores comprising centimetre-sized clay or pulverised gouges and breccias, (2) Fault surfaces can be slippery due to frequent slickensides, and (3) Fault cores are surrounded by damaged zones where the fracture density is 3 to 6 per metre, greater than the 1.5 fractures per metre in host rocks. Packer testings indicate that (1) Damaged zones are more permeable than 90% of host rocks, (2) High permeable zones can be locally discontinued and thus do not form continuous water flow channels, and (3) The fault shows a decrease in transverse permeability of 0.3–0.5 orders of magnitude, which depends on the juxtaposition of less permeable strata, cataclasis and mineralisation of the cores, and clay and shale smear (Fossen 2016).
Another typical structure is outcrop. Weathered geomaterials, especially coals, near outcrops often exhibit high permeabilities due to ubiquitous cleats and joints, enabling enhanced fluid flow pathways as preferential zones for rapid groundwater recharge (Brown and Parizek 1971; Zhang and Peng 2005). Karst-pillar intrusion is a less frequent defect at coal mines, potentially carrying a large volume of water and leading to water inrush into mine workings when mining or tunnelling passes the intrusion structure (Ma et al. 2015, 2018a, 2022). The two structures are more about mine safety and localised hydrology.
Valley is an irregular topography at mine sites, featuring steeply incised terrains where surface water accumulates and mine subsidence mechanics becomes complicated. The valley floor may contain natural fractures through which the hydraulic exchange between surface and subsurface waters becomes frequent (Booth 1986). Valley floor upsidence and valleyside closure can occur due to longwall mining and reshape the valley geomorphology, further altering the localised flow and circulation patterns, as illustrated in Fig. 12.
The upsidence and closure are identified as nonconventional subsidence behaviours and have been investigated for many years in Australia (Hebblewhite 2009; Mills 2007; Zhang et al. 2016). As shown in Fig. 12a, the incised valley structure hinders the continuity of horizontal stress and redirects it from hills into the valley floor. After panel extraction, the floor becomes overstressed and experiences near-surface shearing and buckling to release accumulated energy, accompanied by upsidence behaviour at the base of the valley. Figures 12b, c demonstrate near-surface strata buckling and bedding slab shearing at Waratah Rivulet in NSW, Australia. Valley closure is defined as an inward convexity of valley flanks (Zhang 2014; Zhang et al. 2016), forming due to horizontal compression at topographically low points. In these cases, hillsides move towards the valley centreline, often accompanied by tensile and shear cracking on both high sides (Hebblewhite et al. 2008). Nonconventional subsidence is also captured in some valley terrains in northwestern coalfields, China, as exhibited in Figs. 12d, e.
Previous research has studied the mechanical responses of irregular topography to mine subsidence (Barbato et al. 2016a, b; Fan et al. 2017; Hebblewhite 2009; Zhang et al. 2012, 2013a). Further investigations are expected to clarify the longwall impact on the valley hydrology, especially about the natural water occurrence in valley floor and enhanced hydraulic interaction between surface and subsurface waters due to mining.
The longwall impact on groundwater varies with hydrological settings regarding water distribution, abundance, and conductivity (Hebblewhite et al. 2008; Mills 2022; Walsh et al. 2022). At mine sites, water carriers can be surface watercourses, unconsolidated regolith gravels, and porous bedrocks. Such three carriers form seven types of water occurrence, further up to 14 types by taking into account the connectedness between water and panels, as summarised in Table 3. The single surface water occurrence is a hydrologic type that the water above longwall panels primarily exists in surface containers. Rivers, lakes, and reservoirs are large-scale water bodies posing the greatest threat to longwall mines due to abundant water supplies (Brown and Minchin 2018; Hebblewhite et al. 2008). When no impermeable unit intercepts surface and subsurface interactions or induced fractures propagate to the shallow surface, mine water inrush and surface water loss are highly likely to occur. Other types of containers are less risky because of comparatively small volume and seasonal recharge.
Types | Indirect contact with coal seam | Direct contact with coal seam |
---|---|---|
Single occurrence | ||
Surface |
|
|
Regolith |
|
|
Bedrock |
|
|
Compound occurrence | ||
Surface and regolith |
|
|
Regolith and bedrock |
|
|
Surface and bedrock |
|
|
Surface, regolith, and bedrock |
|
|
Subsurface waters are stored in unconsolidated sand and gravel clusters. Hydrologic surveys suggest that (i) upper and middle regolith water bodies have greater volumes and smoother discharge-recharge conditions than bottom ones, and (ii) Tertiary clays have better water-resisting capability than Quaternary clays because of longer sedimentation and thus lower porosity. As measured, three unconsolidated aquifers in the 360–484 m thick regolith of Pansan Mine, China, have a decreasing water yield from the shallower 1.41 to deeper 0.56 L/(m·s); five unconsolidated aquifers in the 260 thick regolith of Xingtai Mine, China, also show a decreasing yield from the shallower 8.54 to deeper 0.01 L/(m·s), with conductivity decreasing from 73.97 to 0.02 m/d. In deeper area the bedrock water is stored in pores of sandstone and limestone units, sometimes in fissures of geological discontinuities. Pore water mainly behaves percolation under micro- to super-capillary effects, influenced by pore diameters and porosity. Bedrock water should be drained before mining if contacting the coal seam or potentially within the scope of continuous fracturing.
Compound occurrences are more common at mine sites, as depicted in the last four rows of Table 3. The hydrological impact of longwall mining is most pronounced in cases where surface and subsurface waters are hydraulically connected, as surface flows can extensively recharge the groundwater. For instance, the maximum inflow reaches 3000 m3/h at Gaojiabao Mine, China, where longwall panels are merely 84 m below an ultra-thick bedrock aquifer (400 m in thickness and 7.4 MPa in water pressure) recharged by the Weihe River. The paper further classifies groundwater bodies into four structures in terms of the aquifer-aquiclude system and its position relative to the coal seam, as depicted in Fig. 13. The four structures include:
The simple structure has only one aquiclude of a single layer in the overburden, Fig. 13a. The aquiclude shows weaker anti-cracking performance than a multi-layer combination. Aquifer dewatering occurs when fractures penetrate the aquiclude.
The interlayered structure has more than one aquifer-aquiclude structure in the overburden, Fig. 13b. Aquifers are hydraulically separated by impermeable materials in the vertical direction, and thus the groundwater mainly behaves horizontal flow.
The enclosed and semi-enclosed structure represents that water bodies are entirely or partially encapsulated by impermeable materials, Fig. 13c. This structure characterises a limited source of recharge compared to others, and therefore the induced dewatering can be localised.
The unconformable structure forms when the aquifer-aquiclude structure has an unconformable contact with the coal seam, Fig. 13d. This structure may contain one of the above three structures in localised areas. Water barriers should be installed along coal seam boundaries to prevent water inrush.
Besides the above representative structures, mines may have hydrological anomalies associated with faults and Karst intrusions. Identifying the mine groundwater hydraulics should be fundamental for assessing the potential of aquifer dewatering due to sequential panel extractions and hence the likelihood and duration of groundwater recovery.
Groundwater recovery is a component of the hydrological impact of longwall mining, featuring a time dependency as it occurs with fracture closure after mining. Longwall experience of Yangquan coalfields suggests a time course of overburden responses, including the first 13 days for depressurisation development, 24 days for caved rock compaction, and 42 days for goaf reconsolidation stabilisation. Therefore, a groundwater recovery can be 5–11 weeks after the face passes the monitoring point. The Bulianta experience, Fig. 6, represents that the declined water started to recover about 7 weeks after mining.
A complete recovery needs a longer period. Figure 14 presents the water table variation at two mines in Illinois coalfields, US (Booth 1999; Booth et al. 2000). In the Jefferson County case, Well P350 captured a 10 m recovery 6 months after the maximum drawdown, followed by complete recovery in about one year. The adjacent Saline County featured a delayed recovery, where Well P54B recorded a further drawdown even two years after mining, and a complete recovery therein can be very challenging or may not occur if the site has poor transmissivity and inadequate recharge pathways. Besides, the long-term monitoring at Jefferson County captured that the recovered water table was higher than the pre-mining level. This is because the panel extraction resulted in a closed-end “bath-tub” zone in which the aquifer was elevated and well yields were increased (Booth et al. 1998). The difference between the two cases relates to the regional hydrology (Booth et al. 1998), including surface runoff and precipitation supply, aquifer storativity and rechargeability, and regolith water-holding capacity. This case comparison demonstrates the significance of mine hydrogeology identification and long-term piezometric monitoring for longwall impact assessment.
Water drawdown and post-mining recovery at Jefferson County and Saline County, US (Booth 1999). The monitored mining was from December 1988 to April 1989 at Jefferson, and from April 1992 to April 1993 at Saline
Researchers have adopted field measurement, lab test, and numerical modelling methods to understand the mechanism of longwall mining impacting groundwater stability, revealing that longwall mining leads to significant increases in rock porosity and permeability. This section analyses the dewatering mechanism from permeability enhancement and fracture propagation perspectives, considering two different roles they play in groundwater drainage.
In groundwater hydraulics, permeability is often used to measure the ability of rock media to allow fluid to pass through them. The paper uses permeability to describe the changes in fluid flow condition. There have been extensive laboratorial studies about rock permeability variation under external loadings (Wang et al. 1997; Zhang et al. 2018c; Zhu and Wong 1997), which includes a slight decrease due to pore convergence in the elastic stage, significant increase with progressive failure in the yield stage, and fluctuation in the residual stage (Fan et al. 2018; Heiland 2003; Heiland and Raab 2001; Zhao et al. 2017). The overall permeability enhancement features a positive correlation to the specimen fracturing degree (Fan et al. 2018; Wu et al. 2005). The lab-scale fractured rock hydraulics implies that, within a panel domain, the caved rock mass should have the highest permeability due to extreme fracturing, while the overlying intact strata have minor hydraulic changes. Such field-scale understanding has been represented by a number of conceptual models listed in Table 4.
ID | Contributors | Model profile | Empirical descriptions |
---|---|---|---|
I |
| Fractured zone—water drains into the goaf via interconnected fractures Aquiclude zone—minor permeability changes, as a barrier against water inflow from surface Surface fracture zone—permeability increases in the lateral tensile areas due to temporary open fractures | |
II | Tieman et al. (1986) |
| Aquiclude zone—preventing the mine from recharging, with relatively low permeability and minor vertical flow Surface area—significant vertical permeability enhancement due to tensile strain, accompanied by obvious water level decline |
III | Forster et al. (Adhikary and Guo 2015; Adhikary et al. 2017) in 1992 |
| Caved and fractured zones—significant permeability increase, where the water from aquifers flows rapidly into the mine Constrained zone—an increased horizontal permeability, unchanged vertical permeability in the central area, and an increased vertical permeability by both sides |
IV |
| Fractured zone—filled with vertically transmissive fractures Dilated zone—increased storativity, and little or no transmissivity Constrained zone—minor increases in transmissivity or storativity Surface zone—high-angle transmissive cracks and disruptions | |
V | Minns et al. (Coe and Stowe 1984; Minns et al. 1995) in 1995, after Coe and Stowe Model in 1984 |
| Fractured zone—draining to the mine, with slight water recovery after mining Additional zone—acting as an aquitard to limit groundwater flow Near surface zone – slight water drawdown due to increased porosity and permeability |
VI | Guo et al. (2007) in 2007 |
| Fractured zone and below—permeability increase up to 50 (vertical) and 200 (overall) times Constrained zone—permeability increase up to 2 (vertical) and 5–10 (overall) times Elastic zone—overall permeability increase up to 5 times Surface zone—vertical permeability increase up to 10 times |
VII | Tammetta (2013, 2015, 2016) in 2016, after the 2013 model and 2015 model |
| Goaf—acting as a pool where the flow rate is ruled by hydrodynamic factors rather than hydraulic conductivity Collapsed zone—great increase in hydraulic conductivity, where the flow may obey the Richards equation rather than Darcy’s law Disturbed zone—an increased hydraulic conductivity mainly in the horizontal direction |
VIII |
| Caved zone—relatively high porosity and permeability Fractured zone—vertical permeability enhancement with strata depressurisation promoted Constrained zone—permeability enhancement that is negligible vertically but possible horizontally due to bedding plane shearing | |
IX | Seedsman (2019) in 2019 |
| Disrupted zone—potentially minor conductivity increase Minor fracture zone—low fracture conductivity due to clay mineral infillings Potential spanning zone—a barrier of fracture propagation Surface impact zone—groundwater regime may change in the long term |
Hydraulic characteristics of a panel domain can be drawn from varying models in Table 4. Horizontally, the permeability enhancement is symmetrical along the panel centreline and more pronounced near trough edges due to high-angle tensile fracturing, indicated by Models III, V, and IX. Vertically, rock permeability generally increases with depth and to the maximum in the caved zone. The fractured zone features significant permeability enhancement by 1000 (Mills 2022) to 2000 times (Guo et al. 2007) and water head decline due to smooth drainage, ranging from 15 to 60 times of the mining height. The area labelled aquiclude zone (Models I and II), constrained zone (Models III, IV, VI, and VIII), or additional zone (Model V) has minor permeability enhancement but comparatively greater increases in transmissivity and storativity due to bedding plane shearing and parting (Mills 2022), potentially preventing the intrusion of overlying waters into mine voids underneath. Near ground surface the tensile areas lateral of the subsidence trough show an increased permeability, through which rainwater or other surface water sources can infiltrate the ground. Overall, the hydraulic differential forms a gradient along which the upper water tends to drain to the goaf under gravitation and flow momentum effects.
