Download as pdf or txt
Download as pdf or txt
You are on page 1of 55

Page 1 of 55

For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

1 Experimental study on deformation and failure of a mining slope under the action

2 of rainfall

3 Tongwei Taoa, Wenbing Shib,c (), Shaozheng Xiongb, Feng Liangb,c, Jiayong Zhangd

4 aKey Laboratory of Karst Georesources and Environment, Ministry of Education, Guizhou

5 University, Guiyang 550025, China


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

6 bCollege of Resources and Environmental Engineering, Guizhou University, Guiyang 550025,

7 China

8 cMountain Geohazard Prevention R&D Center of Guizhou Province, Guiyang 550025, China

9 dGuizhou Geological Environment Monitoring Institute, Guiyang, Guizhou 550004, China

10 Corresponding author: Wenbing Shi (email: wbshi@gzu.edu.cn)

11 Total word count: 9593

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 2 of 55

12 Abstract: On August 28, 2017, a mining slope collapse in Pusa Village, Guizhou Province,

13 China, resulted in the release of approximately 8,600 m3 of earth and rock, leading to 35

14 fatalities. The deformation and failure mechanisms of the mining slope under the influence of

15 rainfall was studied by a physical model test, using the Pusa Village collapse as an example.

16 Data were collected through strategically placed monitoring points in a physical model,
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

17 forming the basis for a quantitative analysis of slope deformation and failure patterns. The

18 tested results revealed that sequential mining and rainfall activities induce progressively

19 increasing stress variations, exhibiting a cyclical pattern of stress concentration, relaxation,

20 and stabilization. Mining activities initiated the formation of tension and separation fractures

21 within the model, which were subsequently exacerbated by rainfall. Rainwater infiltrating

22 through these fractures and geotechnical bodies resulted in a more pronounced increase in

23 water content in the upper part of the model than in the lower part. Displacement values

24 within the model progressively increased with ongoing coal seam mining, primarily occurring

25 in the mining area's roof and extending to the model's top following rainfall. The failure mode

26 of the Pusa slope under combined mining and rainfall conditions is characterized by the

27 following sequence: bending-pulling-subsidence-creep-dumping. The deformation and failure

28 processes can be categorized into four stages: mining-induced disturbances, fractures

29 extension and extension, formation of a potential collapse surface, destabilization faliure.

30 Keywords: Deformation and failure; Physical model; Collapse; Mining; Rainfall

© The Author(s) or their Institution(s)


Page 3 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

31 1. Introduction

32 Deformation in the overlying rock of mined-out areas, known as mining slope

33 deformation, is a direct or indirect result of underground mining activities, commonly referred

34 to as mining landslides (Xu et al. 2012; Zhao et al. 2016; Dai et al. 2021). Unlike general
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

35 landslides, mining landslides are fundamentally distinct and have garnered worldwide

36 attention due to their considerable risks (Stephen et al. 2006; Kwan et al. 2018; Yu et al.

37 2020). Accelerated by China's booming economy and rising energy demands, the frequency

38 of mining activities and subsequent landslides has increased significantly (Zhang et al. 2020;

39 Chen et al. 2021; Liu et al. 2022; Yang et al. 2022a). Rainfall serves as an additional

40 destabilizing factor, exacerbating the impact of mining activities. Consequently, numerous

41 mining slope destabilization failures have occurred, leading to significant loss of life and

42 property (Hu et al. 2013; Xu et al. 2016; Ma 2017; Locat et al. 2017; Zhang et al. 2022; Tao et

43 al. 2022). The karst region in southwest China is characterized by complex geological

44 features and strong tectonic activity, with soluble rock slopes being widely prevalent. In these

45 mining areas, noticeable karst development and a dense network of joints and fissures are

46 observed on the mining slopes. The topography of these slopes generally exhibits a pattern of

47 being “steep (and hard) on top and gentle (and soft) on the bottom” (Xing et al. 2014; Li et al.

48 2016a; Zheng et al. 2018). These geological and topographical factors, in combination with

49 external factors, such as frequent mining activities and heavy rainfall, create conditions

50 conducive to the collapse of mining slopes and ensuing landslides (Li et al. 2018; Sun et al.

51 2021, 2022; Yang et al. 2022b). Since the 20th century, several landslides in southwest China

52 have been triggered by subterranean mining and rainfall, including but not limited to the Jiwei

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 4 of 55

53 Mountain landslide in Wulong, Chongqing (Feng et al. 2016), Guanling landslide in Guizhou

54 (Zhang et al. 2023a), Zhenxiong landslide in Yunnan (Yin et al. 2017), and Pusa collapse in

55 Guizhou (Zhang et al. 2023b). These events have resulted in severe casualties and have drawn

56 significant social attention. Therefore, an in-depth analysis of the deformation and failure

57 mechanisms of mining slopes influenced by rainfall is of immense practical and


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

58 socio-economic value.

59 Frequent underground mining activities are highly likely to trigger the collapse of the

60 roof in the mined-out areas, giving rise to a “cantilever beam effect” in the rock mass. This

61 effect creates cracks that serve as advantageous pathways for rainwater infiltration. Rainwater

62 permeates the interior of slopes via geotechnical bodies and mining fractures, elevating the

63 saturation levels within these structures. As saturation increases, geotechnical bodies absorb

64 water, leading to an increase in pore water pressure and a decrease in matric suction, thereby

65 reducing the shear strength of geotechnical bodies (Rahimi et al. 2011; Gasmo et al. 2000; Ng

66 et al. 1998, 2003). When slopes reach a critical condition, they become increasingly prone to

67 destabilization over time (Ma et al. 2018; Fernandez et al. 2020). In recent years, a

68 combination of physical model tests and numerical simulations, informed by field geological

69 investigations, has emerged as an effective approach for studying slope stability issues

70 (Mitchell 2011; Scaringi et al. 2018). For example, Jones et al. (1991) used physical model

71 tests to explore landslides induced by underground mining and found that such activities

72 significantly compromised slope stability. Similarly, Tang et al. (2019) performed physical

73 model tests on the Jiwei Mountain landslide in Wulong, Chongqing, China, and proposed a

74 deformation mechanism for it. When adhering to basic scaling laws, physical model tests can

© The Author(s) or their Institution(s)


Page 5 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

75 accurately depict the kinematic characteristics and failure patterns of slopes (Li et al. 2016b;

76 Ray et al. 2020), while also accounting for external factors such as rainfall (Giovanna et al.

77 2020) and underground mining (Dai et al. 2021).

78 Although numerous studies have investigated the failure processes and fracture

79 characteristics of mining slopes, less attention has been paid to the stress and displacement
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

80 variations between geotechnical bodies during mining, as well as the impact of rainfall on

81 mining slopes. To address this gap, the present study employs a physical model test to

82 examine the Pusa collapse in Guizhou as a case example. We investigate changes in stress,

83 fracture development, displacement, and water content within the rock mass during mining

84 and rainfall activities. The study aims to elucidate the failure mechanisms and evolution

85 processes of the Pusa collapse through a four-pronged analysis focusing on stress, fractures,

86 displacements, and water content. The findings offer valuable insights for future research on

87 similar types of slopes in Guizhou, China, and contribute to guiding disaster prevention and

88 mitigation efforts.

89 2. Regional setting

90 2.1 Geological setting and mining activities

91 The study area is situated in the northwestern part of the Guizhou Plateau, featuring

92 lower topography in the north and higher elevations in the west. Mountain ridges

93 predominantly extend in a southwest direction, exhibiting a topography that is “steep at the

94 top and gradual at the bottom.” The steepest slope angle is 70°, whereas the gentler sections

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 6 of 55

95 range between 10–25°, resulting in an average slope angle of 58°. The highest point, Little

96 Eagle Rock, is located in the south with an elevation of 2,175 m, whereas the lowest point lies

97 near the river ditch below the village in the northwest at 1,875 m, making the elevation

98 difference approximately 300 m. The area falls on the southeastern side of the Zhangwei

99 anticline and contains three identified faults (F1, F2, and F3) that have no direct impact on the
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

100 coal seams or the collapse zone (Figure 2) (Wang et al. 2022a; Zhang et al. 2022). Site

101 investigations revealed dissolution fissures and joints at the trailing edge of the collapsed area.

102 These well-developed fissures and joints have fragmented the geotechnical bodies, with the

103 spaces filled with rock debris, leading to poor cementation between the geotechnical bodies

104 and considerable structural failure of the slope (Figures 1b and c).Mining-induced fractures

105 are evident at the trailing edge of the collapse, indicating that rainwater had infiltrated the

106 rock mass, thereby weakening its cohesive strength and further compromising slope stability.

107 Additionally, several karst collapses are observable at the summit, primarily manifesting in

108 the Quaternary residual slope deposits (Q4). These karst collapses are generally round or

109 nearly round in shape, with the largest having a diameter of 4 m (Figures 1d and e).

