University
stringclasses
19 values
Text
stringlengths
458
20.7k
ADE
can provide more practical outputs. The successful application of these algorithms has been reported by other researchers in mining and geotechnical engineering fields (Armaghani et al. 2016; Salimi et al. 2016; Hasanipanah et al. 2017b; Khandelwal et al. 2017). Hence, it is necessary to use state-of-the-art modelling techniques to address the mentioned difficulties and develop new models for predicting rockburst maximum stress and its risk index based on field conditions. As it has been summarized in Fig. 5.4, this study focuses on the following steps: 1) compiling a database based on the true-triaxial unloading tests on different rock types and performing a broad statistical analysis on it to create a homogeneous database and to select the most influential parameters based on an appropriate strategy; 2) Developing genetic-based and decision tree-based models for the prediction of maximum rockburst stress (𝜎 ) and rockburst 𝑅𝐡 risk index (𝐼 ) based on the selected input parameters; 3) validation verification of the 𝑅𝐡 developed models; and 4) conducting a parametric analysis to assess the effect of input parameters on the corresponding outputs. Figure 5.1 Rock ejection and deformation of the supporting system due to strainbursting (Feng et al. 2017) 117
ADE
horizontal in-situ stress in the face to be unloaded; 𝜎 : vertical in-situ stress; 𝜎 : rockburst 𝑣 𝑅𝐡 maximum stress; 𝐼 : rockburst risk index; 𝐷: depth, 𝜌: density, 𝐾: horizontal pressure 𝑅𝐡 coefficient (ratio of average horizontal stresses to the vertical stress due to overburden), 𝑀𝐿𝑅: multiple linear regression; 𝑉𝐼𝐹: variance inflation factor; 𝑅2: coefficient of determination) 5.2. Data Collection and Statistical Analysis In this study, a database containing information about the 139 rockburst laboratory tests conducted on different rock types from 2004 to 2012 at the State Key Laboratory for Geomechanics and Deep Underground Engineering (SKLGDUE), China was compiled. The tested rock samples were gathered from the depth of 200 m to 3375 m. This database consists of many parameters such as rock mechanical properties, in-situ stresses, rock sample depth, rockburst critical depth, rock density, rock specific weight, mineral contents of rocks, loading and unloading rates of the true-triaxial tests, rockburst maximum stress, rockburst risk index, test duration and bursting mechanism. Considering a circular shape for the tunnel crown, the stress concentration factor equal to 2, and the specific weight of 27 kN/m3 for the overburden rock mass, the rockburst critical depth (𝐻 ) was calculated by the following equation: 𝑒 H = 18.52𝜎 (5.1) e 𝑅𝐡 The rockburst risk index (I ) also was calculated for all the samples through the following RB equation (He 2009): H H I = = 0.054 (5.2) RB He ΟƒRB He (2009) defined a new classification for 𝐼 as shown in Table 5.1. Based on this 𝑅𝐡 classification, a 56% of the tested samples have low 𝐼 , 13% of the samples have very high 𝑅𝐡 𝐼 , and the remained 31% of samples have moderate to high 𝐼 . Since all the foregoing 𝑅𝐡 𝑅𝐡 parameters have not been collected during the rockburst tests, there are some missing values in the database. To have a homogeneous database, the missing values (30 records) were eliminated from the primary database, and finally, the results of 109 tests were considered for further analyses. Before developing any model, the presence of natural groups and outliers in the raw database was evaluated using agglomerative hierarchical clustering (AHC) analysis. In fact, the presence of outliers and natural groups can decrease the generality and liability of the developed models (Hudaverdi 2012; Faradonbeh and Monjezi 2017; Shirani Faradonbeh and 120
ADE
Taheri 2019) The AHC is the most common type of clustering techniques which is used in earth sciences (Hudaverdi 2012). The AHC follows a bottom-up procedure that iteratively creates the single object clusters and then these clusters are merged into the larger clusters based on the similarity or dissimilarity criteria. The common criterion for clustering is β€œdistance”, and this means that objects in the same cluster have the least distance from each other, while objects in different clusters are at a great distance from one another. The process of cluster generating and merging is continued until all the objects (datasets) are placed in a single cluster or the pre-defined termination condition is satisfied. For measuring the distance between the objects, the average-linkage function that measures the average distance of any object of one cluster from an object of the other cluster was used to form the clusters (Kaufman and Rousseeuw 2009; Saxena et al. 2017): 1 βˆ‘ βˆ‘ 𝑑(π‘Ž,𝑏) (5.3) π‘Žβˆˆπ΄ π‘βˆˆπ΅ |𝐴||𝐡| where 𝐴 and 𝐡 are two clusters with the sizes of |𝐴| and |𝐡|, respectively. π‘Ž and 𝑏 are objects from the mentioned clusters and 𝑑 is the squared Euclidean distance between two objects. Table 5.1 Rockburst risk index classification, He et al. (2015) Rockburst risk index (𝐼 ) Class 𝑅𝐡 𝐼 <0.6 Low 𝑅𝐡 0.6<𝐼 ≀1.2 Moderate 𝑅𝐡 1.2<𝐼 ≀2.0 High 𝑅𝐡 𝐼 β‰₯2.0 Very high 𝑅𝐡 Fig. 5.5 shows the dendrogram derived from the conducted clustering analysis by AHC. A dendrogram is a tool that represents the relative size of the calculated distances at which the objects and clusters are combined. The objects with the low squared Euclidean distance (high similarity) are close together and vice versa. The X-axis shows the dataset number and the Y- axis shows the rescaled value of the distance. To prevent Fig. 5.5 to be crowded and large, the numbers of the datasets have been summarised on the X-axis. Clearly can be seen from Fig. 5.5 that the whole 109 collected datasets were clustered into one distinct group between the rescaled distances of 0 and 5 except for two cases of 75 and 76 which were placed in the second group. By checking the database, it was found out that the main parameter that caused to grouping is depth, and the members of group 2 belong to the depth of 3375 m which are known as outliers for the current database. Therefore, these two cases were removed from the database to avoid the influence of their distinctive behaviour on the modelling process, and the 121
ADE
5.3. Methods and Results 5.3.1. Stepwise Selection and Elimination Process This section aims to do a systematic stepwise selection and elimination (SSE) analysis to identify the most important parameters on the outputs and reduce the complexity of the developed models. The process of parameter reduction also is carried out using the variable pressure tools of the robust data-mining techniques i.e. GEP and CART. There are several critical statistical terms which have been used in this study for the primary assessment of the database and are defined in the following. Multicollinearity, a high correlation between the independent (predictor) variables, can be considered as one of the most prominent challenges for multiple regressions. The existence of this phenomenon may lead to developing an unstable regression model having high values for variance and covariance coefficients (Sayadi et al. 2012). Variance inflation factor (VIF) is a statistical index to quantify the extent of the multicollinearity between the independent (input) parameters. This index is the ratio of model variance considering several inputs to the variance of the model with a single input parameter. The VIF lower than 10 shows the non-existence of multicollinearity (James et al. 2013). Another important index is Sig. (2-tailed) or p-value of the correlations. The Sig (2-tailed) represents the significance of the correlation at a prescribed alpha level (5%). The Sig. (2- tailed) should be less than or equal to 0.05 to reject the influence of chance factor. The coefficient of determination (denoted by 𝑅2) is another statistical measure for evaluation of the model performance. This index interprets the proportion of the output (dependent) variable’s variance that is predictable from the input (independent) variables. An 𝑅2 of 1 indicates that the regression predictions perfectly fit the data (Montgomery et al. 2012; James et al. 2013; Kumar Sharma and Rai 2017). In the current study, uniaxial compressive strength (π‘ˆπΆπ‘†), Young’s modulus (𝐸), Poisson’s ratio (𝜈), horizontal in-situ stress (𝜎 ), horizontal in-situ stress in the face to be unloaded (𝜎 ), β„Ž1 β„Ž2 vertical in-situ stress (𝜎 ), depth (𝐷), density (𝜌), and horizontal pressure coefficient (𝐾) are 𝑣 known as the input parameters for the maximum rockburst stress (𝜎 ), while all the mentioned 𝑅𝐡 parameters are considered as inputs for the rockburst risk index (𝐼 ). The SPSS software 𝑅𝐡 package 25.0 was used for performing the statistical evaluations. Initially, the database was fed to the software, and the Person’s correlation coefficient (π‘Ÿ) between the input parameters as well as between the inputs and the corresponding outputs was calculated. Table 5.2 lists the calculated correlation values. As can be seen from this table, all the inputs significantly 125
ADE
correlating with 𝜎 (i.e. 𝑆𝑖𝑔.(2βˆ’π‘‘π‘Žπ‘–π‘™π‘’π‘‘) ≀ 0.05), while 𝐷 (depth) with the 𝑆𝑖𝑔.> 0.05 𝑅𝐡 and low correlation coefficient (π‘Ÿ = βˆ’0.128) was removed from the input parameters for further modelling of 𝐼 . The elimination of parameters does not show that they have not any 𝑅𝐡 influence on the output, but simply it means that the effect of those parameters will be minimum in predicting the output. As an initial multicollinearity assessment between input parameters, no one of the correlations exceeds from the condition of π‘Ÿ > 0.90 (Hemmateenejad and Yazdani 2009). However, these input parameters may show multicollinearity when a combination of them are used as regressors in MLR. Based on the above analysis, all the inputs (except parameter 𝐷 for 𝐼 ) were retained for multiple linear regression (MLR). The MLR 𝑅𝐡 models with the possible multicollinearity were developed separately using the selected parameters for both 𝜎 and 𝐼 . 𝑅𝐡 𝑅𝐡 Table 5.3 shows the model summary, calculated coefficients, and the statistical indices for evaluating the developed MLR models. In this stage, according to Fig. 5.4, several conditions including 𝑉𝐼𝐹 < 10, 𝑆𝑑𝑑.π‘’π‘Ÿπ‘Ÿπ‘œπ‘Ÿ ≀ πΆπ‘œπ‘’π‘“π‘“.(𝐡), and πΆπ‘œπ‘’π‘“π‘“.(𝐡) β‰  0 were checked for different inputs to retain them for further evaluations. Considering Table 5.3, for rockburst maximum stress (𝜎 ), the parameters of 𝐾 and 𝜌 have 𝑉𝐼𝐹 > 10 and t-significance higher 𝑅𝐡 than 0.05, respectively, which shows that the effect of these parameters on the 𝜎 is 𝑅𝐡 insignificant. Therefore, these parameters were removed for further modelling of 𝜎 . About 𝑅𝐡 the rockburst risk index (𝐼 ), all the VIF values for inputs are less than 10, but the t- 𝑅𝐡 significance values of the π‘ˆπΆπ‘†, 𝜎 , 𝜎 , and 𝜎 are higher than 0.05. Thus, these parameters 𝑣 β„Ž1 β„Ž2 also were removed from the input set of 𝐼 . In the next step, two stepwise selection and 𝑅𝐡 elimination procedures were performed using the selected inputs for each dependent parameter. In this procedure, a parameter which is entered in the model at the initial stage of selection may be removed at the later stages. In fact, the calculations in this process are like the forward selection and backward procedure (Sarkhosh et al. 2012). Table 5.4 summarises the results of the stepwise selection and elimination procedure carried out using the algorithm provided in the SPSS 25. In this algorithm, the parameters enter the model if the probability (significance level) of its 𝐹 value is less than the Entry value (i.e. 0.05) and are eliminated if the probability is greater than the Removal value (i.e. 0.100). Entry must be less than Removal, and both values must be positive. As given in Table 5.4, during the process of selection and elimination, the correlation coefficient (𝑅) between the measured output and the predicted one was increased from 0.884 (model 1) to 0.910 (model 3) for 𝜎 𝑅𝐡 126