An indispensable condition for mine dewatering is that induced fractures should form a connected and conductive network to allow significant volumes of groundwater flow (Gale 2008), among which the isolated ones have rather limited contribution to smooth drainage (Kowalski et al. 2021). It is therefore necessary to assess the distribution of connected fractures and their connectivity. Lab experiments coupled with interpolated estimations from in situ packer testings have identified influencing factors of fracture connectivity, as summarised in Table 5, where internal factors are about attributes of fractures within a network, and external factors involve the rock environment in which fractures propagate.
Categories | Factors | Definition and correlation to the connectivity |
---|---|---|
Internal | Fracture intensity | Number of fractures per unit area or volume. Existence of more fractures means high likelihoods of connection. Fracture intensity is often expressed as \({P}_{ij}\) (\(i\) for dimensionality and \(j\) for counting unit); for example, \({P}_{10}\) counts the intersections of fracture per unit length, and \({P}_{32}\) the area of fracture per unit volume |
Fracture aperture | Size of the void between fracture surfaces. Wider fractures are more likely to be connected. The aperture favours a cubic contribution to the fracture transmissivity in cases of laminar narrow-channel flow | |
Fracture length | Distance along the fracture path. A longer fracture tends to intersect more fractures, potentially increasing the connectivity. The impact also depends on the fracture intensity, aperture, and orientation | |
Fracture porosity | Volume of voids (infillings excluded) within a fracture, calculated as the ratio of void volume to fracture volume. A high-porosity fracture allows fluid to flow smoothly, increasing the overall connectivity | |
Fracture orientation | Direction of a fracture relative to its host rock, including dip and strike components. Semi-parallel fractures tend to be more connected. The flow has a greater rate when perpendicular to the aperture variation direction | |
Geometric complexity | Fracture waviness and fracture surface roughness, respectively the degree of curvature along the fracture path and the irregularity and asperity of fracture surfaces. Fractures with high waviness and roughness are less connected because the high complicity causes the pathway to be blocked | |
External | Rock property | Strength and deformability of rocks hosting the objective fracture. Brittle rocks allow fractures to be more interconnected and maintained due to the low rock ductility and high fracture toughness. Mineralogically, clay minerals can narrow fractures and decrease fracture connectivity after disintegrating due to water. Existence of natural fractures causes rocks to be more fractured |
Fluid property | Fluid pressure and chemistry. High fluid pressure decreases effective stress and increase fracture connectivity. Acidic fluids can dissolve minerals, widen fractures, and thus increase fracture connectivity | |
Stress regime | Stress level, stress state, and loading history. High tectonic stress causes fractures to be closed more easily. Rocks have more connected fractures under compressive and shear stress but more isolated fractures under tensile stress. Rocks subjected to repeated loading and unloading can be more fractured | |
Time and temperature | Fractures gradually converge due to stress recovery over time. High ground temperature causes rock strength decline, clay mineral deposition, and continuous fracturing to be enhanced |
Internal factors indicate that the interconnectivity of mine fracture network should be greater in the caved zone and minimum in the constrained zone where rock layers sag but remain intact (Hebblewhite 2020). Areas lateral of the subsidence trough have an increased connectivity due to local high-angle fractures (Heritage et al. 2022a). In shallow mining scenarios, the connectivity may show a monotonic decrease with elevation because of through-going fracturing (Brown and Minchin 2018; Brown and Walsh 2022; Meng et al. 2016; Zhao et al. 2019). Influence of external factors implies that (i) stiff and brittle rock masses have greater fracture connectivity than coaly and tuffaceous ones (Galvin 2016), (ii) the high connectivity of caved and fractured zones decreases with goaf reconsolidation over time, accounting for the post-mining recovery of in situ stress and groundwater, and (iii) coal measure rocks in context of narrow pillar configuration or multi seam mining have a higher degree of fragmentation and fracture connectivity due to intensified panel interaction or repeated pressurisation and depressurisation.
The increasing connectivity with depth forms a smooth fracture flow network through which overlying waters drain to the mined-out area underneath. Fracture flow dominates the drainage behaviour, as connected fractures provide steeper pressure drop gradients and smoother fast-flow pathways than rock matrices (Booth and Spande 1992; Hill and Price 1983; Jeanpert et al. 2019). A fundamental model for fracture flow description is the cubic law:
where \(Q\) is flow rate, in \({\text{m}}^{3}/\text{s}\); \(\Delta p/L\) measures the pressure gradient along flow path, in \(\text{pa}/\text{m}\); \(\mu\) is dynamic viscosity, in \(\text{Pa}\bullet \text{s}\); \(l\) is the size of fracture orthogonal to flow, in \(\text{m}\); \(a\) is fracture aperture, in \(\text{m}\). The cubic law assumes laminar flow through a smooth tabular chamber, ideally two parallel plates with distance (aperture) much smaller than length along the fluid flow direction. Snow investigated multiple pathways in jointed rocks and proposed a model for hydraulic conductivity (Snow 1965, 1969):
where \(\rho\) is fluid density, in \(\text{kg}/{\text{m}}^{3}\); \(g\) is gravitational acceleration; \({n}_{i}\) is the unit normal vector of the conduit and \({n}_{j}\) is direction cosines of the unit vector, together determining the orientation of the conduit; \({s}_{i}\) is an average spacing of the joint set along the sampling line; \({\delta }_{ij}\) is the Kronecker delta. The summation is taken over all members of the joint set. Zoorabadi et al. innovated the Snow model to describe the hydraulic conductivity of jointed rock mass (Zoorabadi et al. 2012, 2022):
where \({V}_{R}\) is the volume of rock mass before excavation and \({V}_{r}\) is volume of excavation; \({m}_{k}\) is the number of fractures of the kth set; \({c}_{k}\) is interconnectivity coefficient, a positive number not greater than 1 and calculated via \(({\sum }_{k=1}^{n}\frac{{l}_{k}}{{s}_{j}}\text{sin}{\gamma }_{kj})/{I}_{max}\) (\(k\ne j\), \({l}_{k}\) is the mean trace length of set \(k\), \({s}_{j}\) is the mean spacing of set \(j\), \({\gamma }_{kj}\) is the average angle between fractures of set \(k\) and set \(j\), and \({I}_{max}\) is the total interconnectivity index of the fracture network); \({d}_{k}\) and \({a}_{k}\) are respectively the diameter and aperture of set \(k\); \({n}_{ik}\) and \({n}_{jk}\) are direction cosines of the unit vector normal to each fracture. The Zoorabadi model considers fracture frequency, orientation, size, and aperture, obtaining a more accurate result than the Snow model when applied to estimate the hydraulic conductivity of rock mass (containing five joint sets with micron-scale apertures, with a mean conductivity of \(1.8\times {10}^{-8} \text{m}/\text{s}\)) surrounding a tunnel.
Compared to existing fracture flow descriptions, flow pathways in longwall domain have a broader distribution (15–60 times of mining height), higher frequency (5–10 on the total water loss horizon and 10–15 near the goaf centre Heritage et al. 2022a, b), and larger apertures (up to 100 mm (Price et al. 2022)). Extensive intersections further complicate the pressure drop of fracture flow, and the drainage behaviour in fractured and caved zones may involve high-Reynolds flow and non-Darcian turbulence (Tammetta 2015). There have not been analytical or semi-empirical models applicable to describe fluid flow through extensive interconnected pathways. Addressing the knowledge gap should be based on fracture network identification, which, however, is limited by field data acquisition, especially for bedrock fractures beyond the detection range of acoustic and seismic techniques. In this context, mine fracture observation should be improved, accompanied by continuous fracturing representation in numerical models, to help quantify fracture connectivity and hence establish a description of the fracture flow behaviour.
Empirical equation, field measurement, and numerical modelling are frequently used to evaluate continuous fracturing and the potential of groundwater drainage. This section discusses current applications of the three methods to groundwater drainage assessment and identifies their limitations that should be concerned in future research.
Empirical prediction of mine dewatering utilises HoF prediction equations. Australian longwall practice often adopts two empirical equations, one established by Ditton et al. using HoF datasets from Australian coalfields (Ditton and Merrick 2014), and one by Tammetta using data collected from tens of mines in eight countries (Tammetta 2013). Chinese mines often use equations proposed by China Coal Research Institute (CCRI) according to the longwall experience of eastern coalfields in China (CCRI 1981). Table 6 lists the three equations and their applicability.
Equation ID | Expression | Applicable to |
---|---|---|
Equation 9 | \(\text{HoF}=1.52\times {{w}{\prime}}^{0.401}\times {D}^{0.535}\times {t}^{0.464}\times {{t}{\prime}}^{-0.4}\) | – |
Equation 10 | \(\text{HoF}=1438\times ln(4.315\times {10}^{-5}u+0.9818)+26\) \(u={t}^{1.4}\times w\times {d}^{0.2}\) | – |
Equation 11.1 | \(\text{HoF}=100t/(1.2t+2.0)\pm 8.9\) | Competent rock environment (UCS greater than 40 MPa) |
Equation 11.2 | \(\text{HoF}=100t/(1.6t+3.6)\pm 5.6\) | Medium competent rock environment (UCS from 20 to 40 MPa) |
Equation 11.3 | \(\text{HoF}=100t/(3.1t+5.0)\pm 4.0\) | Soft rock environment (UCS from 10 to 20 MPa) |
Equation 11.4 | \(\text{HoF}=100t/(5.0t+8.0)\pm 3.0\) | Weathered soft rock environment (UCS smaller than 10 MPa) |
Figure 15 compares the three equations along with Eq. 6 in terms of their performance in estimating the height of fracturing at 20 panels in NSW and Bowen Basin coalfields in Australia (Ditton 2014) and 20 panels in Ordos Basin coalfields in China (He et al. 2020). The Ditton Equation has the best performance for Australian cases, potentially because it considers more influencing factors than others. Equation 6 has a better performance for Chinese cases, while less accurate for Australian cases since the equation is established from HoF datasets collected from Chinese coalfields. The CCRI Equation underestimates all the Chinese cases with significant errors, meaning that such an equation based on the eastern longwall experience is no longer applicable to Ordos Basin coalfields where longwall panels usually have a greater geometry and shallower depth.
Comparison of four empirical equations for HoF estimation. Notes The first 20 data points shaded in green are Australian cases, and the following 20 points shaded in red are Chinese cases. The CCRI Equation uses Eq. 11.1 for Australian cases considering a competent rock environment (Ditton 2014), and Eq. 11.2 for Chinese cases, where the overburden is medium competent (He et al. 2020; Ren and Wang 2020). The effective thickness of competent stratum is 20 m for Chinese cases (Fan and Zhang 2015; Zhang et al. 2010, 2011)
The comparison indicates that (i) empirical prediction is often constrained by the background information of the employed equation due to potentially significant differences in mining parameters and geological conditions between mines, and (ii) considering more contributing factors can help improve the accuracy of empirical prediction. Therefore, continuous HoF data collection and high-quality statistical interpretation should be conducted, which are significant for establishing a reliable model for longwall impact estimation.
Continuous measurement and data collection have been a critical part of Australian longwall practice, helping evaluate mine risks, update longwall experience, and provide parameters and criteria for numerical model construction and verification. For example, at Elouera Mine, massive data collected from field mapping, downhole video inspection, geophysical logging, packer testing, and overcore measurements helped the mine control water inflow from the Avon Reservoir to underground workings (Walsh et al. 2022). The Standardised Subsidence Information Management System (SSIMS) established in NSW, Australia, allows users to interrogate the information stored in the Subsidence Database for purposes of coal mining and risk regulation in regard to subsidence mechanics (Li et al. 2022).
Measured data should be managed with explicit classifications that enable efficient data retrieval and interpretation, detection of missing components, and identification of geological difference between panels or mines. The paper designs a structure for a future potential mine groundwater database, as exhibited in Fig. 16. In the database, every panel should contain at least one dataset of ground subsidence, height of fracturing, rock mass conductivity, and water table as key data. Panel geometry, stratigraphic profile, in situ stress, and rock strength are included as basic data, determined from mine drawings, borehole logs, borehole breakout records, and core test reports, respectively. The database should have a wide breadth, requiring that the datasets of sequential panels within a coal seam should be extended to deeper seams, adjacent mines, and even neighbouring coalfields. Another requirement is the coverage in depth, involving regular borehole geophysical logging and packer testing to determine the variation in HoF and hydraulic conductivity with longwall progress and time, accompanied by long-term monitoring of groundwater drawdown and recovery. Many mines in Australian coalfields, such as Dendrobium Mine (Brown and Walsh 2022), United Colliery (Heritage et al. 2022a), and Springvale Mine (Corbett 2022) are examples of extensive field measurement and long-term monitoring.
Numerical modelling is often used to assess rock mass failure and fracture propagation. The approach relies on a methodical workflow including mine data analysis, rock strength determination, in situ stress transformation, coding for model construction, model calibration, and model verification. Compared to empirical predictions, field-scale modelling is time-consuming but potentially more informative by configuring site-specific parameters about panel geometry and geological condition. Many numerical techniques have been developed to assess fluid flow or percolation due to underground excavations, as summarised in Table 7.