110 The general rock angle in the area is 200°∠9°. The exposed stratigraphic layers are as follows

111 (Figure 3):

112 ① Quaternary (Q4): Found at the top and gentler parts of the slope, the lithology is

113 composed of clay and sandy clay, with varying thickness ranging from 0 to 15 m.

114 ② Lower Triassic Yelang Formation (T1y): Located in the middle and upper parts of the

115 slope, the stratum is subdivided into two sections. The upper section (T1y2) is primarily

116 limestone, gray in color, and approximately 50 m thick, featuring more developed karst

© The Author(s) or their Institution(s)


Page 7 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

117 fissures. The lower section (T1y1) is predominantly muddy siltstone, purple in color, and

118 approximately 70 m thick.

119 ③ Upper Permian Changxing-Dalong Formation (P3c+d): Occupying the lower part of

120 the slope and mainly consisting of gray limestone, with a thickness of approximately 18 m.

121 ④ Upper Permian Longtan Formation (P3l): This layer contains a series of coal
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

122 strata—mudstone, siltstone, coal—distributed in the lower part of the slope. A total of 26-44

123 coal seams are identified, among which six—M6, M10, M14, M16, M18, and M20—are

124 considered recoverable. The average thicknesses of the previously mined M16, M14, and

125 M10 seams are 1.48, 1.23, and 2.12 m, respectively (Table 1).

126 The coal seams at the site were mined in the sequence M16-M14-M10, from bottom to

127 top, utilizing consistent mining methods for all three seams: the long-wall coal mining method

128 and gunnery excavation. Most of the M16 seam was extracted prior to 2012, with significant

129 fractures and subsidence observed on the trailing edge of the slope between September and

130 December of 2011. The M14 and M10 seams were mined from 2013 to 2017. Throughout this

131 period, fractures noticeably expanded, and occasional collapses of the slope surface occurred.

132 This establishes a temporal correspondence and causal relationship between the fractures at

133 the trailing edge of the slope and underground mining activities (Yang et al. 2022c). Mining

134 operations at the site concluded in August 2017, by which time seams M16, M14, and M10

135 were essentially depleted. The mined-out area is illustrated in Figure 2.

136 2.2 Characteristics of the collapse

137 On August 28, 2017, a large-scale slope collapse occurred in Pusa Village, located in the

138 Zhangjiawan township of Nayong County, Guizhou Province (26°38′05″E, 105°26′56″N).


7

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 8 of 55

139 The disaster resulted in 26 fatalities, 9 missing persons, and 8 injuries (Fan et al. 2019; Cui et

140 al. 2022). Typical of slope destabilization failure, the rock in the collapse’s origin area was

141 initially destabilized. Subsequent downward movement dislodged loose material in the middle

142 section of the slope and traveled in the direction of 305° as a debris flow. A minor ridge at the

143 leading edge split the rapidly descending mass into two separate flows. One part of the
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

144 collapsed material bypassed the ridge and accumulated at the base of the slope. Field surveys

145 and UAV imagery divide the collapse area into three distinct areas: the collapse source area

146 (Ⅰ), the scraping area (Ⅱ), and the accumulation area (Ⅲ). The elevation of the trailing edge

147 of the collapse is approximately 2,120 m, whereas the base of the slope stands at

148 approximately 1,922 m, resulting in a relative elevation difference of approximately 200 m.

149 The horizontal distance from the collapse source to the front of the accumulation area

150 measures 780 m. The average thickness of the accumulation body is 4 m, with a cubic volume

151 of 8,600 m3, and the maximum width of the accumulation area spans 480 m (Figure 1a).

152 2.3 Rainfall

153 The study area experiences a humid subtropical monsoon climate, marked by relatively

154 abundant rainfall and an average annual temperature ranging between 12 and 15.0°C.

155 Influenced by the monsoon climate, the area sees most of its rainfall during the summer

156 months, with reduced precipitation in winter. The slope collapse occurred in August,

157 following a period of heavy rainfall. Annual rainfall monitoring data indicate that June

158 typically experiences the highest precipitation levels, reaching up to 223 mm, whereas

159 December has the lowest at 22 mm. The monthly rainfall distribution in Nayong County for

160 2017 was as follows: January to May, 161.6 mm; June, 228.7 mm; July, 162.3 mm; and
8

© The Author(s) or their Institution(s)


Page 9 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

161 August’s daily rainfall is illustrated in Figure 4.

162 3. Physical model design

163 3.1 Similar proportions and similar materials


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

164 Maintaining the same size, material, rainfall and test conditions as the prototype is

165 difficult to achieve during physical model testing. Therefore it is only possible to obtain a

166 behavior similar to the real situation by reducing the size of the prototype and keeping certain

167 parameters as similar as possible to those of the prototype. However it is not possible to keep

168 all parameters of the physical model similar to the prototype (Pei et al. 2018; Dai et al. 2021;

169 Yang et al. 2022; Lai et al. 2023; Zhou et al. 2023). Therefore, when we conduct physical

170 model tests, we will try to make some parameters as similarly proportioned as possible to the

171 prototype, depending on the purpose of the study.

172 To comprehensively analyze the stress, fracture development, water content, and

173 displacement characteristics of the slope under mining and rainfall conditions, and to

174 elucidate the deformation and failure mechanisms, a scaled physical model of the Pusa

175 collapse was constructed based on field investigation data (Figure 5). Utilizing similarity

176 theory, the model replicates key processes of the actual event in proportion (Zheng et al. 2015;

177 Tao et al. 2020). The model’s similarity constants are determined based on three aspects of

178 similarity theory: geometric, mechanical, and physical properties. Geometric similarity

179 (expressed as CL) refers to the proportional relationship in size and shape between the model

180 and its actual prototype. Mechanical similarity (Cγ) signifies that the model and the prototype

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 10 of 55

181 share corresponding mechanical properties, such as unit weight. Physical property similarity

182 (Cσ) indicates comparable material properties between the model and the prototype, such as

183 compressive strength, cohesion, and internal friction angle. In this study, the geometric

184 similarity constant CL was set at 400. Accordingly, the model’s dimensions were configured

185 as follows: length 160 cm, width 50 cm, and height 110 cm, with a slope angle of 58 °
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

186 (Figure 5). As gravitational effects in both the model and the actual prototype are equivalent,

187 Cγ is set to 1. The Cσ value is calculated from CLCγ as Cσ = CL ∙ Cγ (Zhang et al. 2005). After

188 calculations, CL is confirmed to be 400. Although it is challenging to maintain identical

189 physical properties between the model and the prototype (Dong et al. 2020; Wang et al.

190 2022b), this study primarily focuses on examining the failure mechanisms of the slope. The

191 deformation and failure of slope are mainly controlled by the compressive strength of rock

192 body; therefore, the similarity index for physical properties in this experiment is based on

193 compressive strength.

194 Based on the information gathered, the fault in the study area did not exert a direct

195 influence on the observed slope collapse. As such, its effects were not incorporated into the

196 physical model. However, rock joints, which play a significant role in slope deformation and

197 failure, were included as model variables. These joints were simulated using mica sheets,

198 which are easily manipulated. The spacing and trace length of these simulated joints were

199 configured based on the established geometric similarity ratio (Figure 6g). The model box,

200 rainfall simulation equipment, and data collection devices are depicted in Figure 6, b-f.

201 For material composition, the physical model was designed to emulate the lithology of

202 the study area, predominantly consisting of limestone, argillaceous siltstone, mudstone, and

10

© The Author(s) or their Institution(s)


Page 11 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

203 coal. In the actual prototype, the stratigraphic thicknesses of the upper section of the Yelang

204 Formation (T1y2), the lower section of the Yelang Formation (T1y1), and the Upper Permian

205 Changxing-Dalong Formation (P3c+d) are 50, 70, and 18 m, respectively, according to the CL

206 equal to 400, the thicknesses corresponding to the lithologies of the physical model can be

207 calculated to be 12.5, 17.5, and 4.5 cm (Figure 6a). Additional substances used in the model
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

208 were categorized as either aggregates or binding materials. Quartz sand and barite powder

209 served as the aggregates, whereas gypsum and cement functioned as binding materials. Borax

210 was added to more closely mimic the properties of the actual materials and to explore the laws

211 governing slope deformation failure. To facilitate the simulation of coal seam excavation,

212 sand was used as a substitute for coal in the model. To ascertain the density and compressive

213 strength of limestone, argillaceous siltstone, mudstone, and coal seams, field samples were

214 initially collected. These were then converted into standard cylindrical specimens with

215 dimensions of Ф50 mm × 100 mm for indoor tests, which included density measurements and

216 uniaxial compression tests. Upon obtaining the parameters of the actual samples, similarity

217 material ratios were determined through orthogonal testing. Subsequently, additional standard

218 cylindrical samples of similar materials with dimensions Ф50 mm × 100 mm were created.