ADE
and from 0.754 (model 1) to 0.821 (model 5) for 𝐼 , respectively. In other words, the 𝑅𝐡 parameters of π‘ˆπΆπ‘†, 𝐸, and 𝜎 can explain 82.8% (𝑅2 = 0.828) variations in 𝜎 . As such, the 𝑣 𝑅𝐡 parameters of 𝜎 , 𝐾, 𝐸, 𝜈, and 𝜌 can explain 67.4% (𝑅2 = 0.674) variations in 𝐼 . Thus, 𝑅𝐡 𝑅𝐡 these parameters were known as the most influential ones among the initial inputs to describe the rockburst parameters. The regression coefficients and the collinearity statistics of the best SSE-based models are shown in Table 5.5. In both models, the VIF factor that shows the multicollinearity is lower than 10, the t-significance is lower than 0.05, and the Std. error values are lower than the regression coefficients which show the reliability of the SSE process in identifying the most influential parameters. Considering the above analyses, the agglomerative hierarchical clustering (AHC) accompanied by the stepwise selection and elimination (SSE) method could provide a homogeneous rockburst database by removing the outliers and decreasing the dimensionality of the problem. This process also can be useful for the complexity reduction of the next predictive models by applying a few input parameters. Due to the high non-linear and complex nature of rockburst hazard (He et al. 2015; Pu et al. 2019; Shirani Faradonbeh and Taheri 2019) there is a need to use the non-linear data-mining algorithms to provide more accurate predictive models for rockburst parameters. To do so, two robust data-driven approaches including the gene expression programming (GEP) as a meta-heuristic algorithm and the classification and regression tree (CART) as a subset of decision tree algorithms were selected for discovering the non-linear latent relationships with more accuracy and lower estimation error. These algorithms despite the various datamining and soft computing techniques such as ANNs, SVM, etc. can provide practical and easy to use outputs for the engineers and the researchers when the true-triaxial testing machine is not available. A summary of the modelling procedure by these techniques is presented in the following sections. 127
ADE
5.3.2. Non-linear Regression Analysis Non-linear regression (NLR) attempts to find a function which is a non-linear combination of the input parameters using a method of successive approximation (Archontoulis and Miguez 2015; Bethea 2018). In geoscience, most of the dependent parameters show a non-linear relationship with the related influential parameters. So, the non-linear regression analysis has been widely used by researchers in the last decades (Armaghani et al. 2016; Jahed Armaghani et al. 2017; Ghasemi 2017). The NLR technique is capable of accommodating a broad range of functions including exponential, power, logarithmic, sigmoid, logistic, trigonometric, Gaussian, etc. that boosts the process of function finding. Another advantage of the NLR is the efficient use of data, i.e. it can provide reasonable estimates of the unknown parameters for a comparatively small data. However, the common NLR technique suffers from several significant drawbacks. In NLR, there is no a closed-form and holistic mathematical structure between the dependent and the independent parameters as there is in multiple linear regression (MLR), while the choice of the model structure is a crucial task to obtain the best solution. In addition, the selection and utilizing the suitable mathematical functions from the large library of functions need an iterative optimization procedure that is not possible in common NLR modellings. Accordingly, the researchers may have to use numerical optimization algorithms to find the best-fitting parameters but still, there is a need to define the starting values for the unknown parameters in these methods. Inappropriate assigning the starting values may cause to getting caught in the local minima rather than finding the global minimum that introduces the least squares estimates (Motulsky and Ransnas 1987; Archontoulis and Miguez 2015; Bethea 2018). For these difficulties, the researchers prefer to use a non-linear regression form that has been used successfully in similar applications. Hereupon, the application of intelligent algorithms is needed to cope with these issues. In the following sections, the process of rockburst assessment using two robust non-linear techniques comprising the gene expression programming (GEP) and classification and regression tree (CART) are explained. 5.3.2.1. Rockburst Assessment Using GEP-based Models Soft computing is the relatively new branch of data-mining methods and can be considered as an alternative to the prevalent hard computing methods for solving the real-world problems (Mitchell 1997; Alavi et al. 2016). Soft computing techniques have been successfully employed in mining, rock mechanics, and geotechnical problems but despite their good performance, they cannot generate practical equations, and their structure needs to be assigned in advance by the 131
ADE
user (Alavi and Gandomi 2011). By inspiring from the Darwinian principle of β€œSurvival of the Fittest” (Nazari and Pacheco Torgal 2013) and the natural evolution, a new subset of soft computing was introduced as the evolutionary algorithm (EA). Generally speaking, EAs work with a randomly generated population of individuals which are then improved using a group of genetic operators (e.g. mutation, crossover and reproduction) and finally, the solutions are encoded into the specific forms such as binary strings in genetic algorithm. The main differences between EAs are related to the method of presenting the solutions, genetic operators, selection mechanism, and the performance measurement method (Ferreira 2002a; Alavi et al. 2016). Gene expression programming (GEP) (Ferreira 2002b) is a well-known evolutionary algorithm that inherits two essential features from its siblings i.e. the use of simple, fixed-length, and linear chromosomes with different shapes and sizes from genetic algorithm (GA) and the expression tree (ET) structure from genetic programming (GP) that improves the robustness of GEP for solving the non-linear problems (Power et al. 2019). The main entities of GEP algorithm are terminal set (input parameters and constant values), function set (e.g. +,βˆ’, Γ—, Γ·), fitness function (for evaluating the generated solutions), and genetic operators (mutation, inversion, transposition, and recombination). A flowchart detailing the GEP modelling procedure is shown in Fig. 5.7. In summary, GEP generates a population of chromosomes (solution/individual) by combining the user-defined terminals and functions. These chromosomes follow a bilingual and unequivocal expression system that is called Karva language (Ferreira 2006). The chromosomes have a specified number of genes (sub -ETs) which are linked together using a linking function (e.g. β€œ/” in Fig. 5.7 that links two genes of a chromosome). Each gene contains two parts of head and tail that the terminals (inputs) and functions (mathematical functions) are placed in them and the genetic operators are applied to these areas to modify the solutions. To have a quick understanding regarding the built-in mathematical equations of chromosomes, the Karva coded programs are then parsed into ETs. Then, the mathematical form of the programs is extracted from ETs and their fitness is evaluated by a fitness function. If the stopping condition(s) such as reaching to a specific number of iterations or the desired fitness value is not met, the selected chromosomes are replicated into a new generation, and the remained ones undergo a modification process using the genetic operators. The above process is repeated and finally, the best solution (predictive model) describing the relationship between the input and output parameters is found. More details about the mechanism of genetic operators and GEP algorithm can be found 132
ADE
in Ferreira.(Ferreira 2006) In the current study, the selected inputs from the SSE analysis were considered as terminal sets to formulate the rockburst parameters nonlinearly as follows: 𝜎 = 𝑓(π‘ˆπΆπ‘†,𝐸,𝜎 ) (5.4) 𝑅𝐡 𝑣 𝐼 = 𝑓(𝜎 ,𝐾,𝜐,𝐸,𝜌) (5.5) 𝑅𝐡 𝑅𝐡 The rockburst database was divided randomly into training and testing subsets. The training set (80 % of the database) was used to train the model and discover the relationship between inputs and outputs, and the remaining datasets were used to validate the performance of the proposed models. It should be noted that the influence of using different groups of training and testing datasets were also evaluated on the accuracy of the models. However, no noticeable change in the results was observed. For evaluating the generated solutions during the GEP modelling, it is necessary to use a fitness function. As mentioned in section 5.3.1, to propose models with lower complexity, it is possible to apply variable pressure tools to compress the developed models as much as possible by eliminating the parameters which have lower importance in a non-linear structure. To this end, the root mean squared error (RMSE) with parsimony pressure was applied to the GEP models of 𝜎 and 𝐼 (Roy et al. 2002). The 𝑅𝐡 𝑅𝐡 𝑅𝑀𝑆𝐸 of a chromosome (solution) 𝑖 is calculated by the following equation: 𝑖 𝑅𝑀𝑆𝐸 = √1 βˆ‘π‘› (𝑃 βˆ’π‘‡)2 (5.6) 𝑖 𝑛 𝑗=1 𝑖𝑗 𝑗 where 𝑃 is the predicted value by the chromosome 𝑖 for the dataset 𝑗, and 𝑇 is the measured 𝑖𝑗 𝑗 value for dataset 𝑗. The 𝑅𝑀𝑆𝐸 varies between 0 and infinity, with 0 corresponding to the ideal. Since the process 𝑖 of selection in GEP algorithm is based on the increase of fitness, Equation (6) cannot be used directly. Thus, the following expression was used for fitness function which obviously ranges between 0 to 1000, with 1000 corresponding to the ideal: 𝑅𝑀𝑆𝐸′ = 1000Γ— 1 (5.7) 𝑖 1+𝑅𝑀𝑆𝐸 𝑖 On the other hand, to apply the parsimony pressure on future models, overall fitness was defined as: 133
ADE
𝑅𝑀𝑆𝐸′′ = 𝑅𝑀𝑆𝐸′ Γ—(1+ 1 Γ— π‘†π‘šπ‘Žπ‘₯βˆ’π‘† 𝑖 ) (5.8) 𝑖 𝑖 5000 π‘†π‘šπ‘Žπ‘₯βˆ’π‘† π‘šπ‘–π‘› where 𝑆 is the size of the GEP program, 𝑆 and 𝑆 are the maximum and minimum 𝑖 π‘šπ‘Žπ‘₯ π‘šπ‘–π‘› program sizes which are calculated by the following equations: 𝑆 = 𝐺(β„Ž+𝑑) (5.9) π‘šπ‘Žπ‘₯ 𝑆 = 𝐺 (5.10) π‘šπ‘–π‘› where 𝐺 is the number of genes, and β„Ž and 𝑑 are the head size and tail size, respectively. A group of trigonometric and straightforward mathematical functions i.e. {+,βˆ’,βˆ— ,/,√,𝐿𝑛,^2,^3,^1/3,𝑠𝑖𝑛,π‘π‘œπ‘ ,π‘‘π‘Žπ‘›} were selected as the function set based on the previous non-linear studies using GEP algorithm (Kayadelen 2011; Faradonbeh and Monjezi 2017; Hoseinian et al. 2017). The other GEP parameters including the number of chromosomes, head size, the number of genes, and the values of genetic operators were changed for different runs to obtain the best solution in such a way that provides not only high accuracy but also less complexity. Table 5.6 presents the architecture of the obtained GEP models for both rockburst maximum stress (𝜎 ) and rockburst risk index (𝐼 ). By applying the parsimony pressure to 𝑅𝐡 𝑅𝐡 the models, the density parameter (𝜌) was identified intelligently as the low-impact parameter in the non-linear form of 𝐼 . Therefore, this parameter was removed by GEP automatically 𝑅𝐡 during modelling and the number of inputs for 𝐼 decreased from 5 to 4. About 𝜎 , the GEP 𝑅𝐡 𝑅𝐡 algorithm identified the three inputs of π‘ˆπΆπ‘†, 𝐸, and 𝜎 as the influential parameters for 𝑣 modelling as formerly proved by SSE analysis. The ability of GEP in identifying the low- influence parameters and excluding them during modelling can be considered as an internal sensitivity analysis that distinguishes GEP from other soft computing techniques. Fig. 5.8 displays the variations of the coefficient of determination (𝑅2) during 5000 generations (iterations) in both training and testing stages of GEP modelling for rockburst parameters. According to this figure, after a few numbers of generations (less than 1000), a rapid increase of 𝑅2 for the generated solutions can be seen which shows the high speed and high capability of GEP algorithm in function finding. From the generation 1000 to 3500, a gentle enhancement in the quality of solutions are visible, and finally, the algorithm converges into an optimum value and its value almost remains constant to reach the stopping condition (i.e. the pre-defined number of generations: 5000). The obtained 𝑅2 values for training and testing stages of 𝜎 𝑅𝐡 are 0.9266 and 0.9398, respectively, while the foregoing values are 0.8824 and 0.9459, respectively for 𝐼 . Figs. 5.9 shows the correlation of the experimentally measured values of 𝑅𝐡 134