Speciality | Technique | Software | Applications |
---|---|---|---|
Geomechanics and geotechnical modelling | Discrete element Method (DEM) | PFC2D/3D | Height of connective fractures and permeability enhancement (Adhikary et al. 2017; Khanal et al. (2019a, 2021); strata fracturing and bedding plane parting (Liu et al. 2019); rock fracturing and porosity variation in overburden (Wang et al. 2017a); rock fracturing and permeability enhancement from multi-scale perspectives (Poulsen et al. 2018) |
UDEC/3DEC | Distribution of fractured zone (Du and Gao 2017; Fan et al. 2019b; Ren and Wang 2020; Wang et al. 2016; Zhang et al. 2017); void ratios of fractures in overburden (Wang et al. 2017b); impact of stress and water pressure on rocks on lab scale (Fan et al. 2019a; Zhang et al. 2018c); fluid flow via fractures based on DFN setups (Liu et al. 2015a); fluid flow in fractures with different morphologies (Ma et al. 2018b) | ||
Finite difference method (FDM) | FLAC2D/3D | Water inrush due to discontinuity activation (Ma et al. 2015, 2018a, 2016; Wu et al. 2004); strata permeability variation and water flow regime (Fan et al. 2020; Meng et al. 2016; Wang et al. 2015); position of water-conducting fractured zone (Guo et al. 2019b; Liu et al. 2015b); fracture distribution and hydraulic conductivity variation (Gale 2005; Heritage et al. 2017; Lu et al. 2020); water–gas seepage via hydraulic fracturing channels in multi-field coupling condition (Fan et al. 2021a, b) | |
Finite element method (FEM) | ABAQUS | Influence of tunnelling on seepage field and groundwater environment (Meye and Shen 2020); influence of groundwater level fluctuation on lateral cantilever structure of pit (Wang et al. 2020c); groundwater inflow while tunnelling and their interaction (Lee and Moon 2020) | |
ANSYS | Location of permeable fractured zone (Wu et al. 2009); water injection seepage property of pore and fracture (Zhang et al. 2020b); water inrush from activated collapse pillar (Liu and Xiong 2007); nonlinear flow via fracture network (Li et al. 2016a; Liu et al. 2016a, b) | ||
COMSOL | Flow regime and hydraulic property of coal measure strata (Chen et al. 2018; Yao et al. 2012); pore pressure and water seepage in coal seam floor with longwall progress (Zhang and Fan 2014); aquifer dewatering due to failure (Li et al. 2018); water inrush around geological structures (Zhu and Wei 2011) | ||
RS2/RS3 | Water seepage on panel roof and its control (Bai and Tu 2016; Sasaoka et al. 2014; Smith et al. 2019); dewatering of shallow unconfined aquifer (Islam et al. 2016) | ||
RFPA | Pattern and pathway of confined water inflow (Yin et al. 2016); groundwater inflow from floor aquifer (Yang et al. 2007) and roof strata (Zhang et al. 2009b) based on flow-stress-damage coupling model; groundwater inrush due to fault activation (Li et al. 2011) | ||
Groundwater hydraulics and hydrological modelling | COSFLOW | Strata movement and permeability variation (Adhikary and Guo 2015; Chen and Guo 2008; Chen and Hu 2009; Khanal et al. 2019b); distribution of stress and pore pressure fields after mining (Guo et al. 2008; Wilkins and Qu 2020) | |
FEFLOW | Hydraulic head regime over a geothermal reservoir area (Hegde et al. 2013); seepage field in coal measure aquifer and its evolution (Dong et al. 2012); prediction of groundwater level variation with pre-mining drainage (Hang et al. 2009) and longwall mining (Vukelić et al. 2016) | ||
Finite difference method (FDM) | GMS | Groundwater flow regime and water level variation due to longwall mining (Wang et al. 2018b; Wu et al. 2019; Xu et al. 2020) | |
MODFLOW | Hydrological impacts of longwall mining on groundwater flow (Booth and Greer 2011; Dhakate et al. 2019; Newman et al. 2016, 2017; Xu et al. 2020); water seepage into mine openings (Zaidel et al. 2010) | ||
Hybrid modelling | FLAC-UDEC | Lab-scale fracture propagation in response to external loading (Mahabadi et al. 2010) | |
FLAC-3DEC | Modelling of coal measure strata, 3DEC for rock fracturing and FLAC3D for stress and rock permeability variation (Wang et al. 2015) | ||
FLAC-PFC | Lab-scale rock fracturing under loading (Adhikary et al. 2017; Cai et al. 2007); surrounding rock mass failure due to excavation (Wei and Ren 2012); water seepage and control (Su et al. 2020); water-resisting property of aquitard under different mining methods (Zhao 2020) |
Representation of rock fracturing and fluid flow can be different between numerical techniques. As compared in Fig. 17, discontinuum model is composed of discretised rigid particle or deformable blocky entities, and in-between physical contacts constituting joints along which adjacent entities can separate and shear to form rock fractures, Fig. 17c. Particle diameters and joints are fabricated in model construction, usually referencing the frequency and orientation of discontinuities captured in lab tests, Fig. 17a, or field measurements, Fig. 17b. Continuum models do not exhibit fractures, instead indicating permeable areas via user-defined criteria such as plastic states (Whittles et al. 2006; Zhang et al. 2019), pore pressure enhancement (Guo et al. 2008; Smith et al. 2019), and damage evolution (Li et al. 2011).
Rock fractures and their discontinuum and continuum representation in numerical modelling. a Shows a fractured mudstone specimen after UCS test b Exhibits fractured cores collected from a borehole at Springvale Mine, Australia (Corbett 2022) c and d Indicate the difference between discontinuum and continuum models in representing rock fracturing and fluid flow
On a panel scale, the continuum method has been applied to investigate overlying rock failure and potential flow pathways. Figure 18a shows three FLAC2D models established in a parametric study (Gale 2008). The modelling obtains that (i) the fracture network is characterised by bedding plane shear and tensile fracturing or mobilisation of existing joints in tensile zones, (ii) flow pathways are direct in the caved zone and gradually tortuous with elevation, and (iii) the height of fracturing, fracture connectivity, and hydraulic conductivity (assuming Darcian flow) increase as greater subsidence occurs. Discrete Element Method (DEM) modelling has increasing applications due to its intrinsic feature in exhibiting rock fracturing. As shown in Fig. 18b, Adhikary et al. simulated a panel extraction within Lidsdale/Lithgow Seam at Springvale Mine and induced continuous fracturing by means of the PFC2D software (Adhikary et al. 2017, 2020). The model incorporated the Discrete Fracture Network (DFN) technology to configure high-angle interfaces along the immediate roof for caving initiation. It was identified that the height of connected fractures was 260 m above the goaf centre and 290 m above the abutment, suggesting no complete groundwater drainage zone forming above the mined panel (Adhikary et al. 2017).
Continuum and discontinuum modellings of panel extraction and induced rock failure and fracture propagation. a Exhibits three FLAC2D models for longwall mining within Whybrow Seam at Hunter Valley, Australia b Exhibits a PFC2D model of panel LW411 (3.2 m in mining height, 315 m in panel width, and 350 m in depth of cover) at Springvale Mine, Australia. The DEM model is 457 by 450 m and contains approximately 503 thousand rigid particles, adopting the built-in ‘flat-joint’ model for physical contacts
Given these present applications, the paper proposes three elements that should be concerned in future panel- to mine-scale modelling.
Size effect in rock parameter determination
Lab test on intact rocks is a major source of strength parameter acquisition, typically using standardised cylinders or cubes with edge lengths less than 200 mm. The intact rock strength should be upscaled with respect to the geometry of rock matrices in panel- to mine-scale models considering the size effect; otherwise, the rock deformation and failure are underestimated (Baghbanan 2008; Castelli et al. 2003; Follington and Isaac 1990; Kong et al. 2021). There have been several equations for rock strength upscaling. Hoek and Brown established an equation \({\sigma }_{m}={\sigma }_{r}{\left(50/{d}_{m}\right)}^{0.18}\) to determine rock mass UCS (\({\sigma }_{m}\)) via cylindrical specimen UCS (\({\sigma }_{r}\)) and rock mass geometry (\({d}_{m}\)) (Hoek and Brown 1980). Yoshinaka et al. proposed \({\sigma }_{m}={\sigma }_{r}{\left(58.1/{d}_{m}\right)}^{k}\), \(k\) from 0.1 to 0.3 for homogenous hard rocks and 0.3 to 0.9 for weathered or microflawed rocks (Yoshinaka et al. 2008). The paper updates a new lab-to-field relationship in rock UCS (Eq. (12)) and elastic modulus (Eq. (13)), by combining coal measure rock (mass) strength collected from labs and mines with some reported data points (Bieniawski 1992; Gale 1992; Pratt et al. 1972; Yoshinaka et al. 2008), as exhibited in Fig. 19.
where \({\sigma }_{m}\) and \({E}_{m}\) are the UCS and elastic modulus of rock matrix, and \({\sigma }_{r}\) and \({E}_{r}\) the UCS and elastic modulus of rock specimen. \({d}_{m}\) is the size of rock matrix, taking the length of the edge normal to the major loading. \({d}_{r}\) is the diameter of cylindrical specimens, or edge length of cubes, using a typical value of 50 mm if no test data available. Equation (12) enlarges the upper limit of UCS upscaling to about 8 m by incorporating the strength of intact core of chain pillars. The size effect also influences the stiffness and strength of physical contacts in DEM models. Such rock joint parameters can be calibrated by simulating uniaxial and triaxial compression tests and fitting the model results to measured values. There have been some workflows for the calibration (Christianson et al. 2006; Gao and Stead 2014).
Rock fracture representation via DFN technology
The DFN technology can be used to fabricate rock joints and identify connected fractures, having two major types of application in fracture propagation and fracture flow assessments:
Pure DFN configuration without mechanical calculation. This method constructs a fracture network by interpreting fracture intensity, orientation, length, and aperture data. Fracture flow is calculated by setting hydraulic boundaries, sometimes accompanied by rock matrix percolation. The pore pressure effect on fracture aperture is neglected. This method is used for understanding fluid flow through jointed rock masses (Liu et al. 2016b; Min et al. 2004; Zoorabadi 2014)
DFN regeneration after mechanical calculation. In model construction, a zero-aperture DFN is built according to fracture statistics to fabricate jointed rock masses. Aperture generates with rock matrices displacing and deforming in mechanical calculations. A new DFN is generated by mapping open contacts, hence forming flow pathways with apertures inherited from contacts. This method has been used in roadway- to panel-scale models (Adhikary et al. 2017, 2020; Khanal et al. 2019a; Poulsen et al. 2018; Wang et al. 2020a; Wei et al. 2022).
The applications indicate that, in model construction, the DFN configuration is controlled by fracture statistics determined from measured fracture attributes. Geological mapping helps detect the trace length, frequency, and orientation, for example the LiDAR system for fractures along outcrops or walls of underground excavations (Vlachopoulos et al. 2020), and ground penetrating radar (GPR) for subsurface fractures (Hawkins et al. 2017; Molron et al. 2020). Borehole geophysical logging uses acoustic and optical scanners to obtain the frequency, orientation, and position of bedrock fractures that intersect the borehole (Brown and Walsh 2022; Corbett 2022; Heritage et al. 2022a; Walsh et al. 2022).
Measured data can be statistically interpreted to determine the probability distribution function (PDF) of each attribute. Lognormal, exponential, gamma, and power-law functions are frequently used for rock fracture statistics (Bonnet et al. 2001). Figure 20 shows a workflow by which the fracture length measured by authors in Ordos Basin coalfields is identified to prefer a lognormal distribution. A primary assessment is first conducted according to the skewness (\({S}_{k}\)) and shape of the probability curve. \({S}_{k}\) measures the asymmetry of the probability distribution of variables (\(X\)) with respect to their mean value (\(\overline{X }\)), expressed as:
where \(n\) is the total number of variables, and \(s\) the standard deviation. \({S}_{k}\) is 0 in normal and 2 in exponential but not a fixed value in lognormal, gamma, and power law distributions. The distribution function should feature a long tail (the Pareto principle), a single mode of a bell-shaped curve, and a bulge left of the peak in power law, lognormal, and gamma distributions, respectively. The primary assessment suggests higher likelihoods of lognormal and gamma distributions. A further data fitting identifies the best-fit distribution to be lognormal with a mean (\(\mu\)) of 0.91 and standard deviation (\(\sigma\)) of 0.44. The lognormal function is validated via K-S test. The workflow is also available for interpreting other fracture attributes.
Model verification and hybrid modelling
Panel- to mine-scale models should have minor errors, usually no more than 10% (Bahrani and Hadjigeorgiou 2018; Gao et al. 2015; Zhang et al. 2020a, 2021), with respect to measurements in chain pillar compression, height of fracturing, and incremental subsidence. This ensures an accurate representation of the mechanical behaviour of rocks from the seam horizon to ground surface. Hydrogeological models should verify groundwater hydraulics when fluid flow is incorporated, for example piezometric variation (Adhikary and Guo 2015; Adhikary et al. 2017; Booth 2002; David 2018; Khanal et al. 2019b; Rapantová et al. 2007, 2012) and mine inflow (Adhikary et al. 2017; Guo et al. 2008; Khanal et al. 2019a).
Different techniques can be combined to form a hybrid modelling method. FDM-DEM coupling has been applied to investigate rock mechanics (Adhikary et al. 2017; Cai et al. 2007; Mahabadi et al. 2010) and field-scale rock mass depressurisation (Wang et al. 2015; Zhao 2020), reducing the long computation time in single DEM simulation. PFC2D and COSFLOW were combined to assess the fluid flow environment above mined panels (Adhikary et al. 2017, 2020; Khanal et al. 2019a), where COSFLOW served for complementing and benchmarking the strata permeability enhancement obtained from PFC2D models. Such hybrid modelling helps address the limitation of geomechanical models in representing mine- to coalfield-scale piezometric variation.
The paper identifies knowledge gaps that should be addressed for obtaining more accurate mine dewatering assessment. Quantification of fracture network interconnectivity and description of fracture flow should be developed at first as they determine whether the hydraulic gradient and mine inflow rate can be accurately captured. In addition, identification of the regional hydrology needs an increased attention, especially for valley terrains, because the induced drawdown and post-mining recovery are influenced by the hydrology itself in regard to (i) aquifer occurrence and water abundance, (ii) groundwater circulation pattern, and (iii) interaction between surface and subsurface waters. The effect of varying retreat rates and stress conditions on continuous fracturing and dewatering can be investigated later, accompanied by the contribution of mineral infillings to fracture closure and long-term water recovery.