219 These were also subjected to density and uniaxial compression tests to finalize the selection of

220 the most suitable material ratios for the physical model. Mechanical parameters and the

221 compatibility of each stratum with similar materials are detailed in Table 2.

222 3.2 Mining and rainfall schemes

223 To simulate coal seam mining activities, sand was used to represent the coal seam. The

224 sand was packed into plastic bags, which were then arranged with intervening coal pillars to
11

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 12 of 55

225 support overburden pressure. The mining process involved cutting open the plastic bag and

226 using a small wooden stick to remove the sand, effectively creating a mined-out area. This

227 simulation method was designed to replicate the mining sequence observed in the actual coal

228 seams: M16, M14, and M10, in ascending order. Specifically, M16 consisted of three mining

229 areas, whereas M14 and M10 each had two. The dimensions for the mining steps were as
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

230 follows: the first step for M16① was 18 cm long, M16② was 20 cm long, M16③ was 20

231 cm long, M14① was 20 cm long, M14② was 20 cm long, M10① was 25 cm long, and

232 M10② was 20 cm long (Figure 6a).

233 The aim of this study is to investigate the deformation and failure mechanisms of the

234 slope under the influence of rainfall. To ensure a pronounced rainfall effect in the physical

235 model, rainfall intensity was determined based on the cumulative precipitation observed in the

236 month preceding the actual event. The cumulative rainfall during this period was categorized

237 into three levels: 25, 100, and 175 mm (Figure 4). Correspondingly, rainfall intensities of 30

238 mm/day, 60 mm/day, and 140 mm/day were selected for the model tests. The first phase of

239 the rainfall simulation began with an increment from 0 mm/day to 30 mm/day; upon reaching

240 30 mm/day, this level was maintained for a few hours before halting the rain. The second

241 phase escalated from 0 mm/day to 60 mm/day and was also held for several hours before

242 ceasing the rain. The final phase ramped up from 0 mm/day to 140 mm/day and continued at

243 this intensity until the slope collapsed.

244 3.3 Data monitoring and collection

245 The monitoring of stress changes, fracture development, and water content commenced

246 as soon as coal seam excavation and rainfall simulation were initiated. Periodic photographs
12

© The Author(s) or their Institution(s)


Page 13 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

247 were taken to chronicle the temporal evolution of fractures. Stress monitoring points were

248 established in each stratum and at the roofs of coal seams, with 10 points in total. Stress data

249 were collected at intervals of 10 s. The depths at which stress monitoring points T1-T10 were

250 buried are 17, 18, 30, 29, 49, 48, 71, 84, 83, and 97 cm, respectively (Figure 6a). Stress

251 sensors were supplied by Jiangsu Test Equipment Manufacturing Co., Ltd., featuring an
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

252 accuracy of ±0.025%FgS , a measuring range of 0–5 Mpa, and a response time of less than 1

253 s. By monitoring water content across various rock layers, the impact of rainwater infiltration

254 on slope stability could be assessed. Five water content monitoring points were established,

255 and data were collected at 10 s intervals. These points, designated S1–S5, were buried at

256 depths of 18, 29, 48, 63, and 82 cm, respectively. The water content sensor, provided by

257 Beijing Heng Rui Chang Tai Technology Co., Ltd, has an accuracy of 2% , a measuring

258 range of 0–100%, and a response time of less than 1 s.

259 3.4 Principle of DPDM technology

260 Digital Picture Deformation Measurement (DPDM) is an image-based measurement

261 technique developed by the China University of Mining and Technology. This method is

262 employed to analyze and process the deformation and trend changes of specified points within

263 an image (Li et al. 2016b). DPDM employs techniques such as image correlation, the

264 center-of-mass method, bilinear interpolation, and quadratic isoparametric unit image

265 calibration to achieve sub-pixel level accuracy in deformation measurements (Li et al. 2019).

266 Its accuracy can be quantified using Eq. (1). In the scale-free method, complex deformation

267 patterns are distilled into two basic forms: translation and rotation. A three-step search

268 algorithm—comprising coarse search, fine translation search, and fine rotation search—is
13

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 14 of 55

269 utilized in conjunction with image correlation. Due to its accessible equipment,

270 straightforward operation, and generalizability, DPDM is increasingly recognized as an

271 essential measurement technique in contemporary experimental mechanics.

Target width
272 Accuracy= (1)
Image width
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

273 In Eq. (1), the target width refers to the model’s dimensions, whereas the image width

274 pertains to the pixel count of the image. Given a target width of 1600 mm and an image width

275 of 3000 pixels, the accuracy is calculated to be 0.53 mm/pixel using Eq. (1).

276 The punctuation-free method employs correlation analysis of speckle images as its

277 foundational principle. It determines the coordinate changes of measurement points before

278 and after deformation by assessing the image correlation within regions where the pixel

279 measurement points are located on a digital photo sequence (Figure 7). Specifically, the

280 method calculates the correlation coefficient for all points (Pd) within a defined search range

281 on the deformed image, relative to a reference point (Pi) on the pre-deformation image. The

282 pixel correlation coefficient is determined using Eq. (2). The point with the highest

283 correlation coefficient is identified as the target point. The displacement of the measurement

284 point is then obtained by calculating the coordinate difference between the reference and

285 target points, and strain calculations are subsequently conducted using finite element methods

286 (Li et al. 2019).

2 k 1 2 k 1

  v ( x, y )  u ( x, y )
x 1 y 1
287 R12  (2)
2 k 1 2 k 1 2 k 1 2 k 1

  v ( x, y )    u ( x, y )
x 1 y 1
2

x 1 y 1
2

288 where R12 is the correlation coefficient between the reference point pixel block and the target

14

© The Author(s) or their Institution(s)


Page 15 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

289 point pixel block, (vi)x,y is the pixel RGB color value of the tracking point, (ui)x,y is the pixel

290 RGB color value of the reference point, 2k+1 is the length or width of the pixel block and the

291 unit is pixel, and i has values 1, 2, 3 and represents the 3 color components of RGB. (Note:

292 The final correlation coefficient is to calculate the respective correlation coefficients for the

293 three R, G, and B color components of, and then calculate the average value.)
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

294 4. Results

295 4.1 Variation law of the stress

296 A total of 10 stress monitoring points were established in the experiment, as shown in

297 Figure 6a. Both coal seam mining and rainfall activities disrupt the original stress equilibrium

298 within the slope. Importantly, the stress values discussed herein are specific to the model’s

299 monitoring points and should not be conflated with the stresses in the actual geological

300 formation—particularly in numerical terms. Nonetheless, variations in stress at these

301 monitoring points can provide insights into the stress behavior of the real-world prototype to

302 some extent. Therefore, we can deduce the deformation and failure mechanisms of the actual

303 slope from these stress variations (Figure 8). Prior to the experiment, the initial stress values

304 at monitoring points 1 through 10 were as follows: 18.2, 18.1, 15.9, 15.5, 12.7, 12.3, 9.5, 6.2,

305 6.0, and 1.0 kPa.

306 During the mining of the M16 coal seam, stress levels at points T3, T5, T6, and T7

307 remained stable from 0 to 8 h. This stability indicates that the early stages of M16 mining did

308 not significantly disturb the model, and the minor fluctuations in stress did not impact these

15

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 16 of 55

309 specific monitoring points. Conversely, the stress levels at points T1, T2, T4, T9, and T10

310 decreased by 0.6, 4.5, 2.3, 1.0, and 1.0 kPa, respectively, within the same timeframe. This

311 reduction was attributed to stress relaxation caused by the collapse of the M16 roof plate and

312 the resulting unloading of the rock mass. At the 10 h, stress values at points T2 and T4

313 suddenly surged by 6.1 and 5.0 kPa, respectively. This increase was attributed to a subsequent
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

314 collapse of the M16 roof plate, which led to subsidence in the upper part of the rock mass and

315 consequently to stress concentration at these points. Between 10 and 30 h, all monitoring

316 points exhibited minor stress fluctuations but largely maintained their previous trends.

317 Analysis suggests that the model was beginning to fracture during this period. Between 30

318 and 40 h, all monitoring points reached a state of stress equilibrium. This equilibrium was due

319 to stress realignment following the disturbance induced by the M16 mining activities.