ADE
𝜎 and 𝐼 versus the predicted ones by the constructed GEP models for training and testing 𝑅𝐡 𝑅𝐡 data groups. As seen, the data points have almost a uniform distribution around the fitted lines in both GEP-based models which show the goodness-of-fit of the models. The developed models and their performance are discussed in more details in sections 5.4 and 5.5. Eventually, the mathematical forms of the proposed GEP models for 𝜎 and 𝐼 were extracted from their 𝑅𝐡 𝑅𝐡 K-expression and ETs as Eqs. 5.11 and 5.12. To avoid the prolongation of the paper, the ETs and their K-expressions have not presented here. 𝜎 = (3 𝜎 +𝐸𝑠𝑖𝑛(𝐸 βˆ’πœŽ )+𝐸)(3 𝐸 +𝜎 +𝐸𝑠𝑖𝑛(𝐸)) 𝐿𝑛(𝐿𝑛(𝜎 )+π‘ˆπΆπ‘†) (5.11) 𝑅𝐡 𝑣 𝑣 𝑣 𝑣 𝑒6 √𝐾(𝐸+4 √𝜈) 𝐼 = (5.12) 𝑅𝐡 (𝜈+𝐾)(πΈβˆ’πœˆ) 𝐿𝑛(πœŽπ‘…π΅) Head Tail Head Tail Create initial population - b Γ— b a b b a b Γ— √ b + a b b b a Gene 1 Gene 2 / Express solutions as ETs - Γ— b Γ— √ b π‘βˆ’(π‘Γ—π‘Ž) Execute each program b a + βˆšπ‘Ž+𝑏×𝑏 a b Evaluate fitness 𝑅𝑀𝑆𝐸′′ 𝑖 1) Mutation (an element is changed to another) - b Γ— b a b b a b Before Yes - b Γ— b Γ— b b a b After TTeerrmmiinnaattee?? 2) Inversion (a fragment is inverted in the head) No Show the solution - b Γ— b a b b a b Before - Γ— b b a b b a b After Select the best solutions 3) Transposition (IS type: a fragment is copied to the head of a gene) - b Γ— b a b b a b Before - b a b a b b a b After Apply genetic operators 4) Recombination (One-point type: two chromosomes exchange a fragment) - b Γ— b a b b a b Before Γ— √ b + a b b b a Create next generation - b b + a b b b a After Γ— √ Γ— b a b b a b Figure 5.7 Process of function finding using GEP algorithm 135
ADE
Fitness function RMSEβ€²β€² RMSEβ€²β€² i i Parsimony pressure Yes Yes Mutation rate 0.01 0.04 Inversion rate 0.1 0.1 Transposition 0.1 0.1 One-point recombination 0.3 0.3 Two-point recombination 0.3 0.3 Gene recombination 0.1 0.1 CART parameter Setting Οƒ I RB RB Initial inputs UCS,E,Οƒ Οƒ ,K,Ο…,E,ρ v RB Excluded parameter - ρ Minimum number of cases 3 3 for parent node Minimum number of cases 1 1 for child node Minimum change of 0.0005 0.0003 impurity level Maximum tree depth 6 5 Number of intervals 10 10 Impurity measure LSD LSD Total number of nodes 27 33 5.3.2.2. Rockburst Assessment Using Classification and Regression Tree (CART) Decision tree as a powerful subset of data-mining techniques has been used in different real- world applications for different aims such as decision making, classification, prediction, pattern recognition, etc. (Kantardzic 2003; Hasanipanah et al. 2017a) A decision tree is a tree comprising a root node (i.e. a parameter that can provide maximum degree of discrimination), some internal nodes representing input parameters, branches which link the nodes together and contain the binary questions regarding the internal nodes, and some leaf nodes representing the solutions (predicted value or a specific class of the dependent parameter). Each path from the root node to a leaf node can be summarised as a rule that this feature makes the decision tree to be known as a rule-based algorithm (Mahjoobi and Etemad-Shahidi 2008). Based on the type of dependent parameter, i.e. being continuous or categorical, the established tree structure is nominated as regression tree (RT) or classification tree (CT), respectively. The decision tree has several subgroups such as ID3 (Quinlan 1986), C4.5, C5.0, CART, CHAID, Exhaustive CHAID, and QUEST (Mahjoobi and Etemad-Shahidi 2008) which have been used for different aims by scholars (Khandelwal et al. 2017; Ghasemi et al. 2017). Among these techniques, the CART algorithm introduced by Breiman et al. (1984) has several advantages that distinguish it among other decision tree algorithms. This algorithm, despite the parametric statistical techniques (e.g. regression analyses), is inherently non-parametric (rule-based), i.e. no 138
ADE
assumption is made with the distribution of values of the independent parameters. On the other hand, CART can handle the highly skewed (multimodal) quantitative data as well as the qualitative parameters with ordinal or non-ordinal structures (Breiman et al. 1984; Salimi et al. 2016). In this algorithm, it is not necessary to eliminate the multicollinearity between the independent parameters. Moreover, CART algorithm can be applied on a database with no homogeneity. CART also can handle the existence of outliers in the raw database by isolating them into a separate node. Because of the mentioned advantages, flexibility, and practical output (tree structure) of this algorithm, it was used in this study for the prediction of rockburst parameters obtained from true-triaxial tests. As a matter of fact, since the output parameters in this study (i.e. 𝜎 and 𝐼 ) are continuous, the aim is to develop two regression trees (RTs) for each 𝑅𝐡 𝑅𝐡 parameter. The process of RT building in CART algorithm focuses mainly on the three following components: (1) a group of questions in the form of 𝑋 ≀ π‘Ž? where 𝑋 is an input parameter and π‘Ž is a constant value in a range that the parameter 𝑋 varies. In CART, the response to this type of question is β€œyes” or β€œno”; (2) the best split on a parameter is determined using a split criterion; (3) calculation of summary statistics for internal nodes. The goal in CART modelling is to create sub-nodes (children) which are more homogeneous and purer than parent nodes based upon the reduction in impurity or improvement score. The term β€œpure” is related to the values of given parameter i.e. in the complete pure node, all cases have a similar value of the splitting parameter and consequently, the node’s variance equal to zero. This issue is compared for all the input parameters and the best improvement is chosen for splitting. This procedure continues until one of the stopping conditions is triggered (Breiman et al. 1984). In CART, the least squared deviation (LSD) is used as an impurity measure. The LSD function for splitting a parent node 𝑑 into two newly generated sub-nodes 𝑑 and 𝑑 can be 𝐿(𝑒𝑓𝑑) 𝑅(π‘–π‘”β„Žπ‘‘) calculated using the following equation (Breiman et al. 1984; Bevilacqua et al. 2003): Ξ¦ = 𝑅2(𝑑)βˆ’π‘ 𝑅2(𝑑 )βˆ’π‘ 𝑅2(𝑑 ) = 1 βˆ‘ [𝑦 βˆ’π‘¦Μ…(𝑑)]2 βˆ’π‘ 1 βˆ‘ [𝑦 βˆ’ (𝑑) 𝐿 𝐿 𝑅 𝑅 𝑁(𝑑) π‘–πœ–π‘‘ 𝑖 𝐿 𝑁(𝑑𝐿) π‘–πœ–π‘‘πΏ 𝑖 𝑦̅(𝑑 )]2 βˆ’π‘ 1 βˆ‘ [𝑦 βˆ’π‘¦Μ…(𝑑 )]2 (5.13) 𝐿 𝑅 𝑁(𝑑𝑅) π‘–πœ–π‘‘π‘… 𝑖 𝑅 where 𝑅2(𝑑 ) is the weighted variance related to the sub-node (child) 𝑑 , 𝑝 is the proportion π‘₯ π‘₯ 𝐿 of cases in parent node 𝑑 which are classified in the left node (𝑑 ), 𝑝 is the proportion of cases 𝐿 𝑅 in parent node 𝑑 which are classified in the right node (𝑑 ), 𝑁(𝑑 ) is the number of cases 𝑅 π‘₯ 139
ADE
classified in sub-node 𝑑 (π‘₯πœ–{𝑅,𝐿}), 𝑦 is the value of the objective parameter for the case 𝑖, π‘₯ 𝑖 𝑦̅(𝑑) is the mean value of parent node, and 𝑦̅(𝑑 ) is the mean value of the sub-node 𝑑 . π‘₯ π‘₯ The best split is obtained by maximizing the Ξ¦ showing the reduction of impurity of an RT (𝑑) model. This splitting process leads to creating a tree structure based on several β€œif-then” rules that make it easy to represent. The splitting process proceeds until each leaf node meets at least one of the stopping criteria. The stopping criteria include: (1) reaching the maximum tree depth; (2) the number of cases (datasets) in the terminal node is less than the predefined minimum parent size; (3) the number of cases in the sub-nodes resulting from the best splits is less than pre-defined minimum child size. The stopping criteria used in this study for CART models are tabulated in Table 6. These criteria and their corresponding values were obtained in such a way that the results provide a good trade-off between the prediction accuracy of regression trees and their complexity (dimension). All these settings also prevent the models from getting stuck in over-fitting problems. The use of a high number of maximum tree depth can lead to producing a large tree structure with high complexity that makes it complicated to use in practice. Additionally, a maximum number of intervals equal to 10 was considered for both models to let the model break down the initial min-max range of each input parameter to different ranges during the splitting process. In this study, the CART models for rockburst parameters (𝜎 and 𝐼 ) were developed using a code written in MatLab R2019a software 𝑅𝐡 𝑅𝐡 environment. To have the same modelling conditions for further assessments, the training and testing datasets used for GEP were fed again to the CART algorithm. According to Table 5.6, for rockburst risk index (𝐼 ) model, like the GEP-based one, the 𝑅𝐡 density (𝜌) parameter has been excluded from the model since the CART benefits from an internal principal component analysis (PCA) that enables it to consider most influential parameters. Figs. 5.10 and 5.11 demonstrate the constructed RTs for 𝜎 and 𝐼 using CART 𝑅𝐡 𝑅𝐡 algorithm, respectively. The tree model of 𝜎 contains 27 nodes and starts with π‘ˆπΆπ‘† as a root 𝑅𝐡 node, while the 𝐼 model has 33 nodes that starts with 𝜎 as the root node. The extraction of 𝑅𝐡 𝑅𝐡 final predicted values of 𝜎 and 𝐼 from these regression trees is an easy task. For instance, 𝑅𝐡 𝑅𝐡 in Fig. 5.10, consider the experimentally measured values of 82.7 MPa, 24.3 GPa, 114.6 MPa, and 108.6 MPa for π‘ˆπΆπ‘†, 𝐸, 𝜎 , and 𝜎 , respectively; by tracking the associated tree structure 𝑣 𝑅𝐡 from the root node (i.e. node1: π‘ˆπΆπ‘†) and the path π‘ˆπΆπ‘† β‰₯ 40.45, 𝜎 < 169.455, 𝜎 β‰₯ 63.3, 𝑣 𝑣 π‘ˆπΆπ‘† < 151.3, 𝜎 < 143.5, and 𝜎 < 118.4, the tree reaches to the leaf node 22 that predicts 𝑣 𝑣 the value of 116.1 for 𝜎 . The same process can be done for 𝐼 as well. As stated at the 𝑅𝐡 𝑅𝐡 140
ADE