Figure 21 presents a framework and lab- to field-scale topics for future research. Methodologically, continuous data collection and field-scale numerical modelling should be enhanced as key technical support, requiring that
A mine groundwater database with HoF and piezometric drawdown as key data should be established, effectively managed, and continuously updated from long-term monitoring at sequential panels.
Empirical predictions should take into account at least panel geometry, depth of cover, and rock strength factors as predictors, keeping a constant revision with the database renewal.
Numerical models should be able to indicate continuous fracturing and hydraulic head decline due to mining, with geological to hydrological verifications to improve the model reliability for parametric studies.
In the first stage, measured fractures are statistically interpreted and reproduced by the DFN technology into field-scale numerical models. Induced fractures that are mutually interconnected and directly or indirectly connected to the goaf should be identified with an appropriate logic and its algorithmic implementation. Typical morphology of intersected fractures within the network can be captured to provide reference for fabricating flow chambers for fracture flow tests in lab. The correlation between hydraulic gradient (J) and flow rate (Q) can be obtained from lab tests and used to benchmark CFD models to investigate the contribution of aperture, frequency, and intersection to head loss and flow nonlinearity via sensitivity analysis. A semi-empirical model should be established to describe fluid flow through large-aperture and interconnected fractures.
In the second stage, hydro-mechanical model should be established by incorporating the fracture flow description and verified against measured water table decline and rock mass conductivity. Two series of parametric studies are needed, including one based on packer testing simulation to investigate the correlation of rock mass conductivity and head drop to fracturing degree, and another based on mine-scale models to quantify the response of continuous fracturing and dewatering to different longwall parameters. HoF data recorded in the database and the correlation between HoF and longwall parameters are combined to develop a new empirical prediction equation. The results of parametric studies and empirical interpretations constitute a methodology to evaluate mine hydrology and groundwater hydraulics variation due to longwall mining. Cases updated into the database can be used to revise the methodology and keep it reliable for mine groundwater depressurisation assessments.
The paper investigates the factors and mechanism of mine dewatering and identifies knowledge gaps to be addressed in future research to gain better understanding and assessment of groundwater depressurisation due to longwall mining.
It is identified that the piezometric drawdown as a direct result of sequential panel extractions can further affect subsurface and surface hydrologic features at mine sites. Statistical analyses of 422 HoF datasets collected from 168 mines indicate that mining height has a greater contribution to the groundwater impact than panel width and depth of cover. The dewatering impact is more pronounced in cases where (i) the seam is extracted using LTCC mining method and (ii) panel interaction is intensified due to narrow pillar configuration within a seam or thin and incompetent interlayers between adjacent seams. Geological and hydrogeological factors such as in situ stress, rock strength and clay mineral infillings, fault structures, valley terrains, and surface and subsurface water interactions also contribute to the longwall impact on groundwater and determine the likelihood and duration of water recovery in the long term.
Interconnected fracture network formed by continuous fracturing dominate the dewatering mechanism, providing steep hydraulic gradients and smooth flow pathways for overlying waters draining to the mined-out area underneath. The fluid flow environment may involve high-Reynolds flow and non-Darcian turbulence due to high fracture frequency and large fracture aperture, complicated than those that have been considered in conventional fracture flow models. In this context, fracture connectivity quantification and fracture flow description should be developed in future research. Continuous field data collection, data management and interpretation, and panel- to mine-scale numerical modelling should be enhanced as technical support. The paper designs a research framework that contains lab- to field-scale topics to narrow the knowledge gaps and achieve improved understanding and assessment of groundwater depressurisation due to longwall mining.
[1] | Ackman TE, Jones JR (1991) Methods to identify and reduce potential surface stream water losses into abandoned underground mines. Environ Geol Water Sci 17(3):227–232. https://doi.org/10.1007/BF01701703 |
[2] | Adhikary D, Guo H (2015) Modelling of longwall mining-induced strata permeability change. Rock Mech Rock Eng 48(1):345–359. https://doi.org/10.1007/s00603-014-0551-7 |
[3] | Adhikary D, Poulsen B, Khanal M (2017) Assessment of longwall mining induced connective fracturing of overburden strata. ACARP Project C24020, CSIRO, Australia |
[4] | Adhikary D, Poulsen B, Wilkins A (2020) Assessment of longwall mining induced connective fracturing. ACARP Project C27045, CSIRO, Australia |
[5] | Ali A, Strezov V, Davies P, Wright I (2017) Environmental impact of coal mining and coal seam gas production on surface water quality in the Sydney basin, Australia. Environ Monitor Assess 189(8):408–423. https://doi.org/10.1007/s10661-017-6110-4 |
[6] | Artugyan L, Urdea P (2014) Groundwater drainage monitoring and karst terrain analysis using spontaneous potential (SP) in Anina Mining Area (Banat Mountains, Romania)—preliminary study. In: Porceedings of Karst Without Boundaries Conference, Trebinje, Bosnia and Herzegovina, pp 11–15 |
[7] | Baghbanan A (2008) Scale and stress effects on hydro-mechanical properties of fractured rock masses. Dissertation, KTH Royal Institute of Technology |
[8] | Bahrani N, Hadjigeorgiou J (2018) Influence of stope excavation on drift convergence and support behavior: insights from 3D continuum and discontinuum models. Rock Mech Rock Eng 51(8):2395–2413. https://doi.org/10.1007/s00603-018-1482-5 |
[9] | Bai L, Wang Y, Zhang K, Yang Y, Bao K, Zhao J, Li X (2022) Spatial variability of soil moisture in a mining subsidence area of northwest China. Int J Coal Sci Technol 9(1):64. https://doi.org/10.1007/s40789-022-00517-w |
[10] | Bai Q, Tu S (2016) Failure analysis of a large span longwall drift under water-rich roofs and its control techniques. Eng Fail Anal 67:15–32. https://doi.org/10.1016/j.engfailanal.2016.05.028 |
[11] | Barbato J, Hebblewhite B, Mitra R, Mills K (2016a) Prediction of horizontal movement and strain at the surface due to longwall coal mining. Int J Rock Mech Min Sci 84:105–118. https://doi.org/10.1016/j.ijrmms.2016.02.006 |
[12] | Barbato J, Hebblewhite B, Mitra R, Mills K (2016b) Review of horizontal surface movements due to longwall coal mining using numerical modelling. In: Proceedings of the 16th Coal Operators’ Conference, Wollongong, NSW, Australia, pp 213–223 |
[13] | Bay RR (1968) The hydrology of several peat deposits in northern Minnesota, U.S.A. In: Proceedings of the Third International Peat Congress, Quebec, Canada, pp 212–218 |
[14] | Bian Z, Lei S, Inyang HI, Chang L, Zhang R, Zhou C, He X (2009) Integrated method of RS and GPR for monitoring the changes in the soil moisture and groundwater environment due to underground coal mining. Environ Geol 57(1):131–142. https://doi.org/10.1007/s00254-008-1289-x |
[15] | Bieniawski Z (1992) A method revisited: coal pillar strength formula based on field investigations. In: Proceedings of the Workshop on Coal Pillar Mechanics and Design, Pittsburgh, US, pp 158–165 |
[16] | Bonnet E, Bour O, Odling NE, Davy P, Main I, Cowie P, Berkowitz B (2001) Scaling of fracture systems in geological media. Rev Geophys 39(3):347–383. https://doi.org/10.1029/1999RG000074 |
[17] | Booth CJ (1986) Strata-movement concepts and the hydrogeological impact of underground coal mining. Groundwater 24(4):507–515. https://doi.org/10.1111/j.1745-6584.1986.tb01030.x |
[18] | Booth CJ (1999) Recovery of groundwater levels after longwall mining. In: Proceedings of the IMWA International Congress, Sevilla, Spain, pp 35–40 |
[19] | Booth CJ (2002) The effects of longwall coal mining on overlying aquifers. Geol Soc Lond Spec Publ 198(1):17–45. https://doi.org/10.1144/GSL.SP.2002.198.01.02 |
[20] | Booth CJ (2006) Groundwater as an environmental constraint of longwall coal mining. Environ Geol 49(6):796–803. https://doi.org/10.1007/s00254-006-0173-9 |
[21] | Booth CJ (2007) Confined-unconfined changes above longwall coal mining due to increases in fracture porosity. Environ Eng Geosci 13(4):355–367. https://doi.org/10.2113/gseegeosci.13.4.355 |
[22] | Booth CJ, Bertsch LP (1999) Groundwater geochemistry in shallow aquifers above longwall mines in Illinois, USA. Hydrogeol J 7(6):561–575. https://doi.org/10.1007/s100400050229 |
[23] | Booth CJ, Curtiss AM, Demaris PJ, Bauer RA (2000) Site-specific variation in the potentiometric response to subsidence above active longwall mining. Environ Eng Geosci 6(4):383–394. https://doi.org/10.2113/gseegeosci.6.4.383 |
[24] | Booth CJ, Curtiss AM, Demaris PJ, Rossendaal DJV (1999) Anomalous increases in piezometric levels in advance of longwall mining subsidence. Environ Eng Geosci 5(4):407–417. https://doi.org/10.2113/gseegeosci.V.4.407 |
[25] | Booth CJ, Greer CB (2011) Application of MODFLOW using TMR and discrete-step modification of hydraulic properties to simulate the hydrogeologic impact of longwall mining subsidence on overlying shallow aquifers. In: Proceedings of Mine Water - Managing the Challenges, Aachen, Germany, pp 211–216 |
[26] | Booth CJ, Spande ED (1992) Potentiometric and aquifer property changes above subsiding longwall mine panels Illinois basin coalfield. Groundwater 30(3):362–368. https://doi.org/10.1111/j.1745-6584.1992.tb02004.x |
[27] | Booth CJ, Spande ED, Pattee CT, Miller JD, Bertsch LP (1998) Positive and negative impacts of longwall mine subsidence on a sandstone aquifer. Environ Geol 34(2–3):223–233. https://doi.org/10.1007/s002540050274 |
[28] | Brown R, Parizek R (1971) Shallow ground-water flow systems beneath strip and deep coal mines at two sites, Clearfield County, Pennsylvania. Special Report SR-84, Department of Environmcntal Resources of the Commonwealth of Pennsylvania |
[29] | Brown S, Minchin W (2018) Dendrobium Mine—end of groundwater assessment for Longwall 13 (Area 3B). Report D18305, HGEO Pty Ltd |
[30] | Brown S, Walsh R (2022) The height of connected fracturing above a longwall coal mine: a field and isotopic investigation. In: Proceedings of the 11th Conference on Mine Subsidence, Pokolbin, NSW, Australia, pp 135–140 |
[31] | China Coal Industry Bureau (2000) Measuring method on height of water flowing fractured zone using losses of drilling fluid. Coal industry standard of the People’s Republic of China MT/T 865-2000, pp 1–3 |
[32] | Butler JJ (2019) The design, performance, and analysis of slug tests, 2nd edn. CRC Press, Boca Raton |
[33] | Cai M, Kaiser PK, Morioka H, Minami M, Maejima T, Tasaka Y, Kurose H (2007) FLAC/PFC coupled numerical simulation of AE in large-scale underground excavations. Int J Rock Mech Min Sci 44(4):550–564. https://doi.org/10.1016/j.ijrmms.2006.09.013 |
[34] | Castelli M, Saetta V, Scavia C (2003) Numerical study of scale effects on the stiffness modulus of rock masses. Int J Geomech 3(2):160–169. https://doi.org/10.1061/(ASCE)1532-3641(2003)3:2(160) |
[35] | CCRI (1981) The law of surface movement and overburden damage in coal mine and its application. China Coal Industry Press, Beijing |
[36] | Chabukdhara M, Singh OP (2016) Coal mining in northeast India: an overview of environmental issues and treatment approaches. Int J Coal Sci Technol 3(2):87–96. https://doi.org/10.1007/s40789-016-0126-1 |
[37] | Chatterjee R, Tarafder G, Paul S (2010) Groundwater quality assessment of Dhanbad district, Jharkhand, India. Bull Eng Geol Env 69(1):137–141. https://doi.org/10.1007/s10064-009-0234-x |