320 During the 40 to 58 h timeframe of mining M14, stress values at points T1, T3, T4, T5,

321 T6, T7, T9, and T10 remained largely unchanged. This suggests that mining of the M14 seam

322 had minimal impact on the model’s stress distribution during this period. However, at 58 h,

323 the stress level at point T4 surged dramatically by 7.5 kPa, attributable to the collapse of the

324 M14 roof plate, which led to stress concentration at this specific location. Concurrently,

325 points T7, T8, and T9 exhibited minor stress increases of 0.9, 1.2, and 0.3 kPa, respectively. It

326 is hypothesized that these changes are a result of off-layer fractures occurring in molded Top

327 (II) and the extrusion of the rock layer. From 58 to 90 h, the stress levels at points T7, T9, and

328 T10 decreased by 1.5, 1.4, and 1.4 kPa, respectively. This decrease can be attributed to

329 ongoing small-scale block dropping, delamination, and collapse of M14’s top slab, facilitated

330 by a weakening of the support in the lower part of Top (II). In contrast, other monitoring

16

© The Author(s) or their Institution(s)


Page 17 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

331 points remained stable, indicating that the conclusion of mining activities in M16 and M14

332 would not induce widespread changes in the model’s stress patterns. During the 90 to 108 h

333 period of mining M10, stress levels at points T3, T6, T7, T8, and T9 increased by 0.5, 0.4, 0.4,

334 0.6, and 0.5 kPa, respectively. Conversely, stress levels at points T1, T2, T4, and T5

335 decreased by 0.3, 0.9, 0.3, and 0.6 kPa, respectively. These changes are assumed to be caused
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

336 by the mining of M10, which led to extensive fracturing around Bottom (III) and consequent

337 stress relaxation. At the same time, Top (II) experienced subsidence, resulting in stress

338 concentration due to geotechnical bodies squeezing each other. The stress at point T10

339 remained stable, suggesting that mining of M10 had not yet impacted the model’s topmost

340 layer. From 108 to 115 h, stress at point T8 increased by 1.3 kPa, whereas points T2, T3, T4,

341 T5, T6, T7, and T9 exhibited stress decreases of 0.6, 0.3, 0.5, 0.6, 0.2, 1.5, and 0.5 kPa,

342 respectively. The decrease in stress at these points indicates a large-scale collapse of the M10

343 roof plate, whereas the stress increase at T8 suggests that the model was undergoing

344 deformation towards a free face. Between 118 and 128 h, stress levels at all monitoring points,

345 except T10, increased. This was attributed to ongoing deformation within the model and

346 downward settlement of the top rock layer, resulting in stress concentration. At the 131 h

347 mark, significant changes in stress were observed at all monitoring points, indicating the

348 expansion of the collapse area and suggesting that the model had entered a stage of significant

349 deformation.

350 Following the rainfall, the roof of the mined-out area collapsed at 142 h, prompting

351 varying degrees of stress changes at all monitoring points. Specifically, the stress values at T7,

352 T9, and T10 surged by 6.1, 6.5, and 6.3 kPa, respectively, whereas stress at T8 declined by

17

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 18 of 55

353 3.5 kPa. These changes are attributed to a significant roof plate collapse that impacted even

354 the top of the slope, as well as the deformation of the modeled Leading Edge (IV) toward the

355 free face. Between 142 and 154 h, stress values at all monitoring points increased, suggesting

356 a global collapse of the model. It is hypothesized that the rainfall led to increased material

357 saturation, reduced matric suction, and elevated pore water pressure, collectively diminishing
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

358 the shear strength and causing the model to collapse. From 154 to 181 h, the stress levels at

359 the monitoring points stabilized, indicating a temporary equilibrium in the model following its

360 global collapse. At 181 h, stress levels at T2 and T8 declined sharply by 11.2 and 5.1 kPa,

361 respectively. This was due to another subsidence event in the roof of the mined-out area,

362 resulting in rock mass failure at T2 and T8 and the detachment of Leading Edge (IV). After

363 181 h, the model underwent large-scale displacement, culminating in a slope collapse at 213 h.

364 Notably, the stress at T2 exhibited significant changes after 181 h, presumably owing to the

365 infiltration of rainwater at this location.

366 Based on the aforementioned analysis and Figure 8, a discernible pattern emerges in

367 stress changes, characterized by cycles of stress concentration, stress relaxation, and stress

368 equilibrium. These monitored stress variations provide an authentic and reliable

369 representation of the internal stress dynamics within the actual prototype. Consequently, it can

370 be inferred that the actual prototype, influenced by both mining activities and rainfall,

371 undergoes sequential phases of stress concentration (due to roof collapse and rock extrusion),

372 stress relaxation (through the creation of fractures and rock layering), and stress equilibrium

373 (as the actual slope attains a state of temporary balance).

18

© The Author(s) or their Institution(s)


Page 19 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

374 4.2 Fracture evolution

375 Following the mining of M16, subsidence occurred in the overburden of the mined-out

376 area, leading to the emergence of fractures in the coal seam’s roof. Two conspicuous

377 delaminated fractures, D1 (48 cm long, 1 mm wide) and D2 (28 cm long, 1 mm wide),
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

378 developed in the upper plate. Fractures D3 (52 cm long, 1 mm wide) and D4 (28 cm long, 0.8

379 mm wide) developed in the base plates of M14 and M10, respectively. In addition, two

380 prominent tensile fractures, T1 and T2, were observed. T1, measuring approximately 28 cm in

381 length and 1 mm in width, progressively extended upward to intersect with separation

382 fractures and smaller tensile fractures. T2, situated 22 cm directly above M16①, had

383 dimensions of 11 cm in length and 1 mm in width. Concurrently, a network of small-scale,

384 interconnected tensile fractures emerged near M16, culminating in the collapse of the upper

385 plate rock. The analysis indicates that the loss of support from M16’s overburden led to rock

386 unloading, triggering fractures once the limit balance state was surpassed (Figure 9a).

387 Upon the completion of M14 mining, the upper plate exhibited a tendency to collapse

388 downward due to a loss of support across a broad area. This resulted in the appearance of new

389 tensile fractures (T3, T4, T5) and a separation fracture (D5) in the M14 upper plate, as well as

390 new separation fractures (D6 and D7) in the M10 upper plate. The length of the D5 fracture

391 extended to approximately 45 cm, matching the length of M14, and its opening width

392 expanded to 5 mm under the rock's own weight. The onset of D5 led to a partial collapse of

393 M14's upper plate. Fractures D6 and D7, approximately 24 cm long, extended horizontally.

394 The slope model exhibited a propensity for deformation and displacement toward the free

395 face, causing an increase in stress in the upper part of the left side of M10. This led to the

19

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 20 of 55

396 formation of a new tensile fracture (T6), measuring 11.2 cm in length and oriented nearly

397 vertically, extending to Top (II) of the model (Figure 9b).

398 Following the completion of M10 mining, slope deformation and failure intensified,

399 culminating in the downward bending and settling of the overburden to create an “arch

400 collapse zone.” The impact of M10 mining was nearly model-wide, with pronounced
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

401 development of mining fractures. This led to the formation of multiple parallel separation

402 fractures from the bottom to the top of the model, along with additional tensile fractures in its

403 middle and upper parts, collectively constituting a "fracture zone.” The extension lengths of

404 the separation fractures in the upper part of M10 ranged from 35 to 45 cm, with widths

405 varying between 1 and 3 mm. Over time, T6 lengthened downward to intersect with a fracture

406 near M10, whereas T1, located at the base of M14, expanded horizontally. The development

407 of T6 is interpreted as a result of the cantilever beam effect, with Rear (I) serving as the fixed

408 end and Top (II), Bottom (III), and Leading Edge (IV) acting as free ends. This led to the

409 fracture occurring at position T6 under gravitational forces (Figure 9c).

410 Upon stabilization of the model's deformation due to mining activities, rainfall was

411 applied. The rainwater infiltrated the model's slope along existing fractures. As anticipated,

412 the erosive capability of the rainwater facilitated the extension and interconnection of

413 fractures within the model's slope. Continuous deformation of the middle and lower

414 geotechnical bodies in the model led to the development of fractures in the shoulder of the

415 slope. Notably, although rainfall did encourage the extension and penetration of fractures, the

416 initial formation of these fractures was primarily driven by mining activities (Figure 9d).

20

© The Author(s) or their Institution(s)


Page 21 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

417 4.3 Rainfall infiltration

418 After stabilizing the stress deformations induced by simulated mining, a rainfall

419 experiment was conducted. The rainfall was divided into three periods with varying lengths

420 and intensities. The initial period extended from 139 to 142 h with an intensity of 30 mm/day,
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

421 the second from 182 to 186 h at 60 mm/day, and the third from 210 to 215 h at 140 mm/day.