sensitivity of the problem. Furthermore, to assess the performance of the developed models in depth, new validation indices have been proposed by other researchers. Golbraikh and Tropsha (2002) defined two indices of π‘˜ and π‘˜β€² to validate the models on testing datasets. In addition, Roy and Roy (2008) proposed an indicator called 𝑅 along with another related parameter of π‘š 𝑅2 to check the predictability of the proposed models. The corresponding values of the π‘˜, π‘˜β€², 𝑂 𝑅 , and 𝑅2 can be calculated based on the measured (β„Ž ) and predicted (𝑑 ) values of the output π‘š 𝑂 𝑖 𝑖 parameters (here are 𝜎 and 𝐼 ). The mathematical expressions of the above indices and their 𝑅𝐡 𝑅𝐡 threshold values are listed in Table 5.7. Taking into account the recommendations, at least a slope of the regression lines (i.e. π‘˜ or π‘˜β€²) through the origin should be close to 1, while π‘˜ is the slope of the regression line when β„Ž is plotted versus 𝑑 , and π‘˜β€² is the regression line when 𝑑 is 𝑖 𝑖 𝑖 plotted versus β„Ž (Golbraikh and Tropsha 2002). The squared correlation coefficient between 𝑖 the predicted and measured values (𝑅2) should be close to 1. The 𝑅 , then, can be calculated 𝑂 π‘š by 𝑅 and 𝑅2 values, and a threshold of > 0.5 is recommended for this index to introduce a 𝑂 model as valid. The foregoing indices were calculated for the developed GEP-based and CART-based models, and their values are listed in Table 5.7. Indeed, the indices of 𝑅, π‘˜, π‘˜β€², 𝑅 , and 𝑅2 were used to verify the validity of the models in testing stage as recommended by π‘š 𝑂 other researchers (Mohammadzadeh et al. 2016; Soleimani et al. 2018). Then, the statistical indices of 𝑅2, 𝑅𝑀𝑆𝐸, and 𝑀𝐴𝐸 were calculated to compare the prediction performance of the GEP and CART models for 𝜎 and 𝐼 based on testing datasets to select the best models. It 𝑅𝐡 𝑅𝐡 can be observed from Table 5.7 that the both proposed models in this study satisfy all the required conditions, and this guarantees that the derived models are strongly credible i.e. the results are not based on chance factor. In addition, comparing the 𝑅2, 𝑅𝑀𝑆𝐸, and 𝑀𝐴𝐸 values of GEP and CART models show that both proposed models have a high degree of accuracy and low estimation error, and subsequently have this capability to be used in practical applications. However, the GEP models of 𝜎 and 𝐼 outperformed the CART models and 𝑅𝐡 𝑅𝐡 have slightly better performance. The next section aims to do a parametric analysis on the selected models (i.e. GEP models) to appraise the effect of the variation of input parameters on the predicted values. 145
ADE
Table 5.7 Statistical indices for the external validation of the developed models Item Formula Threshold 𝜎 𝐼 𝑅𝐡 𝑅𝐡 GEP CART GEP CART 1 βˆ‘π‘› (β„Ž βˆ’β„ŽΜ…)(𝑑 βˆ’π‘‘Μ…) 𝑅>0.8 0.969 0.957 0.972 0.943 𝑅= 𝑖=1 𝑖 𝑖 𝑖 𝑖 βˆšβˆ‘π‘› (β„Ž βˆ’β„ŽΜ…)2βˆ‘π‘› (𝑑 βˆ’π‘‘Μ…)2 𝑖=1 𝑖 𝑖 𝑖=1 𝑖 𝑖 2 βˆ‘π‘› (β„Žπ‘‘) 0.85<π‘˜<1.15 0.934 0.931 0.962 0.981 π‘˜= 𝑖=1 𝑖 𝑖 βˆ‘π‘› β„Ž2 𝑖=1 𝑖 3 βˆ‘π‘› (β„Žπ‘‘) 0.85<π‘˜β€²<1.15 1.046 1.040 1.014 0.971 π‘˜β€²= 𝑖=1 𝑖 𝑖 βˆ‘π‘› 𝑑2 𝑖=1 𝑖 4 𝑅 >0.5 0.763 0.698 0.739 0.596 𝑅 =𝑅2(1βˆ’βˆš|𝑅2βˆ’π‘…2|) π‘š π‘š 𝑂 βˆ‘π‘› (𝑑 βˆ’β„Žπ‘‚)2 Should be close to 1 0.987 0.986 0.997 0.999 𝑅2 =1βˆ’ 𝑖=1 𝑖 𝑖 , 𝑂 βˆ‘π‘› (𝑑 βˆ’π‘‘Μ…)2 𝑖=1 𝑖 𝑖 β„Žπ‘‚ =π‘˜ 𝑑 𝑖 𝑖 5 𝑅2 Should be close to 1 0.939 0.916 0.946 0.889 6 Should be minimum 14.249 16.426 0.195 0.273 𝑛 1 (based on output range) 𝑅𝑀𝑆𝐸= (β„Ž βˆ’π‘‘)2 𝑛 𝑖 𝑖 𝑖=1 7 1 𝑛 Should be minimum 9.803 8.041 0.136 0.187 𝑀𝐴𝐸= 𝑛 |β„Ž π‘–βˆ’π‘‘ 𝑖| (based on output range) 𝑖=1 β„Ž: measured output; 𝑑: predicted output 𝑖 𝑖 5.5. Parametric Analysis To investigate the influence of each input parameter on the predicted values of the corresponding output, a parametric analysis was carried out based on the selected GEP models for 𝜎 and 𝐼 . This analysis also can be another validation for the GEP models by evaluating 𝑅𝐡 𝑅𝐡 how well the results (predicted values) agree with the physical behaviour of the rockburst parameters. To do so, the desired independent parameter should be varied within its range of values, while other independent parameters are constant in their averages. Figs. 5.13 and 5.14 plot the variation of input parameters against the predicted values for rockburst parameters. As it is seen in Fig. 5.13, the 𝜎 increases monotonically in a non-linear fashion with π‘ˆπΆπ‘† and 𝑅𝐡 𝜎 . This result is expected since with the increase of π‘ˆπΆπ‘†, the capacity of the rock to accumulate 𝑣 the strain energy increases, and finally, bursting occurs at a higher stress level violently (Singh 1987). On the other hand, the in-situ stresses, especially the vertical in-situ stress (𝜎 ) are 𝑣 increased in a linear or non-linear relationship with depth (Wagner 2019) and subsequently, due to a high geo-stress state in deep conditions, the 𝜎 is enhanced. However, there are many 𝑅𝐡 fluctuations in 𝜎 values with the increase of Young’s modulus (𝐸) of rocks, but in general, 𝑅𝐡 an increment trend can be seen. It should be mentioned that a parameter may do not show a meaningful relationship solely with the output parameter, while it can be an influential 146
ADE
component in a combination of other parameters in a non-linear form. As mentioned in the GEP modelling section, during the modelling procedure, by applying the variable pressure coefficient, excluding any of the selected three parameters (i.e. π‘ˆπΆπ‘†, 𝐸, and 𝜎 ) from the 𝑣 modelling procedure did not improve the accuracy and complexity of the model. Regarding 𝐼 , a non-linear decreasing trend can be observed for its values with all input 𝑅𝐡 parameters of Young’s modulus (𝐸), Poisson’s ratio (𝜈), horizontal pressure coefficient (𝐾), and rockburst maximum stress (𝜎 ). As can be seen from Fig. 5.14, with the increase of 𝐸 𝑅𝐡 until 20 MPa, the rockburst risk index is decreased suddenly but it remains almost constant with a further increment of 𝐸. Moreover, with the increase of Poisson’s ratio (𝜈) in its range of values, the risk value decreases from 0.473 to 0.40 that according to Table 1, the risk of rockburst occurrence is low. Hence, it seems that the variation of 𝜈 has no significant influence on rockburst risk. However, it is still necessary to do more tests on rocks with a greater range of 𝜈 to check its influence on risk parameter. Generally, the risk of rockburst occurrence for rocks with low strength (or lower 𝜎 ) which are in low depth (or higher 𝐾) is higher than 𝑅𝐡 high-strength rocks in deep conditions. From the results displayed in Figs. 5.13 and 5.14, several non-linear equations between rockburst parameters (𝜎 , and 𝐼 ) and their related 𝑅𝐡 𝑅𝐡 input parameters (except for 𝜎 βˆ’πΈ and 𝐼 βˆ’πΈ) are extracted as follows: 𝑅𝐡 𝑅𝐡 𝜎 = 19.911𝐿𝑛(π‘ˆπΆπ‘†)+10.636, 𝑅2 = 0.9974 (5.14) 𝑅𝐡 𝜎 = βˆ’0.0007𝜎 2 +0.7162𝜎 +41.092, 𝑅2 = 0.9877 (5.15) 𝑅𝐡 𝑣 𝑣 𝐼 = 0.49π‘’βˆ’0.535𝜐, 𝑅2 = 0.9997 (5.16) 𝑅𝐡 𝐼 = βˆ’0.0186𝐾3 +0.2124𝐾2 βˆ’0.7981𝐾 +1.1997, 𝑅2 = 0.9521 (5.17) 𝑅𝐡 𝐼 = 1.2524𝜎 βˆ’0.244, 𝑅2 = 0.9859 (5.18) 𝑅𝐡 𝑅𝐡 According to the above results, clearly can be seen that there is a good correlation between the rockburst parameters and the inputs. These equations provide a series of simple equations for calculating 𝜎 and 𝐼 based on the single rock mechanical parameter as a primary 𝑅𝐡 𝑅𝐡 assessment. These equations may be relevant to investigate rockburst potential. In the end, it is necessary to mention that the developed models in this study are based on the collected datasets and a specific range of values for different parameters. So, for future applications, if the input parameters are out of these ranges, the proposed models should be adjusted again. 147
ADE
2 1 y = -0.0186x3+ 0.2124x2-0.7981x 1.5 0.8 + 1.1997 RΒ² = 0.9521 y = 1.2524x-0.244 B 1 B0.6 RΒ² = 0.9859 R R I I 0.5 0.4 0 0.2 0 2 4 6 0 100 200 300 K Οƒ (MPa) RB Figure 5.14 (Continued) 5.6. Summary and Conclusions As a catastrophic geohazard, rockburst threatens the safety of workers and infrastructures in deep geotechnical conditions. In this study, considering the importance of the stress level that rockburst occurs for different rock types in real stress circumstances, two important rockburst parameters including the maximum rockburst stress (𝜎 ) and rockburst risk index (𝐼 ) were 𝑅𝐡 𝑅𝐡 formulated using the information obtained from true-triaxial unloading tests and two robust data-driven approaches. A comprehensive strategy was applied to the compiled database using the correlation analysis, the agglomerative hierarchical clustering (AHC) technique, and the stepwise selection and elimination (SSE) procedure to provide a homogeneous database free from any outliers, natural groups, and especially, to identify the most influential parameters on 𝜎 and 𝐼 . Then, new non-linear models were developed using robust algorithms of gene 𝑅𝐡 𝑅𝐡 expression programming (GEP) and classification and regression tree (CART). Finally, a parametric analysis was conducted to study the variation of 𝜎 and 𝐼 with the change of 𝑅𝐡 𝑅𝐡 input parameters. The conclusions obtained from this study are presented in the following. The correlation analysis, AHC, SSE, and multiple regression analysis techniques, as recommended and implemented in the current study, have presented promising results by dimension reduction (i.e. eliminating the redundant input parameters) and choosing the statistically significant parameters that affect the rockburst parameters (i.e. 𝜎 and 𝐼 ). This 𝑅𝐡 𝑅𝐡 procedure simplifies the rockburst assessment at the field scale. The obtained dendrogram by AHC analysis (Fig. 5.5) showed that there is no natural group in the compiled database except for two data samples that were identified as outliers and subsequently were eliminated from the original database. Therefore, the database was identified as a homogeneous database for 149
ADE
29 If E in [36.05, 71) and K in [1.721, 5.866) and 𝜎 in [56.8, 255.5) then 𝐼 = 0.185 in 12.8% of cases 𝑅𝐡 𝑅𝐡 30 If E in [14.1, 29.7) and K in [1.721, 5.866) and 𝜎 in [56.8, 255.5) then 𝐼 = 0.292 in 12.8% of cases 𝑅𝐡 𝑅𝐡 31 If E in [29.7, 36.05) and K in [1.721, 5.866) and 𝜎 in [56.8, 255.5) then 𝐼 = 0.547 in 2.3% of cases 𝑅𝐡 𝑅𝐡 32 If 𝜎 in [56.8, 143.6) and E in [36.05, 71) and K in [1.721, 5.866) then 𝐼 = 0.205 in 9.3% of cases 𝑅𝐡 𝑅𝐡 33 If 𝜎 in [143.6, 255.5) and E in [36.05, 71) and K in [1.721, 5.866) then 𝐼 = 0.133 in 3.5% of cases 𝑅𝐡 𝑅𝐡 References Akdag S, Karakus M, Taheri A, et al (2018) Effects of Thermal Damage on Strain Burst Mechanism for Brittle Rocks Under True-Triaxial Loading Conditions. Rock Mechanics and Rock Engineering 51(6):1–26 Alavi AH, Gandomi AH (2011) A robust data mining approach for formulation of geotechnical engineering systems. Engineering Computations 28(3):242–274 Alavi AH, Hasni H, Lajnef N, et al (2016) Damage detection using self-powered wireless sensor data: An evolutionary approach. Measurement: Journal of the International Measurement Confederation 82:254–283 Archontoulis S V, Miguez FE (2015) Nonlinear regression models and applications in agricultural research. Agronomy Journal 107(2):786–798 Armaghani DJ, Faradonbeh RS, Rezaei H, et al (2016) Settlement prediction of the rock- socketed piles through a new technique based on gene expression programming. Neural Computing and Applications 29(11):1115-1125 Bagde MN, PetroΕ‘ V (2005) Fatigue properties of intact sandstone samples subjected to dynamic uniaxial cyclical loading. International Journal of Rock Mechanics and Mining Sciences 42(2):237–250 Barquins M, Petit J-P (1992) Kinetic instabilities during the propagation of a branch crack: effects of loading conditions and internal pressure. Journal of Structural Geology 14(8- 9):893–903 Bethea RM (2018) Statistical methods for engineers and scientists. Routledge Bevilacqua M, Braglia M, Montanari R (2003) The classification and regression tree approach to pump failure rate analysis. Reliability Engineering & System Safety 79(1):59–67 Breiman L, Friedman J, Stone CJ, Olshen RA (1984) Classification and Regression Trees. CRC press 152
ADE
Cai M, Kaiser P (2018) Rockburst support reference bookβ€”volume I: rockburst phenomenon and support characteristics. Laurentian University Chen G, Li T, Wang W, et al (2019) Weakening effects of the presence of water on the brittleness of hard sandstone. Bulletin of Engineering Geology and the Environment 78(3):1471–1483 Chen J, Su G (2010) True triaxial experimental study on hard rock under high geo-stress condition based on excavation and support. In: 2010 International Conference on Mechanic Automation and Control Engineering. pp 1465–1467 Cheon DS, Keon S, Park C, Ryu C (2006) An experimental study on the brittle failure under true triaxial conditions. Tunnelling and underground space technology 21(3-4) Cook NGW (1963) The basic mechanics of rockbursts. Journal of the Southern African Institute of Mining and Metallurgy 64(3):71–81 Cortez P, Embrechts MJ (2011) Opening black box Data Mining models using Sensitivity Analysis. In: IEEE Symposium on Computational Intelligence and Data Mining Faradonbeh RS, Monjezi M (2017) Prediction and minimization of blast-induced ground vibration using two robust meta-heuristic algorithms. Engineering with Computers 1–17 Feng XT, Chen B, Feng G, Zhao Z, Zheng H (2017) Description and engineering phenomenon of rockbursts. In: Rockburst: Mechanisms, Monitoring, Warning, and Mitigation. first ed. Elsevier-Health Sciences Division 3-17 Ferreira C (2006) Gene expression programming: mathematical modeling by an artificial intelligence. Springer Ferreira C (2002) Gene Expression Programming in Problem Solving. In: Roy R, KΓΆppen M, Ovaska S, et al. (eds) Soft Computing and Industry: Recent Applications. Springer London, London, pp 635–653 Ghasemi E (2017) Particle swarm optimization approach for forecasting backbreak induced by bench blasting. Neural Computing and Applications 28(7):1855–1862 Ghasemi E, Kalhori H, Bagherpour R (2017) Stability assessment of hard rock pillars using two intelligent classification techniques: A comparative study. Tunnelling and Underground Space Technology 68:32–37 153