[38] | Chen S, Guo H (2008) Numerical simulation of bed separation development and grout injecting into separations. Geotech Geol Eng 26(4):375–385. https://doi.org/10.1007/s10706-008-9174-7 |
[39] | Chen S, Hu W (2009) A comprehensive study on subsidence control using COSFLOW. Geotech Geol Eng 27(3):305–314. https://doi.org/10.1007/s10706-008-9230-3 |
[40] | Chen Y, Selvadurai APS, Liang W (2018) Computational modelling of groundwater inflow during a longwall coal mining advance: a case study from the Shanxi Province, China. Rock Mech Rock Eng 52(3):917–934. https://doi.org/10.1007/s00603-018-1603-1 |
[41] | Christianson M, Board M, Rigby D (2006) UDEC simulation of triaxial testing of lithophysal tuff. In: Proceedings of the 41st US Symposium on Rock Mechanics, Golden, Colorado, pp 1–8 |
[42] | Coe CJ, Stowe SM (1984) Evaluating the impact of longwall mining on the hydrologic balance. In: Proceedings of Graves. Symp on Surface Mining, Hydrology, Sedimentology, and Reclamation, Lexington, Kentucky, pp 395–403 |
[43] | Corbett P (2022) Groundwater system response to longwall mining at Springvale Mine. In: Proceedings of the 11th Conference on Mine Subsidence, Pokolbin, NSW, Australia, pp 153–170 |
[44] | Cui F, Yang Y, Lai X, Jia C, Shan P (2020) Experimental study on the effect of advancing speed and stoping time on the energy release of overburden in an upward mining coal working face with a hard roof. Sustainability 12(37):1–23. https://doi.org/10.3390/su12010037 |
[45] | Dai H, Li P, Marzhan N, Yan Y, Yuan C, Serik T, Guo J, Zhakypbek Y, Seituly K (2022) Subsidence control method by inversely-inclined slicing and upward mining for ultra-thick steep seams. Int J Min Sci Technol 32(1):103–112. https://doi.org/10.1016/j.ijmst.2021.10.003 |
[46] | Darmody RG, Bauer R, Barkley D, Clarke S, Hamilton D (2014) Agricultural impacts of longwall mine subsidence: the experience in Illinois, USA and Queensland, Australia. Int J Coal Sci Technol 1(2):207–212. https://doi.org/10.1007/s40789-014-0026-1 |
[47] | David K (2018) Groundwater changes in a sedimentary sequence associated with underground mining—a multidisciplinary approach. Dissertation, University of New South Wales |
[48] | Dhakate R, Modi D, Rao VVSG (2019) Impact assessment of coal mining on river water and groundwater and its interaction through hydrological, isotopic characteristics, and simulation flow modeling. Arab J Geosci 12(1):1–8. https://doi.org/10.1007/s12517-018-4110-5 |
[49] | Ditton S (2014) Subsurface fracture zone assessment above the proposed Springvale and Angus Place Mine extension project area longwalls. DgS Report SPV-003/7b, Ditton Geotechnical Services, Australia |
[50] | Ditton S, Merrick NP (2014) A new sub-surface fracture height prediction model for longwall mines in the NSW coalfields. In: Proceedings of Sydney Basin Symposium, Newcastle, NSW, pp 135–136 |
[51] | Dixon DY, Rauch HW (1990) The impact of three longwall coal mines on streamflow in the Appalachian Coalfield. In: Proceedings of the 9th International Conference on Ground Control in Mining, Morgantown, West Virginia, USA, pp 169–182 |
[52] | Dong D, Sun W, Xi S (2012) Optimization of mine drainage capacity using FEFLOW for the No. 14 coal seam of China’s Linnancang coal mine. Mine Water Environ 31(4):353–360. https://doi.org/10.1007/s10230-012-0205-5 |
[53] | Du F, Gao R (2017) Development patterns of fractured water-conducting zones in longwall mining of thick coal seams—a case study on safe mining under the Zhuozhang River. Energies 10(11):1856. https://doi.org/10.3390/en10111856 |
[54] | Fan G, Chen M, Zhang D, Wang Z, Zhang S, Zhang C, Li Q, Cao B (2018) Experimental study on the permeability of weakly cemented rock under different stress states in triaxial compression tests. Geofluids 2018:1–9. https://doi.org/10.1155/2018/9035654 |
[55] | Fan G, Zhang D (2015) Mechanisms of aquifer protection in underground coal mining. Mine Water Environ 34(1):95–104. https://doi.org/10.1007/s10230-014-0298-0 |
[56] | Fan G, Zhang D, Zhang S, Zhang C (2017) Assessment and prevention of water and sand inrush associated with coal mining under a water-filled buried gully: a case study. Mine Water Environ 37(3):565–576. https://doi.org/10.1007/s10230-017-0487-8 |
[57] | Fan G, Zhang D, Zhang S, Zhao Q, Yu W, Liang S (2019a) Influence of stress and crack patterns on the sensitive characteristics of fissure sandstone permeability under hydromechanical coupling. Appl Sci 9(4):641–667. https://doi.org/10.3390/app9040641 |
[58] | Fan G, Zhang D, Zhou L (2011) Suitable layout of gate roads related to slice mining in an ultra-thick unstable coal seam. Min Sci Technol (china) 21(4):563–566. https://doi.org/10.1016/j.mstc.2011.06.017 |
[59] | Fan G, Zhang S, Cao B, Zhang D, Zhang C (2020) Impact of mine panel size on hydraulic permeability of weakly cemented strata. Sustainability 12(6):1–17. https://doi.org/10.3390/su12062396 |
[60] | Fan G, Zhang S, Zhang D, Zhang C, Chen M, Li Q (2019b) An index of aquiclude destabilization for mining-induced roof water inrush forecasting: a case study. Water 11(10):2170–2184. https://doi.org/10.3390/w11102170 |
[61] | Fan L, Ma L, Yu Y, Wang S, Xu Y (2019c) Water-conserving mining influencing factors identification and weight determination in northwest China. Int J Coal Sci Technol 6(1):95–101. https://doi.org/10.1007/s40789-018-0233-2 |
[62] | Fan L, Ma X (2018) A review on investigation of water-preserved coal mining in western China. Int J Coal Sci Technol 5(4):411–416. https://doi.org/10.1007/s40789-018-0223-4 |
[63] | Fan Z, Fan G, Zhang D, Zhang L, Zhang S, Liang S, Yu W (2021a) Optimal injection timing and gas mixture proportion for enhancing coalbed methane recovery. Energy 222:119880. https://doi.org/10.1016/j.energy.2021.119880 |
[64] | Fan Z, Zhang D, Fan G, Zhang L, Zhang S, Liang S, Yu W, Gao S, Guo W (2021b) Non-Darcy thermal-hydraulic-mechanical damage model for enhancing coalbed methane extraction. J Nat Gas Sci Eng 93:104048. https://doi.org/10.1016/j.jngse.2021.104048 |
[65] | Follington IL, Isaac AK (1990) Failure zone development above longwall panels. Min Sci Technol 10(2):103–116. https://doi.org/10.1016/0167-9031(90)90122-9 |
[66] | Fossen H (2016) Structural geology. Cambridge University Press, Cambridge |
[67] | Frimpter MH, Maevsky A (1979) Geohydrologic impacts of coal development in the Narragansett Basin. Massachusetts and Rhode Island. U. S. Government Printing Office, Washington |
[68] | Frost CD, Pearson BN, Ogle KM, Heffern EL, Lyman RM (2002) Sr isotope tracing of aquifer interactions in an area of accelerating coal-bed methane production, Powder River Basin. Wyoming Geol 30(10):923–926. https://doi.org/10.1130/0091-7613(2002)030%3c0923:SITOAI%3e2.0.CO;2 |
[69] | Gale W (2008) Aquifer inflow prediction above longwall panels. ACARP Report C13013, SCT Operations, Australia |
[70] | Gale WJ (1992) A pillar design approach. In: Proceedings of the Workshop on Coal Pillar Mechanics and Design, Pittsburgh, US, pp 188–195 |
[71] | Gale WJ (2005) Application of computer modelling in the understanding of caving and induced hydraulic conductivity about longwall panels. In: Proceedings of the 2005 Coal Operators’ Conference, Wollongong, NSW, Australia, pp 11–15 |
[72] | Galvin JM (2016) Ground engineering—principles and practices for underground coal mining. Springer, Berlin |
[73] | Galvin JM, Hebblewhite BK, Salamon MDG (1999) Coal pillar strength determinations for Australian and South African mining conditions. In: Proceedings of the Second International Workshop on Coal Pillar Mechanics and Design, Pittsburgh, US, pp 63–71 |
[74] | Gao F, Stead D (2014) The application of a modified Voronoi logic to brittle fracture modelling at the laboratory and field scale. Int J Rock Mech Min Sci 68:1–14. https://doi.org/10.1016/j.ijrmms.2014.02.003 |
[75] | Gao F, Stead D, Kang H (2015) Numerical simulation of squeezing failure in a coal mine roadway due to mining-induced stresses. Rock Mech Rock Eng 48(4):1635–1645. https://doi.org/10.1007/s00603-014-0653-2 |
[76] | Ghabraie B, Ren G, Smith JV (2017) Characterising the multi-seam subsidence due to varying mining configuration, insights from physical modelling. Int J Rock Mech Min Sci 93:269–279. https://doi.org/10.1016/j.ijrmms.2017.02.001 |
[77] | Gill DR (2000) Hydrogeologic analysis of streamflow in relation to underground mining in northern West Virginia. Dissertation, West Virginia University |
[78] | Guo H, Adhikary DP, Craig MS (2008) Simulation of mine water inflow and gas emission during longwall mining. Rock Mech Rock Eng 42(1):25–51. https://doi.org/10.1007/s00603-008-0168-9 |
[79] | Guo H, Adhikary DP, Gabeva D (2007) Hydrogeological response to longwall mining. ACARP Report C14033, CSIRO, Australia |
[80] | Guo Q, Yang Y, Xiong X (2016) Using hydrologic simulation to identify contributions of climate change and human activity to runoff changes in the Kuye river basin, China. Environ Earth Sci 75(5):1–10. https://doi.org/10.1007/s12665-016-5280-7 |
[81] | Guo Q, Yang YS, Su N, Li JL, Wang XY (2019a) Impact assessment of climate change and human activities on runoff variation in coal mining watershed, NW China. Water Resour 46(6):871–882. https://doi.org/10.1134/S0097807819060058 |
[82] | Guo W, Zhao G, Lou G, Wang S (2019b) A new method of predicting the height of the fractured water-conducting zone due to high-intensity longwall coal mining in China. Rock Mech Rock Eng 52(8):2789–2802. https://doi.org/10.1007/s00603-018-1567-1 |
[83] | Hang Y, Zhang G, Yang G (2009) Numerical simulation of dewatering thick unconsolidated aquifers for safety of underground coal mining. Min Sci Technol 19(3):312–316. https://doi.org/10.1016/S1674-5264(09)60058-2 |
[84] | Haseltine R (1892) The wasteful methods being practiced in mining coal in the state. Ohio Min J 21:83–101 |
[85] | Hawkins AJ, Becker MW, Tsoflias GP (2017) Evaluation of inert tracers in a bedrock fracture using ground penetrating radar and thermal sensors. Geothermics 67:86–94. https://doi.org/10.1016/j.geothermics.2017.01.006 |
[86] | He X, Zhao Y, Zhang C, Han P (2020) A model to estimate the height of the water-conducting fracture zone for longwall panels in Western China. Mine Water Environ 39(4):823–838. https://doi.org/10.1007/s10230-020-00726-2 |
[87] | Hebblewhite B (2009) Outcomes of the independent inquiry into impacts of underground coal mining on natural features in the Southern Coalfield—an overview. In: Proceedings of the 2009 Coal Operators’ Conference, Wollongong, NSW, Australia, pp 113–124 |
[88] | Hebblewhite B (2020) Fracturing, caving propagation and influence of mining on groundwater above longwall panels—a review of predictive models. Int J Min Sci Technol 30(1):49–54. https://doi.org/10.1016/j.ijmst.2019.12.001 |
[89] | Hebblewhite B, Galvin J, Mackie C, West R, Collins D (2008) Impacts of underground coal mining on natural features in the southern coalfield: strategic review. Report to NSW Department of Planning |
[90] | Hegde C, Rühaak W, Sass I (2013) Evaluation of modelling of flow in fractures. In: Proceedings of International Conference on Advances in Civil Engineering, Pulau Pinang, Malaysia, pp 116–125 |
[91] | Heiland J (2003) Laboratory testing of coupled hydro-mechanical processes during rock deformation. Hydrogeol J 11(1):122–141. https://doi.org/10.1007/s10040-002-0236-2 |
[92] | Heiland J, Raab S (2001) Experimental investigation of the influence of differential stress on permeability of a lower permian (rotliegend) sandstone deformed in the brittle deformation field. Phys Chem Earth Part A Solid Earth Geod 26(1–2):33–38. https://doi.org/10.1016/S1464-1895(01)00019-9 |
[93] | Heras MM-D, Nicolau JM, Espigares T (2008) Vegetation succession in reclaimed coal-mining slopes in a Mediterranean-dry environment. Ecol Eng 34(2):168–178. https://doi.org/10.1016/j.ecoleng.2008.07.017 |
[94] | Heritage Y, Blacka B, James D (2022a) Definition of measured hydraulic conductivity zones above a longwall panel, United Colliery. In: Proceedings of the 11th Conference on Mine Subsidence, Pokolbin, NSW, Australia, pp 141–151 |
[95] | Heritage Y, Blacka B, James D (2022b) Measurement of the caved zone above a longwall panel, United Colliery. In: Proceedings of the 2022 Coal Operators' Conference, Wollongong, NSW, Australia, pp 66–74 |