422 By tracking the changes in water content across various rock layers, we were able to

423 comprehend the rainwater infiltration process within the slope, and subsequently, the

424 deformation and failure mechanisms of the mined slope under the influence of rainfall. Due to

425 variations in layer stacking and material proportions, initial water content differed upon

426 model assembly. Monitoring points for water content, labeled S1 to S5, were distributed from

427 the bottom to the top of the model. Specifically, S1 to S3 were positioned at Bottom (III),

428 whereas S4 and S5 were at Top (Ⅱ), as depicted in Figure 6a. The overall trends for water

429 content changes at points S4 and S5 post-rainfall were generally consistent, as were those at

430 points S1, S2, and S3 (Figure 10). Upon the first rainfall event, water content at S4 and S5

431 quickly escalated, reaching 22.2% and 25.3% respectively. In contrast, S1, S2, and S3 showed

432 more modest increases. These locations, S4 and S5, had a higher density of separation and

433 tension fractures due to mining activities, allowing for rapid rainwater infiltration compared

434 to other parts of the slope. Upon cessation of the first rainfall at 142 h, the water content at all

435 monitored points swiftly declined and stabilized, a shift attributed to the material's water

436 absorption capacities. The stabilized values for S1 to S5 were 12.0, 14.2, 17.2, 19.0, and

437 22.0%, respectively. The reaction rates across all monitoring points were faster during the

438 second rainfall due to its higher intensity of 60 mm/day. Additionally, the first rainfall event

21

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 22 of 55

439 had expanded and penetrated fractures, enhancing the subsequent rainwater infiltration rate.

440 As a result, peak water content levels rose relative to the first rainfall, reaching 16.2, 18.3,

441 19.2, 23.4, and 25.2% for S1 through S5, respectively. During the third rainfall event, with an

442 intensity of 140 mm/day, peak water content levels increased yet again, attaining values of

443 18.6, 20.0, 21.3, 27.0, and 28.8% for S1 to S5, respectively. The water content increase at S3
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

444 paralleled those observed at S4 and S5, suggesting that the downward progression of upper

445 fractures had effectively channeled rainwater infiltration at this particular monitoring location.

446 4.4 Variation of displacement

447 DPDM technology was employed to investigate the variations in the model slope

448 displacement due to mining and rainfall activities. Image processing was feasible only for the

449 pre-collapse state of the model slope because the extensive post-collapse displacement

450 resulted in significant mesh deformation, making DPDM-based image processing

451 challenging.

452 During the mining of M16, deformation was predominantly observed in the roof of the

453 mined-out area, corroborating field observations of roof collapse and separation fractures. As

454 mining progressed across the three mined-out areas of M16, the displacement values showed

455 a gradual increment (Figure 11a–d). Specifically, upon mining the third section (M16③), the

456 maximum displacement spiked to 68.47 mm. Conversely, the mining of M14 resulted in

457 negligible large-scale displacement within the model. Although separation and tension

458 fractures emerged in the roof plate of the M14 mined-out area, as well as in the overlying rock

459 layer and slope foot of M10 (Figure 11e–f), the roof plate itself remained intact. Thus, we

460 infer that the mining of M14 did not induce significant model displacement. When mining
22

© The Author(s) or their Institution(s)


Page 23 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

461 M10, the deformation of the overburden rock in the mined-out area intensified under the

462 action of unloading. The maximum displacement observed in the roof plate in the mined-out

463 area of M10 reached 40 mm, and the affected range extended into the central region of the

464 model mass. Further, the displacement at the slope foot increased due to advanced fissuring

465 triggered by M10 mining (Figure 11g–h). Initial coal seam mining induced only minimal
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

466 model displacement. However, as the mined-out area expanded, both the range and magnitude

467 of displacements increased progressively. Most notably, the roof plate displacement in the

468 mined-out area exhibited a marked rise. During the rainfall event, the displacement range of

469 the model broadened significantly, even reaching the slope's top. This is considered a critical

470 phase in the extensive deformation of a mining slope. At the leading edge of the slope, tension

471 fissures began shifting toward the free face, indicating the onset of slope mass collapse

472 (Figure 11i).

473 5. Analysis of deformation and failure mechanism

474 In summary, the study area exhibits a layered structure of “hard at the top and soft at the

475 bottom” and “steep at the top and gentle at the bottom,” and the slope rocks are intersected by

476 joints conducive to dissolution (Figure 12a–b). Mining operations disrupt the slope's stress

477 equilibrium, generating an extensive fracture network in the overburden and adjacent to the

478 mined-out area. Monitoring data from the physical model reveal that stress redistributes

479 upward over time, resulting in fractures at the slope's trailing edge. These fractures

480 progressively extend downward to intersect with the mined-out area, creating potential sliding

481 planes within the slope (Figure 12c–d). During rainfall events, rainwater infiltrates the slope
23

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 24 of 55

482 through both the geotechnical bodies and existing mining fractures. This leads to increased

483 saturation levels in the slope's geotechnical bodies. As rainwater penetrates deeper, alterations

484 in pore water pressure trigger deformation failure mechanisms within the slope, while also

485 exacerbating the spread of mining-related fractures toward the free face. As the overlying

486 rocks in the modeled mined-out area keep collapsing and extruding, resulting in the closure of
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

487 fractures in the lower portion of the model, rainwater continues to infiltrate and raise.

488 Concurrently, ongoing rainwater infiltration elevates the water table within the slope. Once

489 the geotechnical bodies reaches the critical threshold of expansion and rupture after absorbing

490 water, the volume of the geotechnical bodies will expand and eventually rupture when it

491 continues to absorb water, which greatly affects the stability of the slope (Ng et al. 1998).

492 This process is further facilitated by reductions in both the effective stress and matric suction

493 within the geotechnical bodies, leading to diminished shear strength (Rahimi et al. 2011;

494 Gasmo et al. 2000), and consequently, an increase in deformation failure (Figure 12e).

495 Considering these factors, the slope is likely to experience long-term destabilization (Figures

496 12f–h). Based on both field investigations and physical modeling, this paper concludes that

497 the failure model of the Pusa collapse under the combined effects of mining and rainfall can

498 be characterized by the following sequence: bending-pulling-subsidence-creep-dumping. The

499 deformation and failure of the slope were observed to evolve through four distinct stages

500 (Figure 13):

501 (1) Mining-induced disturbances (Figure 13a)

502 Following the mining of M16, stress reallocation manifested in the roof of the mined-out

503 area. Unloading of the rock mass into the mined-out area led to the development of separation

24

© The Author(s) or their Institution(s)


Page 25 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

504 fractures. Notable variations in stress levels were observed before and after the collapse of the

505 coal seam roof. As mining in the M16 area continued, stress adjustments took place in

506 proximity to the mined-out area, giving rise to minor tension fractures. At this stage, the rock

507 mass above the roof of M14 displayed negligible deformation but experienced slight

508 perturbations.
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

509 (2) Fractures and extensions (Figure 13b)

510 During the mining of M14, the roof of the mined-out area experienced separation,

511 accompanied by the formation of multiple separation and tension fractures. Concurrently, the

512 rock mass at the top of the slope began to incline and deform towards the free face, resulting

513 in tension fractures at the trailing edge of the slope. The mining of M10 further intensified

514 slope deformation and failure, leading to the formation of a localized bubble area and a

515 fracture development area in the upper part. Additionally, the fracture at the trailing edge

516 extended downward into the mined-out area.

517 (3) Formation of a potential collapse surface (Figure 13c)

518 Subsequent to rainfall, water infiltrated the interior of the slope, gradually transitioning

519 the geotechnical bodies from an unsaturated to a saturated state. This infiltration caused an

520 increase in pore water pressure, reducing both the effective stress and matric suction within

521 the geotechnical bodies, thereby lowering the shear strength (Rahimi et al. 2011; Gasmo et al.

522 2000; Ng et al. 2003; Ng et al. 1998). The slope then began a creeping phase, continuing to

523 deform toward the free face. Ongoing rainfall facilitated the upward extension of fractures

524 from the mined-out area, eventually intersecting with the fractures at the slope's top and

525 forming a potential collapse surface.

25

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 26 of 55

526 (4) Destabilization failure (Figure 13d)

527 Due to persistent rainfall and the growing collapse area, fractures widened and

528 lengthened, further fragmenting the geotechnical bodies. These bodies, initially in an

529 unsaturated state, increasingly saturated due to rainwater infiltration. This led to elevated pore

530 water pressure, diminished shear strength, and volumetric expansion, ultimately causing a
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

531 sudden shearing off of the slope's “locking section." This abrupt fracturing of the rock mass

532 released substantial energy, which swiftly transformed into mechanical energy, propelling the

533 slope to slide rapidly in the direction of the debris flow. The slide was arrested by a smaller

534 hill at the base of the slope, causing the debris field to diverge in two directions.

535 6. Discussion

536 How does underground mining and rainfall cause slope deformation and ultimately

537 collapse? Understanding the interplay between underground mining, rainfall, and slope

538 deformation and failure is crucial for disaster prevention and mitigation, particularly in

539 Southwest China. In this study, we used the Pusa collapse as an example to examine this

540 complex relationship. Post-mining of M16, the roof plate underwent stress redistribution and

541 fractures emerged in the vicinity of the mined-out area, although no substantial deformation

542 was observed. Upon mining M14, the top plate separated, leading to multiple tension and

543 separation fractures as the slope began to deform toward the free face. The mining of M10

544 further accelerated slope deformation, introducing new fractures that extended from the

545 trailing edge into the mined-out area. These fractures serve as conduits for rainwater

546 infiltration. During rain events, water seeps into the interior of slopes through fractures and
26

© The Author(s) or their Institution(s)


Page 27 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

547 geotechnical bodies. As these layers absorb water, they become more saturated, resulting in

548 decreased effective stress and matric suction and a subsequent reduction in shear strength.