ADE
Golbraikh A, Tropsha A (2002) Beware of q2! Journal of Molecular Graphics and Modelling 20(4):269–276 Hasanipanah M, Faradonbeh RS, Amnieh HB, et al (2017a) Forecasting blast-induced ground vibration developing a CART model. Engineering with Computers 33(2):307-316 Hasanipanah M, Faradonbeh RS, Armaghani DJ, et al (2017b) Development of a precise model for prediction of blast-induced flyrock using regression tree technique. Environmental Earth Sciences 76(1) He J, Dou L, Gong S, et al (2017) Rock burst assessment and prediction by dynamic and static stress analysis based on micro-seismic monitoring. International Journal of Rock Mechanics and Mining Sciences 100(93):46–53 He M (2009) The mechanism of rockburst and its countermeasure of support. Consultation Report for the Key Technology of Safe and Rapid Construction for Jinping LI Hydropower Station High Overburden and Long Tunnels 23–28 He M, e Sousa LR, Miranda T, Zhu G (2015) Rockburst laboratory tests database - Application of data mining techniques. Engineering Geology 185:116–130 He M, Gong W, Wang J, et al (2014) Development of a novel energy-absorbing bolt with extraordinarily large elongation and constant resistance. International Journal of Rock Mechanics and Mining Sciences 67:29–42 He M, Ren F, Liu D (2018) Rockburst mechanism research and its control. International Journal of Mining Science and Technology 28(5):829–837 He M, Xia H, Jia X, et al (2012) Studies on classification, criteria and control of rockbursts. Journal of Rock Mechanics and Geotechnical Engineering 4(2):97–114 He MC, Miao JL, Feng JL (2010) Rock burst process of limestone and its acoustic emission characteristics under true-triaxial unloading conditions. International Journal of Rock Mechanics and Mining Sciences 47(2):286–298 Hemmateenejad B, Yazdani M (2009) QSPR models for half-wave reduction potential of steroids: A comparative study between feature selection and feature extraction from subsets of or entire set of descriptors. Analytica Chimica Acta 634(1):27–35 Hoseinian FS, Faradonbeh RS, Abdollahzadeh A, et al (2017) Semi-autogenous mill power 154
ADE
model development using gene expression programming. Powder Technology 308:61-69 Hudaverdi T (2012) Application of multivariate analysis for prediction of blast-induced ground vibrations. Soil Dynamics and Earthquake Engineering 43:300–308 Jahed Armaghani D, Faradonbeh RS, Momeni E, et al (2017) Performance prediction of tunnel boring machine through developing a gene expression programming equation. Engineering with Computers 34(1):129-141 James G, Witten D, Hastie T, Tibshirani R (2013) An introduction to statistical learning. Springer Jian Z, Xibing L, Xiuzhi S (2012) Long-term prediction model of rockburst in underground openings using heuristic algorithms and support vector machines. Safety Science 50(4):629–644 Kantardzic M (2003) Data Mining: Concepts, models, methods, and algorithms. John Wiley & Sons Kaufman L, Rousseeuw PJ (2009) Finding groups in data: an introduction to cluster analysis. John Wiley & Sons Kayadelen C (2011) Soil liquefaction modeling by Genetic Expression Programming and Neuro-Fuzzy. Expert Systems with Applications 38(4):4080–4087 Khandelwal M, Armaghani DJ, Faradonbeh RS, et al (2017) Classification and regression tree technique in estimating peak particle velocity caused by blasting. Engineering with Computers 33(1):45–53 Kumar Sharma S, Rai P (2017) Establishment of blasting design parameters influencing mean fragment size using state-of-art statistical tools and techniques. Measurement 96:34–51 Li D, Zhao F, Zheng M (2014) Fractal characteristics of cracks and fragments generated in unloading rockburst tests. International Journal of Mining Science and Technology 24(6):819–823 Li T, Cai MF, Cai M (2007) A review of mining-induced seismicity in China. International Journal of Rock Mechanics and Mining Sciences 44:1149–1171 Linkov AM (1996) Rockbursts and the instability of rock masses. International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts 33(7):727–732 155
ADE
Liu X, Liang Z, Zhang Y, et al (2018) Experimental study on the monitoring of rockburst in tunnels under dry and saturated conditions using AE and infrared monitoring. Tunnelling and Underground Space Technology 82:517–528 Mahjoobi J, Etemad-Shahidi A (2008) An alternative approach for the prediction of significant wave heights based on classification and regression trees. Applied Ocean Research 30(3):172–177 Mitchell TM (1997) Does machine learning really work? AI Magazine 18:11–20 Mohammadzadeh S D, Bolouri Bazaz J, Vafaee Jani Yazd SH, Alavi AH (2016) Deriving an intelligent model for soil compression index utilizing multi-gene genetic programming. Environmental Earth Sciences 75(3):262 Montgomery DC, Peck EA, Vining GG (2012) Introduction to Linear Regression Analysis. Wiley Motulsky HJ, Ransnas LA (1987) Fitting curves to data using nonlinear regression: a practical and nonmathematical review. The FASEB journal 1(5):365–374 Nazari A, Pacheco Torgal F (2013) Modeling the compressive strength of geopolymeric binders by gene expression programming-GEP. Expert Systems with Applications 40(14):5427–5438 Power HE, Gharabaghi B, Bonakdari H, et al (2019) Prediction of wave runup on beaches using Gene-Expression Programming and empirical relationships. Coastal Engineering 144:47–61 Pu Y, Apel DB, Xu H (2019) Rockburst prediction in kimberlite with unsupervised learning method and support vector classifier. Tunnelling and Underground Space Technology 90:12–18 Quinlan JR (1986) Induction of decision trees. Machine Learning 1:81–106 Ranjith PG, Zhao J, Ju M, et al (2017) Opportunities and Challenges in Deep Mining: A Brief Review. Engineering 3(4):546–551 Roy PP, Roy K (2008) On Some Aspects of Variable Selection for Partial Least Squares Regression Models. QSAR & Combinatorial Science 27(3):302–313 Roy R, KΓΆppen M, Ovaska S, et al (2002) Soft Computing and Industry: Recent Applications 156
ADE
Saharan MR, Mitri H (2011) Destress blasting as a mines safety tool: Some fundamental challenges for successful applications. In: Procedia Engineering. Elsevier 26(37-47) pp 37–47 Salimi A, Faradonbeh RS, Monjezi M, Moormann C (2016) TBM performance estimation using a classification and regression tree (CART) technique. Bulletin of Engineering Geology and the Environment, 77(1):429-440 Sarkhosh M, Ghasemi JB, Ayati M (2012) A quantitative structure- property relationship of gas chromatographic/mass spectrometric retention data of 85 volatile organic compounds as air pollutant materials by multivariate methods. Chemistry Central journal 6 Suppl 2:S4–S4 Saxena A, Prasad M, Gupta A, et al (2017) A review of clustering techniques and developments. Neurocomputing 267:664–681 Sayadi AR, Lashgari A, Paraszczak JJ (2012) Hard-rock LHD cost estimation using single and multiple regressions based on principal component analysis. Tunnelling and Underground Space Technology 27(1):133–141 Shi XZ, Zhou J, Dong L, et al (2010) Application of unascertained measurement model to prediction of classification of rockburst intensity. Chinese Journal of Rock Mechanics and Engineering 29(1):2720–2726 Shirani Faradonbeh R, Taheri A (2019) Long-term prediction of rockburst hazard in deep underground openings using three robust data mining techniques. Engineering with Computers 35(2):659–675 Singh SP (1987) The influence of rock properties on the occurrence and control of rockbursts. Mining Science and Technology 5(1):11–18 SMITH GN (1986) Probability and statistics in civil engineering. Collins Professional and Technical Books 244 Soleimani S, Rajaei S, Jiao P, et al (2018) New prediction models for unconfined compressive strength of geopolymer stabilized soil using multi-gen genetic programming. Measurement 113:99–107 Sousa LR (2010) Risk Analysis for Tunneling Projects. Massachusetts Institute of Technology, Cambridge 157
ADE
Su G, Chen Z, Ju JW, Jiang J (2017a) Influence of temperature on the strainburst characteristics of granite under true triaxial loading conditions. Engineering Geology 222:38–52 Su G, Jiang J, Zhai S, Zhang G (2017b) Influence of Tunnel Axis Stress on Strainburst: An Experimental Study. Rock Mechanics and Rock Engineering 50(6):1551–1567 Sun X, Xu H, Zheng L, et al (2016) An experimental investigation on acoustic emission characteristics of sandstone rockburst with different moisture contents. Science China Technological Sciences 59(10):1549–1558 Tang C, Wang J, Zhang J (2010) Preliminary engineering application of microseismic monitoring technique to rockburst prediction in tunneling of Jinping II project. Journal of Rock Mechanics and Geotechnical Engineering 2(3):193–208 Wagner H (2019) Deep Mining: A Rock Engineering Challenge. Rock Mechanics and Rock Engineering, 1417-1446 Wang J-A, Park HD (2001) Comprehensive prediction of rockburst based on analysis of strain energy in rocks. Tunnelling and Underground Space Technology 16(1):49–57 Wang X, Huang R (1998) Analysis of deformation and failure features characteristics of rock under unloading conditions and their effects on rock burst. Mountain Research 16(4):281– 285 Wang Y, He M, Liu D, Gao Y (2019) Rockburst in Sandstone Containing Elliptic Holes with Varying Axial Ratios. Advances in Materials Science and Engineering Weng L, Huang L, Taheri A, Li X (2017) Rockburst characteristics and numerical simulation based on a strain energy density index: A case study of a roadway in Linglong gold mine, China. Tunnelling and Underground Space Technology 69:223–232 Xu LS (2003) Research on the experimental rock mechanics of rockburst under unloading condition. Journal of Chongqing Jiaotong University 22(1):1–4 Zhao XG, Cai M (2015) Influence of specimen height-to-width ratio on the strainburst characteristics of Tianhu granite under true-triaxial unloading conditions. Canadian Geotechnical Journal 52(7):890–902 Zhao XG, Wang J, Cai M, et al (2014) Influence of Unloading Rate on the Strainburst Characteristics of Beishan Granite Under True-Triaxial Unloading Conditions. Rock 158
ADE