[96] | Heritage Y, Gale W, Rippon A (2017) Connectivity of mining induced fractures below longwall panels: a modelling approach. In: Proceedings of the 2017 Coal Operators’ Conference, Wollongong, NSW, Australia, pp 41–50 |
[97] | Hill JG, Price DR (1983) The impact of deep mining on an overlying aquifer in western Pennsylvania. Groundw Monit Remediat 3(1):138–143. https://doi.org/10.1111/j.1745-6592.1983.tb00875.x |
[98] | Hoek E, Brown E (1980) Underground excavations in rock. CRC Press, Boca Raton |
[99] | Hou E, Wen Q, Ye Z, Chen W, Wei J (2020) Height prediction of water-flowing fracture zone with a genetic-algorithm support-vector-machine method. Int J Coal Sci Technol 7(4):740–751. https://doi.org/10.1007/s40789-020-00363-8 |
[100] | Huang Q, Wu B, Cheng W, Lei B, Shi H, Chen L (2018) Investigation of permeability evolution in the lower slice during thick seam slicing mining and gas drainage: a case study from the Dahuangshan coalmine in China. J Nat Gas Sci Eng 52:141–154. https://doi.org/10.1016/j.jngse.2018.01.036 |
[101] | Hutcheson SM, Kipp JA, Dinger JS, Sendlein LVA, Carey DI (2000) Effects of longwall mining on hydrology, Leslie County, Kentucky part 2: during-mining conditions. Kentucky Geological Survey Report of Investigations, University of Kentucky |
[102] | Cac III, Goode DJ, Bartles MD, Risser DW, Galeone DG (2014) Surface water and groundwater interactions in an extensively mined watershed, upper Schuylkill River, Pennsylvania, USA. Hydrol Process 28(10):3574–3601. https://doi.org/10.1002/hyp.9885 |
[103] | Islam MR, Shinjo R, Faruque MO, Shimada H, Howladar MF (2016) Finite element method (FEM) groundwater inflow modeling associated with an unconfined aquifer into the open-pit coalmine of the Phulbari area, NW Bangladesh. Arab J Geosci 9(5):394–405. https://doi.org/10.1007/s12517-016-2402-1 |
[104] | Jankowski J, Knights P (2010) Surface water–groundwater interaction in the fractured sandstone aquifer impacted by mining-induced subsidence: 1. Hydrology and hydrogeology. Biuletyn Panstwowego Instytutu Geologicznego 441:33–42 |
[105] | Jeanpert J, Iseppi M, Adler PM, Genthon P, Sevin B, Thovert JF, Dewandel B, Join JL (2019) Fracture controlled permeability of ultramafic basement aquifers. Inferences from the Koniambo massif, New Caledonia. Eng Geol 256:67–83. https://doi.org/10.1016/j.enggeo.2019.05.006 |
[106] | Ji Y, Zhang M, Wang Y, Wang P, Wang A, Wu Y, Xu H, Zhang Y (2014) Microwave-photonic sensor for remote water-level monitoring based on chaotic laser. Int J Bifurc Chaos 24(3):1–7. https://doi.org/10.1142/S0218127414500321 |
[107] | Jing Z, Wang J, Zhu Y, Feng Y (2018) Effects of land subsidence resulted from coal mining on soil nutrient distributions in a loess area of China. J Clean Prod 177(10):350–361. https://doi.org/10.1016/j.jclepro.2017.12.191 |
[108] | Kang H, Zhang X, Si L, Wu Y, Gao F (2010) In-situ stress measurements and stress distribution characteristics in underground coal mines in China. Eng Geol 116(3):333–345. https://doi.org/10.1016/j.enggeo.2010.09.015 |
[109] | Karacan CO, Diamond WP, Esterhuizen GS, Schatzel SJ (2005) Numerical analysis of the impact of longwall panel width on methane emissions and performance of gob gas ventholes. In: Proceedings of the 2005 International Coalbed Methane Symposium, Tuscaloosa, Alabama, pp 1–28 |
[110] | Karacan CO, Goodman G (2009) Hydraulic conductivity changes and influencing factors in longwall overburden determined by slug tests in gob gas ventholes. Int J Rock Mech Min Sci 46(7):1162–1174. https://doi.org/10.1016/j.ijrmms.2009.02.005 |
[111] | Kay D, Barbato J, Brassington G, Somer Bd (2006) Impacts of longwall mining to rivers and cliffs in the Southern Coalfield. In: Proceedings of the 2006 Coal Operators’ Conference, Wollongong, NSW, Australia, pp 327–336 |
[112] | Kendorski FS (1993) Effect of high-extraction coal mining on surface and ground waters. In: Proceedings of the 12th Conference on Ground Control in Mining, Morgantown, West Virginia, USA, pp 412–425 |
[113] | Kendorski FS (2006) Effect of full-extraction underground mining on ground and surface waters: a 25-year retrospective. In: Proceedings of the 25th International Conference on Ground Control in Mining, Lombard, Illinois, USA, pp 1–6 |
[114] | Kendorski FS, Khosla I, Singh MM (1979) Criteria for determining when a body of surface water constitutes a hazard to mining. Final Report to U.S. Bureau of Mines on Contract J0285011, Engineers International |
[115] | Khanal M, Adhikary D, Poulsen B, Guo H (2019a) Mine site specific longwall-induced permeability changes: a case study. Geotech Geol Eng 37(4):3271–3282. https://doi.org/10.1007/s10706-019-00842-z |
[116] | Khanal M, Guo H, Adhikary D (2019b) 3D numerical study of underground coal mining induced strata deformation and subsequent permeability change. Geotech Geol Eng 37(1):235–249. https://doi.org/10.1007/s10706-018-0605-9 |
[117] | Khanal M, Poulsen B, Adhikary D, Balusu R, Wilkins A, Belle B (2021) Numerical study of stability and connectivity of vertical goaf drainage holes. Geotech Geol Eng 39(3):2669–2679. https://doi.org/10.1007/s10706-020-01650-6 |
[118] | Kong X, Liu Q, Lu H (2021) Effects of rock specimen size on mechanical properties in laboratory testing. J Geotech Geoenviron Eng 147(5):04021013. https://doi.org/10.1061/(ASCE)GT.1943-5606.0002478 |
[119] | Kowalski ACG, Mendonça CA, Ofterdinger US (2021) Fracture flow characterization with low-noise spontaneous potential logging. Groundwater 59(1):16–23. https://doi.org/10.1111/gwat.13009 |
[120] | Kumar V, Chandra A, Usmani Z (2017) Impact of coal mining on soil properties and their efficient eco-restoration. Int J Energy Technol Policy 13(1/2):158–165. https://doi.org/10.1504/IJETP.2017.080613 |
[121] | Lechner AM, Baumgartl T, Matthew P, Glenn V (2016) The impact of underground longwall mining on prime agricultural land: a review and research agenda. Land Degrad Dev 27(6):1650–1663. https://doi.org/10.1002/ldr.2303 |
[122] | Lee S, Moon J-S (2020) Excessive groundwater inflow during TBM tunneling in limestone formation. Tunn Undergr Space Technol 96:103217. https://doi.org/10.1016/j.tust.2019.103217 |
[123] | Lei S, Bian Z, Daniels JL, He X (2010) Spatio-temporal variation of vegetation in an arid and vulnerable coal mining region. Min Sci Technol 20(3):485–490. https://doi.org/10.1016/S1674-5264(09)60230-1 |
[124] | Li B, Liu R, Jiang Y (2016a) Influences of hydraulic gradient, surface roughness, intersecting angle, and scale effect on nonlinear flow behavior at single fracture intersections. J Hydrol 538:440–453. https://doi.org/10.1016/j.jhydrol.2016.04.053 |
[125] | Li D, Tian Z, Ma Y, Gu J, Ji Y, Li S (2019) Application of grounded electrical source airborne transient electromagnetic (GREATEM) system in goaf water detection. J Environ Eng Geophys 24(3):387–397. https://doi.org/10.2113/JEEG24.3.387 |
[126] | Li G, Paquet R, Steuary P, Ramage R, Perceval J (2022) An introduction to the standardised subsidence information management system. In: Proceedings of the 11th Conference on Mine Subsidence, Pokolbin, NSW, Australia, pp 39–56 |
[127] | Li L, Yang T, Liang Z, Zhu W, Tang C (2011) Numerical investigation of groundwater outbursts near faults in underground coal mines. Int J Coal Geol 85(3):276–288. https://doi.org/10.1016/j.coal.2010.12.006 |
[128] | Li P, Ma L, Wu Y, Zhang L, Hao Y (2018) Concurrent mining during construction and water-filling of a goaf groundwater reservoir in a coal mine. Mine Water Environ 37(4):763–773. https://doi.org/10.1007/s10230-018-0537-x |
[129] | Li S, Chen YF, Li ZJ, Zhang K (2016b) Applying a statistical method to streamflow reduction caused by underground mining for coal in the Kuye River basin. Sci China Technol Sci 59(12):1911–1920. https://doi.org/10.1007/s11431-016-0393-4 |
[130] | Liu R, Jiang Y, Li B (2016a) Effects of intersection and dead-end of fractures on nonlinear flow and particle transport in rock fracture networks. Geosci J 20(3):415–426. https://doi.org/10.1007/s12303-015-0057-7 |
[131] | Liu R, Jiang Y, Li B, Wang X (2015a) A fractal model for characterizing fluid flow in fractured rock masses based on randomly distributed rock fracture networks. Comput Geotech 65:45–55. https://doi.org/10.1016/j.compgeo.2014.11.004 |
[132] | Liu R, Li B, Jiang Y (2016b) Critical hydraulic gradient for nonlinear flow through rock fracture networks: the roles of aperture, surface roughness, and number of intersections. Adv Water Resour 88:53–65. https://doi.org/10.1016/j.advwatres.2015.12.002 |
[133] | Liu X, Tan Y, Ning J, Tian C, Wang J (2015b) The height of water-conducting fractured zones in longwall mining of shallow coal seams. Geotech Geol Eng 33(3):693–700. https://doi.org/10.1007/s10706-015-9851-2 |
[134] | Liu Y, Wu L, Yue H, Ma B, Meng Z (2016c) Effect of underground coal mining on land surface soil moisture in desert mining area. Coal Sci Technol 44(11):197–202 |
[135] | Liu Z, Xiong C (2007) Numerical simulation study on water inrush mechanism from collapse column. Chin J Rock Mech Eng 26(2):4013–4018 |
[136] | Liu Z, Zhong X, Qin B, Ren H, Gao A (2019) Redevelopment of fractures and permeability changes after multi-seam mining of shallow closely spaced coal seams. Arch Min Sci 64(4):671–686. https://doi.org/10.24425/ams.2019.129376 |
[137] | Lu H, Shan N, Zhang Y, Liang X (2020) Effect of strain-dependent hydraulic conductivity of coal rock on groundwater inrush in mining. Geofluids 8887392:1–15. https://doi.org/10.1155/2020/8887392 |
[138] | Luan J, Zhang Y, Tian J, Meresa H, Liu D (2020) Coal mining impacts on catchment runoff. J Hydrol 589(125101):1–13. https://doi.org/10.1016/j.jhydrol.2020.125101 |
[139] | Luo AK, Hou Y, Hu XY (2018) Mining influence on underground water resources in arid and semiarid regions. In: Proceedings of IOP Conference Series: Earth and Environmental Science, Ordos, China, pp 1–8 |
[140] | Luo Y, Peng SS, Preusse A, Henog C, Wings R, Smka A, Smka A (2001) Effect of face advance rates on the characteristics of subsidence processes associated with US and Germany longwall mining operations. In: Proceedings of the 20th International Conference on Ground Control in Mining, Morgantown, West Virginia, pp 140–148 |
[141] | Ma D, Bai H, Wang Y (2015) Mechanical behavior of a coal seam penetrated by a karst collapse pillar: mining-induced groundwater inrush risk. Nat Hazards 75(3):2137–2151. https://doi.org/10.1007/s11069-014-1416-9 |
[142] | Ma D, Cai X, Li Q, Duan H (2018a) In-situ and numerical investigation of groundwater inrush hazard from grouted karst collapse pillar in longwall mining. Water 10(9):1–17. https://doi.org/10.3390/w10091187 |
[143] | Ma D, Duan H, Zhang J, Bai H (2022) A state-of-the-art review on rock seepage mechanism of water inrush disaster in coal mines. Int J Coal Sci Technol 9(1):50. https://doi.org/10.1007/s40789-022-00525-w |
[144] | Ma D, Miao X, Bai H, Huang J, Pu H, Wu Y, Zhang G, Li J (2016) Effect of mining on shear sidewall groundwater inrush hazard caused by seepage instability of the penetrated karst collapse pillar. Nat Hazards 82(1):73–93. https://doi.org/10.1007/s11069-016-2180-9 |
[145] | Ma Y, Zhang Y, Yu Z, Huang Y, Zhang C (2018b) Heat transfer by water flowing through rough fractures and distribution of local heat transfer coefficient along the flow direction. Int J Heat Mass Transf 119:139–147. https://doi.org/10.1016/j.ijheatmasstransfer.2017.11.102 |
[146] | Mahabadi OK, Cottrell BE, Grasselli G (2010) An example of realistic modelling of rock dynamics problems: FEM/DEM simulation of dynamic Brazilian test on Barre granite. Rock Mech Rock Eng 43(6):707–716. https://doi.org/10.1007/s00603-010-0092-7 |
[147] | Mather JD, Gray DA, Jenkins DG (1969) The use of tracers to investigate the relationship between mining subsidence and groundwater occurrence at Aberfan, South Wales. J Hydrol 9(2):136–154. https://doi.org/10.1016/0022-1694(69)90073-0 |
[148] | Meng Z, Shi X, Li G (2016) Deformation, failure and permeability of coal-bearing strata during longwall mining. Eng Geol 208(24):69–80. https://doi.org/10.1016/j.enggeo.2016.04.029 |
[149] | Meye SM, Shen Z (2020) Numerical simulation of the influence of tunnel construction by mining method on the seepage field in weathered granite stratum. Engineering 12(6):424–456. https://doi.org/10.4236/eng.2020.126031 |
[150] | Mills KW (2007) Subsidence impacts on river channels and opportunities for control. In: Proceedings of the 7th Conference on Mine Subsidence, Wollongong, NSW, Australia, pp 207–217 |