549 These findings align with the physical modeling tests conducted by Yang et al. (2022c) and

550 Zhang et al. (2023b).

551 However, it is important to acknowledge that a physical model cannot fully replicate the
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

552 conditions of a real-world scenario. Due to the larger bonding force of the model material, the

553 model strength and resistance to external loads, the model is not easy to be deformed during

554 the test, resulting in less obvious fracture evolution, so the fracture evolution law is still to be

555 discussed. The primary variables considered in this study were density and compressive

556 strength, and future research could include a broader range of physical parameters such as

557 angle of internal friction, modulus of elasticity and tensile strength, while similar metrics

558 could consider deformation time and deformation velocity. Due to site constraints, the size of

559 the physical model is small, so the stress behavior in the model can only maintain a similar

560 relationship with the actual prototype, and in the future, larger size physical model tests and

561 simulations using centrifuges can be carried out to obtain more accurate deformation and

562 damage laws. The selection of rainfall intensity for the model tests was based solely on data

563 from the month preceding the actual collapse; future studies could explore the impact of

564 rainfall over extended periods. Due to the limited equipment conditions, only the moisture

565 content of the model was monitored during the test without monitoring the pore water

566 pressure, which affects the effective stress and shear strength of the geotechnical body during

567 rainfall infiltration, and plays a crucial role in the stability of the slope, and the influence of

568 pore water pressure should be taken into account for further research in this area in the future.

27

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 28 of 55

569 7. Conclusions

570 In this study, we elucidated the failure mechanisms and processes leading to the

571 destabilization and collapse of the Pusa mining layer, drawing upon both a physical model
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

572 and test data. We examined the stress patterns, fracture development, water content, and

573 displacement characteristics of the slope under the combined influences of mining and

574 rainfall. The deformation failure behavior obtained from the physical model established by

575 the above parameters is consistent with the actual prototype, which illustrates the accuracy of

576 the method. The following key conclusions were reached:

577 1. During the mining of M16, a notable stress alteration occurred in the roof of the

578 mined-out area. As mining activities extended to M14 and M10, stress changes in the model

579 became increasingly pronounced. Each subsequent rainfall test further accentuated the stress

580 responses, displaying a "sudden rise and sudden fall" pattern.

581 2. Mining activities disrupted the original stress equilibrium in the rock mass, resulting

582 in the development of separation and tension fractures. Subsequent rainfall events exacerbated

583 the expansion and extension of these fractures. Both mining and rainfall contributed to the

584 incremental increase in displacement values and range, with the maximum displacement

585 being observed in the roof of the mined-out area.

586 3. Rainwater permeated the slope from the fractures and geotechnical bodies, more

587 profoundly affecting the upper rock mass than the lower sections. This led to an increase in

588 saturation, a decrease in matric suction, a rise in pore water pressure, and ultimately a

589 reduction in shear strength. When the pore water pressure reached a certain threshold, the
28

© The Author(s) or their Institution(s)


Page 29 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

590 slope shifted toward the free face, resulting in creeping and sliding deformations.

591 4. The failure mode of the Pusa slope under the influence of mining and rainfall can be

592 described as bending-pulling-subside-creep-dumping. The findings from this physical model

593 outline a four-stage process of slope failure: mining-induced disturbances, fracture initiation

594 and extension, formation of a potential collapse surface, and destabilization failure.
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

595 Data availability statement

596 Data generated or analyzed during this study are available from the corresponding author

597 upon reasonable request.

598 Author contributions

599 Conceptualization: TWT, WBS, SZX

600 Data curation: TWT, SZX

601 Formal Analysis: TWT, WBS, SZX, FL, JYZ

602 Funding acquisition: WBS

603 Investigation: TWT, WBS, SZX, FL, JYZ

604 Methodology: TWT, WBS, SZX, FL, JYZ

605 Project administration: TWT, WBS, SZX, FL

606 Resources: TWT, WBS, SZX, FL, JYZ

607 Software: TWT

29

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 30 of 55

608 Supervision: TWT, WBS

609 Validation: TWT, WBS

610 Visualization: TWT, WBS, SZX

611 Writing – original draft: TWT


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

612 Writing – review & editing: TWT, WBS, SZX

613 Competing interests

614 The authors declare there are no competing interests.

615 Funding statement

616 This study is financially supported by the National Natural Science Foundation of China (Gra

617 nt No. 42067046), Science and Technology Planning Project of Guizhou Province, China (Gr

618 ant No. QKHJC-ZK[2021]YB228), and Startup Research Foundation for High-Level Talents

619 of Guizhou University (Grant No. 2017077)

30

© The Author(s) or their Institution(s)


Page 31 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

620 References

621 Chen, L.Q., Zhao, C.Y., Li, B., He, K., Ren, C.F., Liu, X.J., and Liu, D.L. 2021. Deformation

622 monitoring and failure mode research of mining-induced Jianshanying landslide in karst

623 mountain area, China with ALOS/PALSAR-2 images. Landslides 18:2739-2750.


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

624 doi:10.1007/s10346-021-01678-6.

625 Cui, F.P., Li, B., Xiong, C., Yang, Z.P., Peng, J.Q., Li, J.S., and Li, H.W. 2022. Dynamic

626 triggering mechanism of the Pusa mining-induced landslide in Nayong County,

627 Guizhou Province, China. Geomatics, Natural Hazards and Risk 13:123-147.

628 doi:10.1080/19475705.2021.2017020.

629 Dai, Z.Y., Zhang, L., Wang, Y.L., Jiang, Z.B., and Xu, S.Q. 2021. Deformation and failure

630 response characteristics and stability analysis of bedding rock slope after underground

631 adverse slope mining. Bulletin of Engineering Geology and the

632 Environment 80:4405-4422. doi:10.1007/s10064-021-02258-7.

633 Dong, M.L., Zhang, F.M., Lv, J.Q., Hu, M.J., and Li, Z.N. 2020. Study on deformation and

634 failure law of soft-hard rock interbedding toppling slope base on similar test. Bulletin of

635 Engineering Geology and the Environment 79:4625-4637.

636 doi:10.1007/s10064-020-01845-4.

637 Fan, X.M., Xu, Q., Scaringi, G., Zheng, G., Huang, R.Q., Dai, L.X., and Ju, Y.Z. 2019. The

638 “long” runout rock avalanche in Pusa, China, on August 28, 2017: a preliminary report.

639 Landslides 16:139-154. doi:10.1007/s10346-018-1084-z.

640 Feng, Z., Li, B., Cai, Q.P., and Wen, C.J. 2016. Initiation mechanism of the Jiweishan

641 landslide in Chongqing, Southwestern China. Environmental and Engineering

31

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 32 of 55

642 Geoscience 22:341-351. doi:10.2113/gseegeosci.22.4.341.

643 Fernandez, P.R., Granda, G.R., Krzemień, A., Cortes, S.G., and Valverde, G.F. 2020.

644 Subsidence versus natural landslides when dealing with property failure liabilities in

645 underground coal mines. International Journal of Rock Mechanics and Mining

646 126:104175. doi:10. 1016/j. ijrmms.2019. 104175.


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

647 Giovanna, C., Damiano, E., Greco, R., Olivares, L., and Spolverino, G. 2020. Physical

648 modeling investigation of rainfall infiltration in steep layered volcanoclastic slopes.

649 Journal of Hydrology 580, 0022-1694. doi:10.1016/j.jhydrol.2019.124199.

650 Gasmo, J.M., Rahardjo, H., and Leong, E.C. 2000. Infiltration effects on stability of a residual

651 soil slope. Computers and Geotechnics 26(2), 145-165.

652 doi:10.1016/S0266-352X(99)00035-X.

653 Hu, X.L., Zhang, M., Sun, M.J., Huang, K.X., and Song, Y.J. 2013. Deformation

654 characteristics and failure mode of the Zhujiadian landslide in the Three Gorges

655 Reservoir, China. Bulletin of Engineering Geology and the Environment 74:1-12.

656 doi:10.1007/s10064-013-0552-x.

657 Kwan, J.S.H., Sze, E.H.Y., and Lam, C. 2018. Finite element analysis for rockfall and debris

658 flow mitigation works. Canadian Geotechnical

659 Journal 56:1225-1250. doi:10.1139/cgj-2017-0628.

660 Jones, D.B., Siddle, H.J., Reddish, D.J., and Whittaker, B.N. 1991. Landslides and

661 undermining: slope stability interaction with mining subsidence behaviour. Proc 7th

662 ISRM Int Cong Rock Mechan Aachen 893–898. doi:10. 1016/0148- 9062(93) 93140-S.

663 Li, B., Feng, Z., Wang, G.Z., and Wang, W.P. 2016a. Processes and behaviors of block topple

32

© The Author(s) or their Institution(s)


Page 33 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

664 avalanches resulting from carbonate slope failures due to underground mining.