Title of Paper Post-peak behaviour of rocks under cyclic loading using a double-criteria damage-controlled test method Publication Status Published Accepted for Publication Submitted for Publication U npublished and Unsubmitted work written in manuscript style Publication Details Shirani Faradonbeh R, Taheri A, Karakus M (2021) Post-peak behaviour of rocks under cyclic loading using a double-criteria damage-controlled tests method. Bulletin of Engineering Geology and the Environment 80(2):1713–1727 Principal Author Name of Principal Author (Candidate) Roohollah Shirani Faradonbeh Contribution to the Paper Literature review, conducting the laboratory tests, analysis of the results and preparation of the manuscript Overall percentage (%) 80% Certification: This paper reports on original research I conducted during the period of my Higher Degree by Research candidature and is not subject to any obligations or contractual agreements with a third party that would constrain its inclusion in this thesis. I am the primary author of this paper. Signature Date 17 June 2021 Co-Author Contributions By signing the Statement of Authorship, each author certifies that: i. the candidate’s stated contribution to the publication is accurate (as detailed above); ii. permission is granted for the candidate in include the publication in the thesis; and iii. the sum of all co-author contributions is equal to 100% less the candidate’s stated contribution. Name of Co-Author Abbas Taheri Contribution to the Paper Research supervision, review and revision of the manuscript Signature Date 21 June 2021 Name of Co-Author Murat Karakus Contribution to the Paper Review and revision of the manuscript Signature Date 21 June 2021 Chapter 6 160
ADE
Post-Peak Behaviour of Rocks Under Cyclic Loading Using a Double-Criteria Damage- Controlled Test Method Abstract Cyclic loading-induced hazards are severe instability problems concerning surface and underground geotechnical projects. Therefore, it is crucial to understand the rock failure mechanism under cyclic loading. An innovative double-criteria damage-controlled testing method was proposed in this study to capture the complete stress-strain response of porous limestone, especially the post-peak behaviour, under systematic cyclic loading. The proposed test method was successful in applying the pre-peak cyclic loading and then in controlling the self-sustaining failure of rock during the post-peak cyclic loading. The results showed that the strength of the rock specimens slightly increased with an increase in the fatigue life in the pre- peak region due to cyclic loading-induced hardening. Additionally, a combination of class I and class II behaviours was observed in the post-peak region during the cyclic loading tests; the class II behaviour was more dominant by the increase in fatigue life in the pre-peak region. Damage evolution was assessed based on several parameters, such as the elastic modulus, energy dissipation ratio, damage variable and crack damage threshold stress, both in the pre- peak and post-peak regions. It was found that when the cyclic loading stress is not close to the peak strength, due to a coupled mechanism of dilatant microcracking and grain crushing and pore filling, quasi-elastic behaviour dominates the cyclic loading history, causing more elastic strain energy to accumulate in the specimens. Keywords Cyclic loading, Pre-peak and post-peak behaviour, Damage, Crack damage threshold stress, Strength hardening 6.1. Introduction Surface and underground structures are usually exposed to environmental and human-induced cyclic loadings such as earthquakes, wind, volcanism, drilling and blasting, mechanical 161
ADE
excavation and mining seismicity, which threaten their long-term stability (Taheri et al. 2016; Munoz et al. 2016a). Therefore, it is necessary to evaluate the time-dependent behaviour of rocks under cyclic loading. In rock engineering, understanding the fatigue response of rocks is of particular interest since rock stability conditions vary significantly under cyclic loading. A great majority of rock fatigue studies have reported on the reduction in rock strength due to cyclic loading (Bagde and PetroΕ‘ 2005). However, there are very few studies that have illustrated strength hardening when the cyclic stress level is low enough to prevent failure during cyclic loading (Burdine 1963; Singh 1989; Ma et al. 2013; Taheri et al. 2017). Unlike the static and quasi-static loadings, which the applied load/deformation increases/decreases continuously, cyclic loading is described by a time-dependent displacement/load signal with a repetitive pattern. The loading rate in cyclic experiments is relatively high and propagates waves, and their superposition causes a stress distribution different from that induced by quasi- static loading (Cho et al. 2003). In recent decades, many studies have investigated the mechanical behaviour of rocks under different cyclic loading histories and loading conditions. Most of these studies have reported the results of tests performed under uniaxial compression (Attewell and Sandford 1974; Eberhardt et al. 1999), which can replicate the stress state in mining pillars and around galleries. Other studies have focused on triaxial compression conditions with different confining pressures (Munoz et al. 2016a; Zhou et al. 2019) and indirect tensile tests (Ghamgosar and Erarslan 2016), which are useful to calibrate the advanced constitutive laws and to estimate the tensile strength of a material, respectively. In addition, few cyclic studies of flexural tests (three-point and four-point) (Cardani and Meda 2004) and freeze-thaw tests (Zhang et al. 2019a) can be found in the literature. In prior studies, the fatigue properties of rocks were found to be dependent on the loading stress level, amplitude, frequency, waveform and loading and unloading rate. Rock behaviour in the post-peak region under uniaxial compression is characterised by either class I or class II behaviour (Fig. 6.1). The former is defined by a negative post-peak modulus describing a stable fracture propagation and the need to do more work on the specimen to degrade its load-bearing capacity, while the latter represents a positive post-peak modulus (i.e., snap-back behaviour) describing a self-sustaining (brittle) failure (Wawersik and Fairhurst 1970; Munoz et al. 2016b). The proper measurement of the post-peak behaviour of rocks can be a useful tool for quantifying the post-peak fracture energy and rock brittleness that can be employed to optimise the designation of surface and underground structures and to mitigate possible hazards (Akinbinu 2016). For instance, to evaluate the proneness and intensity of the 162
ADE
rockburst phenomenon near underground excavation in deep underground conditions, post- peak analysis of the rocks in terms of strain energy evolution is required. In other words, the rockburst hazard in deep underground openings is associated with not only internal strain energy accumulation but also seismic disturbances induced by external sources (Xuefeng et al. 2010). Therefore, the post-peak response of rocks subjected to cyclic loading can unveil the mechanism of geotechnical hazards such as rockburst and provide practical tools for their assessment. As shown in Fig. 6.2, the cyclic loading of rock can be undertaken following two main loading methods: 1. Systematic cyclic loading: These tests have a constant loading amplitude, π΄π‘šπ‘.(𝜎 ), π‘Ž and can be conducted as single-level (Fig. 6.3a) or multi-level (Fig. 6.3b) testes under load-controlled or displacement-controlled (i.e., axial and lateral displacement- controlled) loading conditions. In both load-controlled and displacement-controlled conditions, the post-peak behaviour cannot be obtained, as the axial load level is the only criterion to define the amount of the load that a specimen should be subjected to during cyclic loading, until failure or even after failure. As a result, the specimen fails during cyclic loading in an uncontrolled manner, and the post-peak response cannot be obtained. Figs. 6.4a-d demonstrate the single-level and multi-level systematic cyclic tests conducted by different researchers under load-controlled and displacement- controlled conditions. As shown in these figures, in all the tests, failure occurred in an uncontrolled manner, and post-peak behaviour was not obtained. Prior systematic cyclic loading studies found that failure occurs at a stress level lower than the determined monotonic strength owing to the strength weakening process. As such, the accumulation of irreversible deformation (plastic strains) is not constant during the experiment, while the hysteresis loops follow a loose-dense-loose law (Xiao et al. 2009). 2. Damage-controlled cyclic loading: These tests involving incremental loading amplitude can be conducted in a load-based mode (Fig. 6.3c) or displacement-based mode (Fig. 6.3d). The former can be conducted either in load-controlled or displacement-controlled loading conditions (i.e., axial and lateral displacement-controlled). However, the post- peak response cannot be obtained, as the specimen might experience an uncontrolled failure when it is forced to reach a pre-defined stress level. Figs. 6.4e and f show representative results. A displacement-based test can be undertaken in either axial or lateral displacement-controlled conditions. In this type of damage-controlled test, axial 163
ADE
stress is reversed when a certain amount of axial or lateral displacement is achieved in a loading cycle. Munoz et al. (2016b) showed that under uniaxial loading conditions, soft, medium-strong and strong rocks demonstrate either class II or a combination of class I and class II post-peak behaviours. As a result, the post-peak response cannot be adequately measured when the test is controlled by axial displacement (Fig. 6.4g). However, by using lateral strain to control the amount of damage in a damage- controlled test, the post-peak behaviour of a brittle rock can be achieved successfully (Fig. 6.4h). From prior damage-controlled cyclic loading studies, it is reported that failure occurs at a stress level close to or lower than the determined monotonic strength. Moreover, the rate of strain accumulation under this type of loading is lower than that during systematic cyclic tests (Cerfontaine and Collin 2018). It should be noted that previous studies have mostly focused on the influence of cyclic loadings on the mechanical rock properties and damage evolution in the pre-peak region. There are, however, a few studies investigating failure behaviour and deformation localisation during post-peak cyclic loading (e.g., Munoz and Taheri 2017a, 2019). Given the above, to the best of our knowledge, no study has investigated the post-peak response of rocks subjected to pre-peak systematic cyclic loading. This is because failure cannot be controlled when a constant axial load is achieved in every cycle in a systematic cyclic loading. In addition, in a damage- controlled test in which the lateral displacement is used to control the damage, an axial load is reversed when a certain amount of lateral strain occurs. Therefore, systematic cyclic loading cannot be applied in such a way that the load is always reversed at a constant stress level in the pre-peak region. However, rock material in engineering applications (e.g., mining pillars in deep underground conditions) may be subjected to systematic pre-peak cyclic loading and then post-failure cyclic loading. Thus, it is significant to investigate the behaviour of rock subjected to this loading condition. In this study, for the first time, a new cyclic test method considering two cyclic loading control criteria is proposed to capture the complete response of rocks, especially the post-peak behaviour, under cyclic loading. The proposed test method is a combination of multi-level systematic cyclic loading and lateral displacement-based damage- controlled cyclic loading to control both the damage and the rate of cyclic loading (see Fig. 6.2). A critical analysis is carried out to investigate damage evolution in both the pre-peak and post-peak regions, and the influences of pre-peak cyclic loading on the peak strength, crack damage threshold stress and rock stiffness are evaluated in more detail. 164
ADE