[151] | Mills KW (2022) A perspective on the mechanics of mine subsidence above longwall panels. In: Proceedings of the 11th Conference on Mine Subsidence, Pokolbin, NSW, Australia, pp 1–37 |
[152] | Mills KW, O’Grady P (1998) Impact of longwall width on overburden behaviour. In: Proceedings of the 1998 Coal Operators’ Conference, Wollongong, NSW, Australia, pp 147–155 |
[153] | Min K-B, Rutqvist J, Tsang C-F, Jing L (2004) Stress-dependent permeability of fractured rock masses: a numerical study. Int J Rock Mech Min Sci 41(7):1191–1210. https://doi.org/10.1016/j.ijrmms.2004.05.005 |
[154] | Minns SA, Kipp JA, Carey DI, Dinger JS, Sendlein LVA (1995) Effects of longwall mining on hydrogeology, Leslie County, Kentucky part 1: pre-mining conditions. Kentucky Geological Survey Report of Investigations, University of Kentucky |
[155] | Molron J, Linde N, Baron L, Selroos J-O, Darcel C, Davy P (2020) Which fractures are imaged with Ground Penetrating Radar? Results from an experiment in the Äspö Hardrock Laboratory, Sweden. Eng Geol 273:105674. https://doi.org/10.1016/j.enggeo.2020.105674 |
[156] | Mondal D, Roy PNS, Kumar M (2020) Monitoring the strata behavior in the destressed zone of a shallow Indian longwall panel with hard sandstone cover using mine-microseismicity and borehole televiewer data. Eng Geol 271:105593. https://doi.org/10.1016/j.enggeo.2020.105593 |
[157] | Moridi MA, Sharifzadeh M, Kawamura Y, Jang HD (2018) Development of wireless sensor networks for underground communication and monitoring systems (the cases of underground mine environments). Tunn Undergr Space Technol 73:127–138. https://doi.org/10.1016/j.tust.2017.12.015 |
[158] | Newman C, Agioutantis Z, Leon GBJ (2016) Evaluation of potential impacts to stream and ground water due to underground coal mining. In: Proceedings of the 35th International Conference on Ground Control in Mining, Morgantown, West Virginia, USA, pp 1–8 |
[159] | Newman C, Agioutantis Z, Leon GBJ (2017) Assessment of potential impacts to surface and subsurface water bodies due to longwall mining. Int J Min Sci Technol 27(1):57–64. https://doi.org/10.1016/j.ijmst.2016.11.016 |
[160] | Niwas S, Tezkan B, Israil M (2011) Aquifer hydraulic conductivity estimation from surface geoelectrical measurements for Krauthausen test site, Germany. Hydrogeol J 19(2):307–315. https://doi.org/10.1007/s10040-010-0689-7 |
[161] | Palchik V (2002) Influence of physical characteristics of weak rock mass on height of caved zone over abandoned subsurface coal mines. Environ Geol 42(1):92–101. https://doi.org/10.1007/s00254-002-0542-y |
[162] | Palchik V (2003) Formation of fractured zones in overburden due to longwall mining. Environ Geol 44(1):28–38. https://doi.org/10.1007/s00254-002-0732-7 |
[163] | PanIgrahy BP, Singh PK, Tiwari AK, Kumar B (2015) Variation in groundwater quality with seasonal fluctuation in Jharia coal mine region, Jharkhand. India. Curr World Environ 10(1):171–178. https://doi.org/10.12944/CWE.10.1.19 |
[164] | Poulsen BA, Adhikary D, Guo H (2018) Simulating mining-induced strata permeability changes. Eng Geol 237(10):208–216. https://doi.org/10.1016/j.enggeo.2018.03.001 |
[165] | Pratt HR, Black AD, Brown WS, Brace WF (1972) The effect of speciment size on the mechanical properties of unjointed diorite. Int J Rock Mech Min Sci Geomech Abstr 9(4):513–516. https://doi.org/10.1016/0148-9062(72)90042-3 |
[166] | Price G, Blacka B, Baut NL (2022) Permeation grouting of a subsidence impacted watercourse in the southern coalfields. In: Proceedings of the 11th Conference on Mine Subsidence, Pokolbin, NSW, Australia, pp 327–337 |
[167] | Qin D, Wang X, Zhang D, Guan W, Zhang L, Xu M (2019) Occurrence characteristic and mining technology of ultra-thick coal seam in Xinjiang. China Sustain 11(22):6470. https://doi.org/10.3390/su11226470 |
[168] | Rapantová N, Grmela A, Vojtek D, Halir J, Michalek B (2007) Ground water flow modelling applications in mining hydrogeology. Mine Water Environ 26(4):264–270. https://doi.org/10.1007/s10230-007-0017-1 |
[169] | Rapantová N, Krzeszowski Ś, Grmela A, Wolkersdorfer C (2012) Quantitative assessment of mine water sources based on the general mixing equation and multivariate statistics. Mine Water Environ 31(4):252–265. https://doi.org/10.1007/s10230-012-0192-6 |
[170] | Ren Z, Wang N (2020) The overburden strata caving characteristics and height determination of water conducting fracture zone in fully mechanized caving mining of extra thick coal seam. Geotech Geol Eng 38(1):329–341. https://doi.org/10.1007/s10706-019-01019-4 |
[171] | Rezaei M, Hossaini MF, Majdi A (2015a) Determination of longwall mining-induced stress using the strain energy method. Rock Mech Rock Eng 48(6):2421–2433. https://doi.org/10.1007/s00603-014-0704-8 |
[172] | Rezaei M, Hossaini MF, Majdi A (2015b) A time-independent energy model to determine the height of destressed zone above the mined panel in longwall coal mining. Tunn Undergr Space Technol 47:81–92. https://doi.org/10.1016/j.tust.2015.01.001 |
[173] | Sasaoka T, Shimada H, Lin NZ, Takamoto H, Matsui K, Kramadibrata S, Sulistianto B (2014) Geotechnical issues in the application of rock bolting technology for the development of underground coal mines in Indonesia. Int J Min Reclam Environ 28(3):150–172. https://doi.org/10.1080/17480930.2013.804258 |
[174] | Seedsman R (2019) Interpretations of mine water pump-out data and revisions to caving and fracturing models for longwalls. Mine Water Environ 38(3):676–685. https://doi.org/10.1007/s10230-019-00627-z |
[175] | Sgambat JP, LaBella EA, Roebuck S (1980) Effects of underground coal mining on ground water in the eastern United States. Interagency Energy-Environment Research and Development, Industrial Environmental Research Laboratory, U.S. Environmental Protection Agency |
[176] | Sidle RC, Kamil I, Sharma A, Yamashita S (2000) Stream response to subsidence from underground coal mining in central Utah. Environ Geol 39(3–4):279–291. https://doi.org/10.1007/s002540050007 |
[177] | Smerdon BD, Smith LA, Harrington GA, Gardner WP, Piane CD, Sarout J (2014) Estimating the hydraulic properties of an aquitard from in situ pore pressure measurements. Hydrogeol J 22(8):1875–1887. https://doi.org/10.1007/s10040-014-1161-x |
[178] | Smith JA, Ramandi HL, Zhang C, Timms W (2019) Analysis of the influence of groundwater and the stress regime on bolt behaviour in underground coal mines. Int J Coal Sci Technol 6(2):286–300. https://doi.org/10.1007/s40789-019-0246-5 |
[179] | Snow DT (1965) A parallel plate model of fractured permeable media. University of California, Berkeley |
[180] | Snow DT (1969) Anisotropie permeability of fractured media. Water Resour Res 5(6):1273–1289. https://doi.org/10.1029/WR005i006p01273 |
[181] | Su H, Li H, Zhang L, Wen Z (2020) Particle flow code method-based seepage behavior analysis and control effect evaluation for soil levee. Eng Comput 36(1):97–114. https://doi.org/10.1007/s00366-018-0687-2 |
[182] | Sun K, Fan L, Xia Y, Li C, Chen J, Gao S, Wu B, Peng J, Ji Y (2021) Impact of coal mining on groundwater of Luohe Formation in Binchang mining area. Int J Coal Sci Technol 8(1):88–102. https://doi.org/10.1007/s40789-020-00366-5 |
[183] | Tammetta P (2013) Estimation of the height of complete groundwater drainage above mined longwall panels. Groundwater 52(5):723–734. https://doi.org/10.1111/gwat.12003 |
[184] | Tammetta P (2015) Estimation of the change in hydraulic conductivity above mined longwall panels. Groundwater 53(1):122–129. https://doi.org/10.1111/gwat.12153 |
[185] | Tammetta P (2016) Estimation of the change in storage capacity above mined longwall panels. Ground Water 54(5):646–655. https://doi.org/10.1111/gwat.12405 |
[186] | Tieman GE, Rauch HW (1986) Study of dewatering effects at an underground longwall mine site in the Pittsburgh Seam of the Northern Appalachian Coalfield. In: Proceedings of Bureau of Mines Technology Transfer Seminar, Pittsburgh, Pennsylvania, USA, pp 72–89 |
[187] | Tiwari AK, Maio MD, Singh PK, Singh AK (2016) Hydrogeochemical characterization and groundwater quality assessment in a coal mining area, India. Arab J Geosci 9(3):1–17. https://doi.org/10.1007/s12517-015-2209-5 |
[188] | Trotter T, Frazier P (2007) Monitoring the impact of longwall mine subsidence on surface environments using remote sensing. In: Proceedings of the 7th Conference on Mine Subsidence, Wollongong, NSW, Australia, pp 101–109 |
[189] | Unlu T, Akcin H, Yilmaz O (2013) An integrated approach for the prediction of subsidence for coal mining basins. Eng Geol 166(8):186–203. https://doi.org/10.1016/j.enggeo.2013.07.014 |
[190] | Vakili A, Hebblewhite BK (2010) A new cavability assessment criterion for Longwall Top Coal Caving. Int J Rock Mech Min Sci 47(8):1317–1329. https://doi.org/10.1016/j.ijrmms.2010.08.010 |
[191] | Vishwakarma AK, Rai R, Sonkar AK, Behera T, Shrivastva BK (2020) Assessment of impacts of coal mining–induced subsidence on native flora and native forest land: a brief review. In: Sustainable development practices using geoinformatics, Scrivener Publishing LLC, pp 141–152 |
[192] | Vlachopoulos N, Carrapatoso C, Vazaios I, Forbes B (2020) Integration among LiDAR-based structural input, DFN generation and DOS technology to describe the rock mass fracture system. In: Proceedings of ISRM International Symposium-EUROCK 2020, pp 1–8 |
[193] | Vukelić Ž, Dervarič E, Šporin J, Vižintin G (2016) The development of dewatering predictions of the Velenje coalmine. Energies 9(9):702. https://doi.org/10.3390/en9090702 |
[194] | Wahler WA (1979) Dewatering active underground coal mines: technical aspects and cost-effectiveness. Interagency Energy-Environment Research and Development, Industrial Environmental Research Laboratory. Environmental Protection Agency, Washington |
[195] | Walsh R, Brown S, Mills K, Heritage R (2022) Geological and hydrological characteristics of the Elouera Fault and the impacts of mining. In: Proceedings of the 11th Conference on Mine Subsidence, Pokolbin, NSW, Australia, pp 121–133 |
[196] | Wang C, Wang Y, Lu S (2000) Deformational behaviour of roadways in soft rocks in underground coal mines and principles for stability control. Int J Rock Mech Min Sci 37(6):937–946. https://doi.org/10.1016/S1365-1609(00)00026-5 |
[197] | Wang C, Zhang C, Zhao X, Liao L, Zhang S (2018a) Dynamic structural evolution of overlying strata during shallow coal seam longwall mining. Int J Rock Mech Min Sci 103:20–32. https://doi.org/10.1016/j.ijrmms.2018.01.014 |
[198] | Wang F, Tu S, Zhang C, Zhang Y, Bai Q (2016) Evolution mechanism of water-flowing zones and control technology for longwall mining in shallow coal seams beneath gully topography. Environ Earth Sci 75(19):1309. https://doi.org/10.1007/s12665-016-6121-4 |
[199] | Wang G, Wu M, Wang R, Xu H, Song X (2017a) Height of the mining-induced fractured zone above a coal face. Eng Geol 216(12):140–152. https://doi.org/10.1016/j.enggeo.2016.11.024 |
[200] | Wang L, Chen W, Tan X, Tan X, Yang J, Yang D, Zhang X (2020a) Numerical investigation on the stability of deforming fractured rocks using discrete fracture networks: a case study of underground excavation. Bull Eng Geol Env 79(1):133–151. https://doi.org/10.1007/s10064-019-01536-9 |
[201] | Wang M, Song Z, Gou P, Li J, Zhang S, Zheng D (2020b) Effect of mining speed on stability control of overburden structure in fully mechanized coal face. J China Univ Min Technol 49(3):463–470 |
[202] | Wang S, Li X, Wang S (2017b) Separation and fracturing in overlying strata disturbed by longwall mining in a mineral deposit seam. Eng Geol 226(30):257–266. https://doi.org/10.1016/j.enggeo.2017.06.015 |
[203] | Wang T, David C, Zhu W (1997) The transition from brittle faulting to cataclastic flow in porous sandstones: mechanical deformation. J Geophys Res 102(B2):3009–3025. https://doi.org/10.1029/96JB03281 |
[204] | Wang W, Cheng Y, Wang H, Li W, Wang L (2015) Coupled disaster-causing mechanisms of strata pressure behavior and abnormal gas emissions in underground coal extraction. Environ Earth Sci 74(9):6717–6735. https://doi.org/10.1007/s12665-015-4682-2 |
[205] | Wang X, Qin D, Zhang D, Guan W, Xu M, Wang X, Zhang C (2019) Evolution characteristics of overburden strata structure for ultra-thick coal seam multi-layer mining in Xinjiang East Junggar Basin. Energies 12(2):332. https://doi.org/10.3390/en12020332 |