665 Environmental Earth Sciences 75:1435-1441. doi:10.1007/s12665-016-5529-1.

666 Li, Z., He, X., Dou, L., and Song, D. 2018. Comparison of rockburst occurrence during

667 extraction of thick coal seams using top-coal caving versus slicing mining

668 methods. Canadian Geotechnical Journal 55:1433-1450. doi:10.1139/cgj-2016-0631.


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

669 Li, Y.H., Zhang, Q., Lin, Z.B., and Wang, X.D. 2016b. Spatiotemporal evolution rule of rocks

670 fracture surrounding gob-side roadway with model experiments. International Journal

671 of Mining Science and Technology. 26(5):895–902. doi:10.1016/j.ijmst.2016.05.031.

672 Li, Y.H., Tang, X.J., Yang, S., and Chen, J.W. 2019. Evolution of the broken rock zone in the

673 mixed ground tunnel based on the DSCM. Tunnelling and Underground Space

674 Technology. 84:248–258. doi:10. 1016/j. tust. 2018. 11. 017.

675 Liu, Z.Y., Mei, G., and Sun, Y.J. 2022. Investigating deformation patterns of a

676 mining-induced landslide using multisource remote sensing: the Songmugou landslide

677 in Shanxi Province, China. Bulletin of Engineering Geology and the Environment

678 81:226. doi:10.1007/s10064-022-02699-8.

679 Locat, A., Locat, P., Demers, D., Leroueil, S., Robitaille, D., and Lefebvre, G. 2017. The

680 Saint-Jude landslide of 10 May 2010, Quebec, Canada: Investigation and

681 characterization of the landslide and its failure mechanism. Canadian Geotechnical

682 Journal 54:1357-1374. doi:10.1139/cgj-2017-0085.

683 Lai, Q., Zhao, J., Shi, B. et al. 2023. Deformation evolution of landslides induced by coal

684 mining in mountainous areas: case study of the Madaling landslide, Guizhou,

685 China. Landslides 20, 2003-2016 (2023). doi:org/10.1007/s10346-023-02069-9

33

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 34 of 55

686 Ma, G.T., Hu, X.W., Yin, Y.P., Luo, G., and Pan, Y.X. 2018. Failure mechanisms and

687 development of catastrophic rockslides triggered by precipitation and open-pit mining

688 in Emei, Sichuan, China. Landslides 15:1401-1414. doi:10.1007/s10346-018-0981-5.

689 Ma, Y.T. 2017. Physical simulation study on genetic mechanism of mining landslide under

690 rainfall condition-take Madaling landslide in Guizhou province as an example.


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

691 Chengdu University of Technology. (in Chinese).

692 Ng, C.W.W., Zhan, L.T., Bao, C.G., Fredlund, D.G., and Gong, B.W. 2003. Performance of

693 an unsaturated expansive soil slope subjected to artificial rainfall infiltration.

694 Géotechnique 53(2), 143-157. doi:10.1680/geot.2003.53.2.143.

695 Ng, C.W.W., and Shi, Q. 1998. A numerical investigation of the stability of unsaturated soil

696 slopes subjected to transient seepage. Computers and Geotechnics 22(1), 1-28.

697 doi:10.1016/S0266-352X(97)00036-0

698 Rahimi, A., Rahimi, H., and Leong, E.C. 2011. Effect of Antecedent Rainfall Patterns on

699 Rainfall-Induced Slope Failure. Journal of Geotechnical and Geoenvironmental

700 Engineering 137(5), 483-491. doi: 10.1061/(ASCE)GT.1943-5606.0000451

701 Kennedy, R., Take, W.A., and Siemens, G. 2020. Geotechnical centrifuge modelling of

702 retrogressive sensitive clay landslides. Canadian Geotechnical Journal. 58(10),

703 1452-1465. doi: 10.1139/cgj-2019-0677

704 Mitchell, R.J. 2011. Centrifuge modelling as a consulting tool. Canadian Geotechnical

705 Journal. 28(1), 162-167. doi:org/10.1139/t91-018

706 Pei, H.F., Zhang, S.Q., Borana, L., Zhao, Y, and Yin, J.H. 2018. Slope stability analysis based

707 on real-time displacement measurements. Measurement 131:686-693.

34

© The Author(s) or their Institution(s)


Page 35 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

708 doi.org/10.1016/j.measurement.2018.09.019.

709 Stephen, F.G., Cortland, F.E., Douglas, C.P., and Alexander, R. 2006. Coal and the

710 environment. American Geological Institute, Alexandria.

711 Sun, S.W., Pang, B., Hu, J.B., Yang, Z.X., and Zhong, X.Y. 2021. Characteristics and

712 mechanism of a landslide at Anqian iron mine, China. Landslides 18:2593-2607.


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

713 doi:10.1007/s10346-021-01671-z.

714 Sun, S.W., Liu, L., Hu, J.B., and Ding, H. 2022. Failure characteristics and mechanism of a

715 rain-triggered landslide in the northern longwall of Fushun west open pit,

716 China. Landslides 19:2439-2458. doi:10.1007/s10346-022-01926-3.

717 Tao, T., Shi, W., Liang, F., and Wang, X. 2022. Failure mechanism and evolution of the

718 Jinhaihu landslide in Bijie City, China, on January 3, 2022. Landslides 19:2727-2736.

719 doi:10.1007/s10346-022-01957-w.

720 Tao, Z.G., Shu, Y., Yang, X.J., Peng, Y.Y., Chen, Q.H., and Zhang, H.J. 2020. Physical

721 model test study on shear strength characteristics of slope sliding surface in Nanfen

722 open-pit mine. International Journal of Mining Science and Technology 30:421-429.

723 doi:10.1016/j.ijmst.2020.05.006.

724 Tang, J., Dai, Z., Wang, Y., and Zhang, L. 2019. Fracture failure of consequent bedding rock

725 slopes after underground mining in mountainous area. Rock Mech Rock Eng

726 52(8):2853–2870. doi:10. 1007/ s00603- 019- 01876-8

727 Wang, X., Xiao, Y., Shi, W., Ren, J., Liang, F., Lu, J., Li, H., and Yu, X. 2022a. Forensic

728 analysis and numerical simulation of a catastrophic landslide of dissolved and fractured

729 rock slope subject to underground mining. Landslides 19:1045-1067.

35

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 36 of 55

730 doi:10.1007/s10346-021-01842-y.

731 Wang, Y.D., Wang, X., Zhang, J.S., Yang, B.S., Zhu, J.W., and Wang, Z.P. 2022b. Similar

732 experimental study on retaining waterproof coal pillar in composite strata

733 mining. Scientific Reports 12:1366. doi:10.1038/s41598-022-05369-7.

734 Xing, A., Wang, G., Li, B., Jiang, Y., Feng, Z., and Kamai, T. 2014. Long-runout mechanism
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

735 and landsliding behaviour of large catastrophic landslide triggered by heavy rainfall in

736 Guanling, Guizhou, China. Canadian Geotechnical Journal. 52(7):

737 971-981. doi:10.1139/cgj-2014-0122.

738 Xu, Q., Liu, H., Ran, J., Li, W., and Sun, X. 2016. Field monitoring of groundwater responses

739 to heavy rainfalls and the early warning of the Kualiangzi landslide in Sichuan Basin,

740 southwestern China. Landslides 13:1555-1570. doi:10.1007/s10346-016-0717-3.

741 Xu, Q., Shang, Y.J., van Asch, Th.W.J., Wang, S.J, Zhang, Z.Y, and Dong, X.J. 2012.

742 Observations from the large, rapid Yigong rock slide–debris avalanche, Southeast

743 Tibet. Canadian Geotechnical Journal 49:589-606. doi:10.1139/t2012-021.

744 Yang, D.D., Qiu, H.J., Ma, S.Y., Liu, Z.J., Du, C., Zhu, Y.R., and Cao, M.M. 2022a. Slow

745 surface subsidence and its impact on shallow loess landslides in a coal mining area.

746 CATENA 209:105830. doi:10.1016/j.catena.2021.105830.

747 Yang, C.W., Shi, W.B., Peng, X.W., Zhang, S.B., and Wang, X.M. 2022b. Numerical

748 simulation of layered anti-inclined mining slopes based on different free face

749 characteristics. Bulletin of Engineering Geology and the Environment 81:359.

750 doi:10.1007/s10064-022-02855-0.