cyclic loading load-controlled test (Li et al. 2019), d multi-level systematic cyclic loading axial displacement-controlled test (Liu et al. 2014), e load-based damage controlled cyclic loading load-controlled test (Guo et al. 2018), f load-based damage controlled cyclic loading axial displacement-controlled test (Heap et al. 2010), g displacement-based damage controlled cyclic loading axial displacement-controlled test and (Wang et al. 2019) h displacement-based damage controlled cyclic loading lateral displacement-controlled test (Munoz and Taheri 2019) 6.2. Experimental Methodology 6.2.1. Tuffeau Limestone Specimens Tuffeau limestone is used in this study to undertake double-criteria damage-controlled cyclic loading tests (Fig. 6.5a). The name of this rock comes from the Latin word tofus, meaning spongy rock. This yellowish-white sedimentary rock is a local limestone of the Loire Valley in France and was deposited in the middle Turonian of the Upper Cretaceous, approximately 90 million years ago. This rock type is usually extracted from surface and underground quarries and is used mostly in the building industry (Beck and Al-Mukhtar 2014). X-ray diffraction (XRD) (Fig. 6.5b) and scanning electronic microscopy (SEM) analyses (Fig. 6.5c) were carried out on collected limestone specimens to identify their mineralogical components and microstructural characteristics. Two main crystalline phases, calcite (CaCO ) (β‰…50%) and 3 silica (SiO ) (β‰…30%), which has the two forms of quartz and opal cristobalite-tridymite (opal- 2 CT), were identified. Other phases, such as mica and clayey minerals (e.g., muscovite, biotite, smectite, and glauconite) (β‰…20%), are disseminated in this limestone. Tuffeau limestone has an average density of 1.43 g/cm3 and is a lightweight and fine-grained limestone with a complex porous network (total porosity of 45Β±5%). The arrangement of grains with different sizes contributes to the creation of micropores and macropores within the rock texture (Al- Mukhtar and Beck 2006). The rock specimen in Fig. 6.5c has a heterogeneous porous structure, and the microcracks, microcavities, and quartz are the main components controlling the macrofailure of the specimen under loading conditions. The cylindrical rock specimens with diameters and lengths of 42 mm and 100 mm, respectively (i.e., an aspect ratio of 2.4), were cored from a single rock block and prepared to be smooth and straight according to the ISRM standards (Fairhurst and Hudson 1999) to minimise the end friction effects and to ensure a uniform stress state within the specimen during loading. Additionally, the diameter of the 168
ADE
tests to capture the rock behaviour before and after peak stress. The axial load (acquired by a load cell), axial strain (acquired by a pair of LVDTs), and lateral strain (acquired by a chain extensometer) were recorded simultaneously during the tests by a data acquisition system at a rate of 10 data points per second (see Fig. 6.6a). Five uniaxial monotonic tests were conducted under the lateral strain rate of 0.02Γ—10-4/s to satisfy the static to quasi-static loading conditions (Munoz and Taheri 2017b). These monotonic tests provide a reference for defining the stress levels of cyclic uniaxial compression tests. The time history of the loading (𝜎 ), axial strain π‘Ž (πœ€ ), and lateral strain (πœ€ ) for a typical monotonic loading test is shown in Fig. 6.6b. As seen π‘Ž 𝑙 in this figure, in the pre-peak and the post-peak regions, the lateral strain (πœ€ ) increases 𝑙 monotonically with time, maintaining a constant lateral strain rate throughout the test, and the complete post-peak response is obtained in a straightforward manner using the lateral strain- controlled technique. Fig. 6.6c shows the normalised stress-strain curves obtained from the five uniaxial monotonic tests. The specimens have an average monotonic compressive strength and Young’s modulus of 7.39 MPa and 1.67 GPa. As seen from Fig. 6.6c, in the post-peak region, the axial stress and axial strain fluctuate successively due to the coupled mechanism of strength degradation induced by the coalescence of the macrocracks and strength recovery induced by interlocking the sides of the macrocracks. However, the total behaviour of all the conducted monotonic tests in the post-peak region is a combination of class I and class II behaviours, which is consistent with the results reported by Munoz et al. (2016a). Additionally, the conducted monotonic tests exhibit similar behaviour both in the pre-peak and the post-peak regions, which shows the low discrepancy among the tested specimens. 170
ADE
(here, 6 MPa) is reached. In this stage, the axial stress and lateral strain feedback signals received from the load cell and the chain extensometer, respectively, are continuously compared with the program signals (i.e., the user-defined values) and the errors, if any, are adjusted by the servo-controller. By doing so, it is guaranteed that the axial load is always applied under a constant lateral strain rate and that the axial load does not exceed the initial stress level defined for cyclic loading. Thereafter, the specimen is unloaded until the axial stress is equal to 0.07 MPa, ensuring that the specimen is always in complete contact with the loading platens. b) Afterwards, cyclic loading is applied under a constant lateral strain rate for a specific number of cycles (i.e., 400 cycles). Two criteria are adopted to control the failure: a maximum axial stress level that can be achieved and a maximum lateral strain amplitude that a Tuffeau limestone specimen is allowed to experience in a cycle during loading, π΄π‘šπ‘.(πœ€ ). In this study, the initial maximum stress level (i.e., the first 𝑙 criterion) is adopted to be equal to 6.0 MPa. The optimum values for π΄π‘šπ‘.(πœ€ ) and the 𝑙 loading rate (π‘‘πœ€ /𝑑𝑑) were determined based on a previous study conducted by Munoz 𝑙 and Taheri (2017a) on Tuffeau limestone and the results obtained from the trial tests to avoid the sudden failure of a specimen in an uncontrolled manner. Therefore, different loading rates and π΄π‘šπ‘.(πœ€ ) values were evaluated by performing four trial cyclic 𝑙 loading tests, and finally, π΄π‘šπ‘.(πœ€ ) = 17Γ—10βˆ’4 and π‘‘πœ€ /𝑑𝑑 = 2Γ—10βˆ’4/𝑠 were 𝑙 𝑙 obtained by balancing the capability of the methodology in capturing the post-peak behaviour of the rock and completing the test in the shortest possible time. The axial load is reversed when at least one criterion is met. By following the closed-loop procedure shown in Fig. 6.7, the test is continued until the specimen fails or until 400 cycles are completed. c) If the specimen does not fail after 400 cycles, the specimen is monotonically loaded under a constant lateral strain rate of 0.02Γ—10-4/s until the specimen is under an axial load of 6.5 MPa (i.e., a 0.5 MPa increase in the stress level compared to the previous stress level in this multi-level cycling loading scheme). If the specimen fails during monotonic loading, the complete post-peak behaviour is measured during lateral strain- controlled loading. d) The procedure explained in b and c is repeated until the specimen fails. Fig. 6.8 shows typical results for a Tuffeau limestone specimen. As shown in this figure, after initial monotonic loading under the constant loading rate of 0.02Γ—10-4/s, the prescribed axial 172
ADE
stress level (i.e., 6 MPa) is reached. Afterwards, the specimen is unloaded monotonically, and then cyclic loading is applied under a constant lateral strain rate of 2Γ—10-4/s. At the first step of cyclic loading, the amplitude of lateral strain, π΄π‘šπ‘.(πœ€ ), is relatively low (6Γ—10-4/s after 200 𝑙 cycles), and the first criterion is always met during cyclic loading (i.e., the stress level remaining below 6 MPa). As the specimen does not fail after 400 cycles, the axial load is increased monotonically to the second stress level (i.e., 6.5 MPa), and the cyclic loading procedure is repeated. As shown in Fig. 8, in the second series of cyclic loading at the onset of the failure, the lateral strain amplitude, π΄π‘šπ‘.(πœ€ ), is equal to 17Γ—10-4. After this cycle, the 𝑙 second criterion controls the cyclic loading, and the strength degradation during post-peak cyclic loading is observed until complete failure. By doing so, the complete post-peak behaviour of the Tuffeau limestone under systematic cyclic loading can be successfully observed. 173
ADE
10 SL 1= 6 MPa SL 2= 6.5 MPa 10 0 0.07 MPa de/d=0.02 l t -20 de/d=0.02 l t 6 MPa 6 MPa -40 6.5 MPa Failure point de/d= 2 l t -60 Amp. (e l)= 6 de l/d t= 2 -4 de/d= 2Β΄10 L t Time Amp (e)= 16 l Amp (e)= 17 l Amp (e)= 17 l Time -80 0 2000 4000 6000 8000 Time, t (s) Figure 6.8 Typical time-history of axial stress and lateral strain during a double-criteria cyclic 6.3. Experimental Results 6.3.1. Complete Stress-Strain Response In this study, three multi-level systematic cyclic loading tests were conducted using the methodology explained above to evaluate the applicability of the proposed testing method in capturing the failure behaviour of the soft and porous Tuffeau limestone. Fig. 6.9 displays the axial stress-strain relations obtained for these tests, in which 6 MPa was defined as the initial stress level, and the specimens were subjected to systematic cyclic loading at different stress levels, taking 0.5 MPa as the stress increment between consecutive cyclic loading steps. The envelope curves showing the overall behaviour of the specimens in the post-peak region were drawn by connecting the loci of the indicator stresses (π‘ž , the maximum stress of each cycle). 𝑖 As seen from Fig. 6.9, the overall post-peak behaviour of the specimens is characterised by the combination of class I and class II; however, the class I behaviour is more dominant in specimen TL6 (Fig. 6.9a) than in specimens TL7 and TL8 (Figs. 6.9b and c). Table 6.1 summarises the results of the cyclic loading tests. As listed in Table 6.1 and shown in Fig. 6.9, the different cycle numbers and stress levels are recorded for the three specimens before failure; for example, specimen TL8 experienced 2906 cycles before failure, and its failure occurred at 175 4- )aPM( as ,sserts laixA ) 01Β΄( le ,niarts laretaL sserts laixA niarts laretaL sserts laixA niarts laretaL
ADE
6.3.2. Fatigue Damage Evolution Damage can be characterised by the process of generation, propagation and coalescence of mesoscopic defects and voids through solid materials. Damage can be described by the degradation of some material properties, such as stiffness, residual strength, and P-wave velocity. Additionally, damage during cyclic loading can be investigated by the corresponding irreversible strain, dissipative energy, electrical resistance, and acoustic emission counts (Xiao et al. 2010; Taheri and Tatsuoka 2012). The incremental accumulation of plastic deformation during cyclic loading contributes to the degradation of the cohesive strength and stiffness of the rocks. Therefore, the irreversible strain can be regarded as a suitable indicator for fatigue damage assessment. Hence, a damage variable (𝐷) was defined based on the accumulation of irreversible axial strain (πœ€π‘–π‘Ÿπ‘Ÿ) (see Fig. 6.10) after each loading and unloading cycle as follows: π‘Ž βˆ‘π‘š (πœ€π‘–π‘Ÿπ‘Ÿ) 𝐷 = 𝑖=1 π‘Ž 𝑖 (6.1) βˆ‘π‘› (πœ€π‘–π‘Ÿπ‘Ÿ) 𝑖=1 π‘Ž 𝑖 where 𝑖 is the cycle number, βˆ‘π‘€ (πœ€π‘–π‘Ÿπ‘Ÿ) is the accumulation of irreversible strain after π‘š 𝑖=1 π‘Ž 𝑖 cycles, and βˆ‘π‘› (πœ€π‘–π‘Ÿπ‘Ÿ) is the total cumulative irreversible strain during the entire multi-level 𝑖=1 π‘Ž 𝑖 systematic cyclic loading test. Rock deformability and its failure mechanism are closely related to energy dissipation. Therefore, the energy trends during the rock deformation process can reflect the rock damage mechanism (Zhang et al. 2019b). As shown in Fig. 6.10, a part of the total work done on the unit volume of a specimen (π‘ˆ ) by the external force during a loading-unloading cycle is stored 𝑑 in the specimen as elastic energy (π‘ˆ ); the remaining is released as dissipated energy (π‘ˆ ) due 𝑒 𝑑 to plastic deformation and rock damage. Because of the complexity in energy conversion during rock deformation and failure, subtle energies (thermal energy, acoustic emission energy, kinetic energy, etc.) are usually ignored to simplify the energy equation as follows (Zhou et al. 2019): π‘ˆ = π‘ˆ +π‘ˆ (6.2) 𝑑 𝑒 𝑑 πœ€" π‘ˆ 𝑑 = ∫ 0 𝜎 π‘Ž π‘‘πœ€ π‘Ž πœ€" (6.3) π‘ˆ = ∫ 𝜎 π‘‘πœ€ 𝑒 πœ€β€² π‘Ž π‘Ž {π‘ˆ = π‘ˆ βˆ’π‘ˆ 𝑑 𝑑 𝑒 178