[206] | Wang X, Zhang Y, Zhang Z (2018b) Influence of coal mining on permeability coefficient and water level of losse aquifer in thick loess covered area. Water Resour Power 36(2):85–88 |
[207] | Wang Y, Li B, Chen C, Jia H (2020c) Influence of groundwater level fluctuation on lateral deformation of cantilever enclosure structure of pit-in-pit. Mar Georesour Geotechnol 38(1):108–113. https://doi.org/10.1080/1064119X.2018.1558471 |
[208] | Wei J, Wu F, Yin H, Guo J, Xie D, Xiao L, Zhi H, Lefticariu L (2017) Formation and height of the interconnected fractures zone after extraction of thick coal seams with weak overburden in western China. Mine Water Environ 36(1):59–66. https://doi.org/10.1007/s10230-016-0396-2 |
[209] | Wei S, Li Y, Shang Y, Sun Y, Li K (2022) Stability analysis of fractured rock mass around underground excavations based on a three-dimensional discrete fracture network. Environ Earth Sci 81(3):65. https://doi.org/10.1007/s12665-021-10145-3 |
[210] | Wei S, Ren H (2012) PFC/FLAC coupled numerical simulation of excavation damage zone in deep schist tunnel. Appl Mech Mater 236–237:622–626. https://doi.org/10.4028/www.scientific.net/AMM.236-237.622 |
[211] | Whittles DN, Lowndes IS, Kingman SW, Yates C, Jobling S (2006) Influence of geotechnical factors on gas flow experienced in a UK longwall coal mine panel. Int J Rock Mech Min Sci 43(3):369–387. https://doi.org/10.1016/j.ijrmms.2005.07.006 |
[212] | Wilkins A, Qu Q (2020) A formalism to compute permeability changes in anisotropic fractured rocks due to arbitrary deformations. Int J Rock Mech Min Sci 125:104159. https://doi.org/10.1016/j.ijrmms.2019.104159 |
[213] | Wu C, Wu X, Zhu G, Qian C (2019) Predicting mine water inflow and groundwater levels for coal mining operations in the Pangpangta coalfield. China Environ Earth Sci 78(5):130–142. https://doi.org/10.1007/s12665-019-8098-2 |
[214] | Wu Q, Wang M, Wu X (2004) Investigations of groundwater bursting into coal mine seam floors from fault zones. Int J Rock Mech Min Sci 41(4):557–571. https://doi.org/10.1016/j.ijrmms.2003.01.004 |
[215] | Wu X, Wang X, Jiang X, Duan Q, Sun Y, Tian H, Jia Z (2009) A study on coal mining under large reservoir areas. Environ Geol 57(3):675–683. https://doi.org/10.1007/s00254-008-1346-5 |
[216] | Wu Y, Cao G, Ding W (2005) Experimental study on relation between seepage and stress of sandstone in CT scale. J Rock Mech Eng 24(23):4203–4209 |
[217] | Xie P, Luo Y, Wu Y, Gao X, Luo S, Zeng Y (2020) Roof deformation associated with mining of two panels in steeply dipping coal seam using subsurface subsidence prediction model and physical simulation experiment. Min Metall Explor 37(2):581–591. https://doi.org/10.1007/s42461-019-00156-x |
[218] | Xu J, Zhu W, Xu J, Wu J, Li Y (2021) High-intensity longwall mining-induced ground subsidence in Shendong coalfield, China. Int J Rock Mech Min Sci 141:104730. https://doi.org/10.1016/j.ijrmms.2021.104730 |
[219] | Xu S, Zhang Y, Shi H, Zhang Z, Chen J (2020) Impacts of aquitard properties on an overlying unconsolidated aquifer in a mining area of the loess plateau: case study of the Changcun Colliery, Shanxi. Mine Water Environ 39(1):121–134. https://doi.org/10.1007/s10230-019-00649-7 |
[220] | Yang D, Bian Z, Lei S (2016) Impact on soil physical qualities by the subsidence of coal mining: a case study in Western China. Environ Earth Sci 75(8):1–14. https://doi.org/10.1007/s12665-016-5439-2 |
[221] | Yang D, Ning Z, Li Y, Lv Z, Qiao Y (2021) In situ stress measurement and analysis of the stress accumulation levels in coal mines in the northern Ordos Basin, China. Int J Coal Sci Technol 8(6):1316–1335. https://doi.org/10.1007/s40789-021-00407-7 |
[222] | Yang TH, Liu J, Zhu WC, Elsworth D, Tham LG, Tang CA (2007) A coupled flow-stress-damage model for groundwater outbursts from an underlying aquifer into mining excavations. Int J Rock Mech Min Sci 44(1):87–97. https://doi.org/10.1016/j.ijrmms.2006.04.012 |
[223] | Yang X, Wen G, Dai L, Sun H, Li X (2019) Ground subsidence and surface cracks evolution from shallow-buried close-distance multi-seam mining: a case study in bulianta coal mine. Rock Mech Rock Eng 52(8):2835–2852. https://doi.org/10.1007/s00603-018-1726-4 |
[224] | Yang Y, Erskine PD, Zhang S, Wang Y, Bian Z, Lei S (2018) Effects of underground mining on vegetation and environmental patterns in a semi-arid watershed with implications for resilience management. Environ Earth Sci 77(17):1–12. https://doi.org/10.1007/s12665-018-7796-5 |
[225] | Yao B, Bai H, Zhang B (2012) Numerical simulation on the risk of roof water inrush in Wuyang Coal Mine. Int J Min Sci Technol 22(2):273–277. https://doi.org/10.1016/j.ijmst.2012.03.006 |
[226] | Yin H, Wei J, Lefticariu L, Guo J, Xie D, Li Z, Zhao P (2016) Numerical simulation of water flow from the coal seam floor in a deep longwall mine in China. Mine Water Environ 35(2):243–252. https://doi.org/10.1007/s10230-016-0385-5 |
[227] | Yin H, Zhao H, Xie D, Sang S, Shi Y, Tian M (2019) Mechanism of mine water inrush from overlying porous aquifer in Quaternary: a case study in Xinhe Coal Mine of Shandong Province. China Arab J Geosci 12(5):1–12. https://doi.org/10.1007/s12517-019-4325-0 |
[228] | Yoshinaka R, Osada M, Park H, Sasaki T, Sasaki K (2008) Practical determination of mechanical design parameters of intact rock considering scale effect. Eng Geol 96(3):173–186. https://doi.org/10.1016/j.enggeo.2007.10.008 |
[229] | Zaidel J, Markham B, Bleiker D (2010) Simulating seepage into mine shafts and tunnels with MODFLOW. Groundwater 48(3):390–400. https://doi.org/10.1111/j.1745-6584.2009.00659.x |
[230] | Zhang C (2014) Evaluation of mechanisms contributing to valley closure subsidence effects under irregular topographic conditions. Dissertation, Univeristy of New South Wales |
[231] | Zhang C, Mitra R, Hebblewhite B (2012) Review of numerical modelling evaluation of mechanisms for valley closure subsidence under irregular topographic condition. In: Proceedings of ISRM International Symposium, Stockholm, Sweden, pp 1–15 |
[232] | Zhang C, Mitra R, Hebblewhite B (2013a) Evaluation of valley closure subsidence effects under irregular topographic conditions. Min Technol 122(3):172–183. https://doi.org/10.1179/1743286313Y.0000000036 |
[233] | Zhang C, Mitra R, Oh J, Canbulat I, Hebblewhite B (2018a) Numerical analysis on mining-induced fracture development around river valleys. Int J Min Reclam Environ 32(7):463–485. https://doi.org/10.1080/17480930.2017.1293495 |
[234] | Zhang C, Mitra R, Oh J, Hebblewhite B (2016) Analysis of mining-induced valley closure movements. Rock Mech Rock Eng 49(5):1923–1941. https://doi.org/10.1007/s00603-015-0880-1 |
[235] | Zhang D, Fan G, Liu Y, Ma L (2010) Field trials of aquifer protection in longwall mining of shallow coal seams in China. Int J Rock Mech Min Sci 47(6):908–914. https://doi.org/10.1016/j.ijrmms.2010.06.018 |
[236] | Zhang D, Fan G, Ma L, Wang A, Liu Y (2009a) Harmony of large-scale underground mining and surface ecological environment protection in desert district—a case study in Shendong mining area, northwest of China. Procedia Earth Planet Sci 1(1):1114–1120. https://doi.org/10.1016/j.proeps.2009.09.171 |
[237] | Zhang D, Fan G, Ma L, Wang X (2011) Aquifer protection during longwall mining of shallow coal seams: a case study in the Shendong Coalfield of China. Int J Coal Geol 86(2–3):190–196. https://doi.org/10.1016/j.coal.2011.01.006 |
[238] | Zhang G, Chen L, Wen Z, Chen M, Tao G, Li Y, Zuo H (2020a) Squeezing failure behavior of roof-coal masses in a gob-side entry driven under unstable overlying strata. Energy Sci Eng 8(7):2443–2456. https://doi.org/10.1002/ese3.678 |
[239] | Zhang HQ, He YN, Tang CA, Ahmad B, Han LJ (2009b) Application of an improved flow-stress-damage model to the criticality assessment of water inrush in a mine: a case study. Rock Mech Rock Eng 42(6):911–930. https://doi.org/10.1007/s00603-008-0004-2 |
[240] | Zhang J, Peng S (2005) Water inrush and environmental impact of shallow seam mining. Environ Geol 48(8):1068–1076. https://doi.org/10.1007/s00254-005-0045-8 |
[241] | Zhang R, Jiang Z, Sun Q, Zhu S (2013b) The relationship between the deformation mechanism and permeability on brittle rock. Nat Hazards 66(2):1179–1187. https://doi.org/10.1007/s11069-012-0543-4 |
[242] | Zhang S, Fan G, Jiang S, Fan Z, Li S, Ni H (2021) Dual-hazard control mechanism of burst-prone and spontaneous combustion coalface considering effect of retreat speed. Energy Rep 7:278–288. https://doi.org/10.1016/j.egyr.2020.12.033 |
[243] | Zhang S, Fan G, Zhang D, Li Q (2018b) Physical simulation research on evolution laws of clay aquifuge stability during slice mining. Environ Earth Sci 77(7):278–287. https://doi.org/10.1007/s12665-018-7401-y |
[244] | Zhang S, Tang S, Zhang D, Fan G, Wang Z (2017) Determination of the height of the water-conducting fractured zone in difficult geological structures: a case study in Zhao Gu No. 1 coal seam. Sustainability 9(7):1077–1095. https://doi.org/10.3390/su9071077 |
[245] | Zhang S, Zhang D, Wang Z, Chen M (2018c) Influence of stress and water pressure on the permeability of fissured sandstone under hydromechanical coupling. Mine Water Environ 37(4):774–785. https://doi.org/10.1007/s10230-018-0540-2 |
[246] | Zhang T, Gan Q, Zhao Y, Zhu G, Nie X, Yang K, Li J (2019) Investigations into mining-induced stress-fracture-seepage field coupling effect considering the response of key stratum and composite aquifer. Rock Mech Rock Eng 52(10):4017–4031. https://doi.org/10.1007/s00603-019-01811-x |
[247] | Zhang W, Zhou G, Zhang G, Duan J, Kong Y, Jiang T (2020b) Study on 3D spatial characterization analysis and water injection seepage numerical simulation of coal micro-pore/fracture. Energy Sources Part A Recov Util Environ Effects 2020:1–15. https://doi.org/10.1080/15567036.2020.1808742 |
[248] | Zhang Y, Fan C (2014) Seepage field-strain field coupling analysis for rock masses of coal seam floor during mining based on COMSOL. Adv Mater Res 1010–1012:1467–1470. https://doi.org/10.4028/www.scientific.net/AMR.1010-1012.1467 |
[249] | Zhang Y, Tu S, Bai Q, Li J (2013c) Overburden fracture evolution laws and water-controlling technologies in mining very thick coal seam under water-rich roof. Int J Min Sci Technol 23(5):693–700. https://doi.org/10.1016/j.ijmst.2013.08.013 |
[250] | Zhao C, Jin D, Geng J, Sun Q (2019) Numerical simulation of the groundwater system for mining shallow buried coal seams in the ecologically fragile areas of Western China. Mine Water Environ 38(1):158–165. https://doi.org/10.1007/s10230-018-0551-z |
[251] | Zhao Q (2020) Research on evolution law of mining-induced water resisting property of the two groups of clay aquicludes in thick coal seam mining in Jinjitan coal mine. Dissertation, China University of Mining and Technology |
[252] | Zhao Y, Tang J, Chen Y, Zhang L, Wang W, Wan W, Liao J (2017) Hydromechanical coupling tests for mechanical and permeability characteristics of fractured limestone in complete stress–strain process. Environ Earth Sci 76(24):1–18. https://doi.org/10.1007/s12665-016-6322-x |
[253] | Zhou Q (2017) Modeling and risk assessment of fault-induced mine water inrush. Dissertation, Technical University of Madrid |
[254] | Zhu W, Wong T (1997) The transition from brittle faulting to cataclastic flow: permeability evolution. J Geophys Res 102(B2):3027–3041. https://doi.org/10.1029/96JB03282 |
[255] | Zhu WC, Wei CH (2011) Numerical simulation on mining-induced water inrushes related to geologic structures using a damage-based hydromechanical model. Environ Earth Sci 62(1):43–54. https://doi.org/10.1007/s12665-010-0494-6 |
[256] | Zoorabadi M (2014) Hydraulic conductivity of jointed rocks. Dissertation, UNSW Sydney |
[257] | Zoorabadi M, Indraratna B, Nemcik J (2012) A new equation for the equivalent hydraulic conductivity of rock mass around a tunnel. Int J Rock Mech Min Sci 54:125–128. https://doi.org/10.1016/j.ijrmms.2012.05.017 |
[258] | Zoorabadi M, Saydam S, Timms W, Hebblewhite B (2022) Analytical methods to estimate the hydraulic conductivity of jointed rocks. Hydrogeol J 30:1. https://doi.org/10.1007/s10040-021-02413-0 |
21 September 2023
30 November 2023
13 June 2024
November -0001
https://doi.org/10.1007/s40789-024-00716-7