751 Yang, Z.P., Zhao, Q., Liu, X.R., Yin, Z.M., Zhao, Y.L., and Li, X.Y. 2022c. Experimental

36

© The Author(s) or their Institution(s)


Page 37 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

752 study on the movement and failure characteristics of karst mountain with deep and large

753 fissures induced by coal seam mining. Rock Mechanics and Rock Engineering

754 55:4839-4867. doi:10.1007/s00603-022-02910-y.

755 Yu, X.Y., and Mao, X.W. 2020. A preliminary discrimination model of a deep mining

756 landslide and its application in the Guanwen coal mine. Bulletin of Engineering
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

757 Geology and the Environment 79:485-493. doi:10.1007/s10064-019-01565-4.

758 Yin, Y., Xing, A., Wang, G., Feng, Z., Li, B., and Jiang, Y. 2017. Experimental and

759 numerical investigations of a catastrophic long-runout landslide in Zhenxiong, Yunnan,

760 southwestern China. Landslides 14, 649–659. doi:10.1007/s10346-016-0729-z.

761 Zhao, J.J., Jian, J., Ma, Y.T., and Zhang, X.D. 2016. Physical modeling on deformation

762 characteristics of overlying rock mass above mined-out area in gently inclined coal

763 seam. Journal of China Coal Society 41:1369-1374. (in Chinese). doi:10.13225

764 /j.cnki.jccs.2015.1408.

765 Zhang, C., Li, T.F., and Han, X.D. 2020. Slope failure mechanism affected by mining

766 subsidence: a case study of highway slopes in Yangquan, Shanxi Province, China. IOP

767 Conference Series: Earth and Environmental Science 570:022067.

768 doi:10.1088/1755-1315/570/2/022067.

769 Zhang, S., Shi, W., Wang, Y., Liang, F., Zhang, J., and Wang, X. 2023a. Numerical

770 evaluation of the deformation and failure mechanisms and movement processes of the

771 Guanling landslide in Guizhou, China. Landslides 20, 1747–1762.

772 doi:10.1007/s10346-023-02059-x.

773 Zhang, S., Shi, W., Yang, C., Wang, Y., and Yu, X. 2023b. Experimental evaluation of gentle

37

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 38 of 55

774 anti-dip slope deformation and fracture network under the action of underground

775 mining. Landslides 20, 381–408 (2023). doi:10.1007/s10346-022-01976-7.

776 Zhang, Q.Y., Li, S.C., and Jiao, Y.Y. 2005. Numerical method of rock mass analysis and

777 principle of geomechanical model test and engineering application. Beijing: China

778 Water Resources and Hydropower Press. (in Chinese).


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

779 Zhang, Z.G., Lv, X.L., Mao, M.D., Pan, Y.T., Fang, L., and Wu, Z.T. 2022. Mechanical

780 response for rainfall-induced landslides on jointed gas pipelines. Computers and

781 Geotechnics 146:0266-352X. doi:10.1016/j.compgeo.2022.104708.

782 Zheng, G., Xu, Q., Ju, Y.Z., Li, W.L., Zhou, X.P., and Peng, S.Q. 2018. The Pusacun

783 rockavalanche on August 28, 2017 in Zhangjiawan Nayongxian, Guizhou:

784 characteristics and failure mechanism. Journal of Engineering Geology 26:223-240. (in

785 Chinese). doi:10.13544 /j.cnki.jeg.2018.01.023.

786 Zhou, H.X., Che, A.L., Chen, J.C., and Yuan, G.L. 2023. Study on structural damage

787 evolution of excavated slope subjected to earthquake and rainfall using electrical

788 resistivity measurement. Soil Dynamics and Earthquake Engineering 170:0267-7261.

789 doi.org/10.1016/j.soildyn.2023.107908

38

© The Author(s) or their Institution(s)


Page 39 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

Figure captions

Figure 1 Images of areas under study: (a) drone aerial photography: slide direction is 305°;

Ⅰ, Ⅱ, Ⅲ are source, scraping, and accumulation areas, respectively; (b) discontinuities in

collapse area; (c) dissolution fissures; (d) close-up view of mining-induced fracture; (e) karst

collapse.
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

Figure 2 Map of area under study: 1, First member of Yelang Formation of Lower Triassic; 2,

Second member of Yelang Formation of Lower Triassic; 3, Changxing–Dalong Formation of

Upper Permian; 4, Upper Permian Longtan Formation; 5, Quaternary deposits; 6, fault; 7,

coal seam; 8, outline of Pusa collapse; 9, section line; 10, M14 mined-out area; 11, M10

mined-out area; 12, M16 mined-out area.

Figure 3 Profile of Pusa collapse along principal sliding direction (A-A' in Figure 2): 1,

limestone; 2, muddy siltstone; 3, mudstone; 4, angle; 5, coal seam; 6, mined-out area; 7, fault;

8, karst fissures; 9, second member of the Yelang Formation of the Lower Triassic; 10, First

member of the Yelang Formation of Lower Triassic; 11, Changxing–Dalong Formation of

Upper Permian; 12, Upper Permian Longtan Formation; 13, Original topography; 14, Present

topography; 15, Accumulation body

Figure 4 Rainfall (mm) during the month preceding collapse.

Figure 5 Schematic of the physical model. 1, stress monitoring point; 2, water content

monitoring point; 3, limestone; 4, argillaceous siltstone; 5, mudstone; 6, mined-out area; 7,

coal seam.

Figure 6 Experimental setup. (a) Distribution of monitoring points: T, stress; S, water

content. (b-g) Experimental equipment: (b) rainfall simulation. (c) water sensor; (d) stress

© The Author(s) or their Institution(s)


For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 40 of 55

sensor; (e) model box; (f) data acquisition equipment; (g) joints made of mica sheets.

Figure 7 Principle of punctuation-free method correlation analysis

Figure 8 Variation law of rock stress during mining.

Figure 9 Fracture evolution under mining and rainfall (D, separation fracture; T, tensile

fracture). (a) excavation of simulated coal seam M16; (b) excavation of M14; (c) excavation
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

of M10; (d) rainfall

Figure 10 Change in water content during three periods of simulated rainfall.

Figure 11 Process of total displacement during the Pusa collapse (the red rectangle represents

truss used in model building process; the area was not displaced during the test.).

Figure 12 The model-predicted initiation and evolution of slope collapse. The red lines in

Fig. 11a indicate the rock boundary, and the blue lines indicate the joint. The red lines in Fig.

11c and d indicate layer separation fracture, and the blue lines indicate tensile fracture.

Figure 13 Slope deformation and failure process: (a) mining disturbance; (b) fracture

extension and extension; (c) formation of a potential collapse surface; (d) destabilization

failure.

© The Author(s) or their Institution(s)


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 41 of 55

Figures
Figure 1

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Figure 2

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Page 42 of 55
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 43 of 55

Figure 3

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Figure 4

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Page 44 of 55
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 45 of 55

Figure 5

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Figure 6

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Page 46 of 55
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 47 of 55

Figure 7

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Figure 8

10

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Page 48 of 55
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 49 of 55

Figure 9

11

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Figure 10

12

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Page 50 of 55
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 51 of 55

Figure 11

13

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Figure 12

14

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
Page 52 of 55
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 53 of 55

Figure 13

15

© The Author(s) or their Institution(s)


Canadian Geotechnical Journal (Author Accepted Manuscript)
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript) Page 54 of 55

Tables

Table 1 Characteristics of main coal seams at the Pusa site

Coal seam Thickness (m) Dip angle (°) Top bed Bottom bed

M16 0.76–2.04 10 silty mudstone mudstone

M14 0.77–1.95 10 silty mudstone mudstone


Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

M10 0.79–2.95 10 silty mudstone silty mudstone

© The Author(s) or their Institution(s)


Page 55 of 55
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Canadian Geotechnical Journal (Author Accepted Manuscript)

Table 2 Mechanic parameters of lithology and similar materials

Quartz
Compressive
Density sand: Gypsum: Barite
Lithology strength Note
(g/cm3) Cementing Cement powder/Aggregate
(KPa)
material
Can. Geotech. J. Downloaded from cdnsciencepub.com by NAT INST OF TECHNOLOGY on 03/30/24

Limestone 2.71 83.6×103 / / /

Argillaceous
2.21 47.7×103 / / / Actual
siltstone
samples
Mudstone 1.83 29.8×103 / / /

Coal seams 1.62 22.7×103 / / /

Material 1 2.69 232 14:1 5:1 0.5

Material 2 2.14 123 5:1 1:0 0.5 Similar

Material 3 1.75 73 3:2 1:0 0.1 samples

Material 4 1.57 62 12:1 10:1 0.1

Materials 1-4 of the model were composed such that they simulated limestone, argillaceous

siltstone, mudstone, and coal, respectively. Cγ is 1.007 and Cσ is 360 for material 1; Cγ is

1.033 and Cσ is 387 for material 2; Cγ is 1.058 and Cσ is 408 for material 3; Cγ is 1.032 and Cσ

is 366 for material 4.

© The Author(s) or their Institution(s)

You might also like