ADE
Fig. 6.11 summarises the evolution of the damage variable (𝐷), elastic modulus (𝐸), and energy dissipation ratio (𝐾 = π‘ˆ /π‘ˆ ) as damage parameters for specimen TL6. A similar trend was 𝑑 𝑑 observed for the other tested specimens. As demonstrated in Fig. 6.11, the total behaviour of damage parameters under multi-level systematic cyclic loading conditions can be divided into four stages. In stage I, the damage variable (𝐷) increases slightly and is accompanied by the rapid increase in stiffness (𝐸) from 1.46 GPa to 2.23 GPa, corresponding to the closure of existing defects and expansion of the yield surface (Taheri and Tatsuoka 2015). Furthermore, the energy dissipation ratio (𝐾) decreases suddenly in this stage, which indicates that the elastic energy (π‘ˆ ) accumulates more rapidly than the dissipated energy (π‘ˆ ). Stage II, which is the 𝑒 𝑑 majority of the damage evolution process, shows a nearly unchanging behaviour for all three damage parameters 𝐷, 𝐸, and 𝐾. In this stage, although the specimen has experienced 400 cycles, no notable damage is incurred in the specimen. This stage can be interpreted as a balance between the two mechanisms of dilatant microcracking, which reduces the rock stiffness, and grain crushing and pore collapse, which improves the rock stiffness. This balanced state between two competing inelastic procedures results in a quasi-elastic behaviour of the damage parameters in such a way that the deformation seems elastic, and no more energy is dissipated in this stage. In stage III, during the transition to the second stress level via a monotonic loading, the elastic modulus first increases for several cycles. This increase may be related to the change in the strain rate from 2Γ—10-4/s to 0.02Γ—10-4/s for monotonic loading, which allows the existing microcracks and pores to be more compacted and ultimately results in a small stiffening (Peng et al. 2019). Then, the elastic modulus decreases gradually due to the dilatant cracking that degrades the axial stiffness and simultaneously allows more energy to be dissipated (see the trend of 𝐾 in Fig. 6.11). In stage IV, the specimen enters the post-peak region due to the coalescence of the microcracks and the generation of macrocracks through the specimen, and the degradation process of the specimen increases dramatically. According to Fig. 6.11, the energy dissipation ratio (𝐾) and damage variable (𝐷) increase rapidly in this stage, while the stiffness of the specimen decreases until the residual state is reached. 179
ADE
constant and very close to the maximum stress in each cycle. When transitioning to the higher stress levels using a monotonic loading, 𝜎 increases to reach a stationary state at each stress 𝑐𝑑 level. The results presented in Fig. 6.9 show that by applying 400 cycles at each stress level, the closed microvoids and micropores are not re-opened during pre-peak cyclic loading until the cyclic loading damages the rock at the last stress level. Thus, when the cyclic loading stress level is not high enough to cause the specimen to fail, the specimen does not switch from a compaction-dominated state to a dilatancy-dominated state but instead acts as an elastic material. According to Fig. 6.9a, specimen TL6 shows dilatant behaviour in the pre-peak region, in the second cyclic loading stage, by a sudden drop in 𝜎 due to the re-opening of 𝑐𝑑 closed cracks and the generation of new cracks. Degradation of 𝜎 continues in the post-peak 𝑐𝑑 region, followed by strength degradation until the specimen starts to show a residual strength state where 𝜎 increases to reach a stable condition. For specimens TL7 (Fig. 6.9b) and TL8 𝑐𝑑 (Fig. 6.9c), the drop in 𝜎 occurs very close to and at the failure point, respectively. This, in 𝑐𝑑 turn, causes a sudden release of stored elastic strain energy in a self-sustaining manner. 6.4. Strength Hardening Behaviour As mentioned earlier, in the cyclic loading tests, an increase in the peak strength of specimens TL7 and TL8 was observed with the increase in fatigue life in the pre-peak region. The discrepancy among specimens may partially contribute to this trend in the results. Considering the previous findings (Burdine 1963; Singh 1989; Ma et al. 2013; Taheri et al. 2017) and the results of cyclic loading tests in this study, the authors believe that the increase in the peak strength of specimens TL7 and TL8 is due to not only this discrepancy but also the cyclic loading. This phenomenon should be investigated in future studies by undertaking more specific cyclic loading tests. The hardening behaviour, however, is discussed briefly below. As discussed in section 6.3.2 and shown in Fig. 6.11, during pre-peak systematic cyclic loading, when the stress level is not high enough to cause the specimen to fail due to fatigue, a quasi- elastic behaviour dominates the damage evolution process. In this stage, some mesoscopic elements with lower strength and stiffness may reach their maximum load-bearing capacity, and the weak bonding between the grains breaks, producing fine materials. However, as the stress level is not close to the failure point, due to the slippage and dislocation of the produced fine materials, the existing microfissures and pores are filled during cyclic loading. This may result in more compaction of the specimen and, consequently, strength hardening. This behaviour can also be confirmed by the variation in crack damage threshold stress (𝜎 ) during 𝑐𝑑 181
ADE
cyclic loading (see Fig. 6.9). As explained in section 3.3, specimen TL8, which experienced more loading and unloading cycles in the pre-peak region than the other specimens did, is mostly in the compaction-dominated stage; dilation occurs at the failure point, followed by the sudden decrease in 𝜎 . This, in turn, resulted in the strength improvement of specimen TL8. 𝑐𝑑 However, for specimen TL6 with a shorter fatigue life, dilation occurred earlier in the pre-peak region. The process of rock compaction and porosity reduction in highly porous rock material may be similar to the mechanism explained by Baud et al. (2017). Fig. 6.12 shows the backscattered SEM images of a highly porous limestone in intact and deformed conditions under the same confining pressure of 9 MPa at different axial strain levels. As shown in this figure, when the intact specimen (Fig. 6.12a) deforms to 14% strain, microcracks are created in the calcite grains, and most of the fossil shells are broken and pulverised, while the quartz grains largely remain intact (Fig. 6.12b). With the further deformation of the specimen to 27% strain (Fig. 6.12c), the majority of the calcite grains are broken, and all of the fossil shells are pulverized, resulting in the existing pores being filled and the creation of compacted zones through the specimen. This grain packing is more evident in Fig. 6.12d, at a larger scale. The stress may concentrate more around the compacted areas, which behave elastically during loading and may contribute to the specimens exhibiting more brittle failure. Figure 6.12 Backscattered SEM images of a porous limestone in a intact and triaxial compression conditions for b 14% and c, d 27% axial strain (modified from Baud et al. (2017)) 182
ADE
6.5. Conclusions An innovative testing methodology considering two criteria was proposed in this study to describe the post-peak behaviour of rocks subjected to systematic cyclic loading. Regarding this, the Tuffeau limestone was selected to evaluate the capability of the proposed testing method in capturing the full stress-strain response of soft rocks. After obtaining the optimum values for the loading rate (π‘‘πœ€ /𝑑𝑑) and π΄π‘šπ‘.(πœ€ ) during a trial procedure, three main multi- 𝑙 𝑙 level systematic cyclic loading tests were conducted on Tuffeau limestone specimens using the proposed damage-controlled test method. The evolution of different parameters, including the peak strength, damage variable, elastic modulus and crack damage threshold stress, was evaluated comprehensively with the results of the conducted cyclic loading tests. The following conclusions were drawn from this study: 1. The proposed double-criteria damage-controlled testing method was successful in capturing the class II post-peak behaviour of Tuffeau limestone subjected to multi- level systematic cyclic loading. This testing method can provide new insights regarding the damage evolution of rocks in the post-peak region under systematic cyclic loading conditions, which was not previously achievable. The test method was successfully performed on Tuffeau limestone, which is a soft rock. The application of the method still needs to be examined on stronger rock types. 2. The whole process of cyclic loading tests conducted in this study can be summarised into several stages: a) The rock specimen initially stiffens and shows elastic behaviour due to the initial compaction, which is accompanied by the reduction in the energy dissipation. b) Due to a balance between the grain-crushing and pore collapse processes during compaction, a quasi-elastic behaviour dominates the whole test. c) The stiffness of the specimen decreases gradually due to dilatant microcracking, which dissipates more energy. d) With the generation and coalescence of microcracks, the rocks transition from a dilatant state, characterised by the rapid increase in damage and energy dissipation, and stiffness reduction. 3. The evolution of the crack damage threshold stress (𝜎 ) during cyclic loading showed 𝑐𝑑 that the specimens do not switch from a compaction-dominated to a dilatancy- dominated state when the cyclic loading stress level is not high enough to cause the specimen to fail. This results in a constant 𝜎 that is very close to the unloading stress 𝑐𝑑 in each cycle. 183
ADE
4. An increase in strength with an increase in fatigue life was observed for the highly porous Tuffeau limestone. According to the variation in the damage parameters, stiffness and crack damage threshold stress during the systematic cyclic loading tests, this hardening behaviour can be due to the further compaction of a rock specimen with increasing number of cycles in the pre-peak region. Indeed, the weak bonding between the grains may break down during cycling loading, and the fine materials produced in this process may fill the existing micropores and microfissures, which can result in a porosity reduction and hardening behaviour. Acknowledgments The authors would like to thank the laboratory staff, in particular, Simon Golding and Dale Hodson, for their aids in conducting the tests. Funding The first author acknowledges the University of Adelaide for providing the research fund (Beacon of Enlightenment PhD Scholarship) to conduct this study. References Akinbinu VA (2016) Class I and Class II rocks: implication of self-sustaining fracturing in brittle compression. Geotechnical and Geological Engineering 34(3):877–887 Al-Mukhtar M, Beck K (2006) Physical-mechanical characterisation of hydraulic and non- hydraulic lime based mortars for a French porous limestone. arXiv Prepr physics/0609108 Attewell PB, Sandford MR (1974) Intrinsic shear strength of a brittle, anisotropic rockβ€”I: experimental and mechanical interpretation. In: International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts. Elsevier, 11(11):423–430 Bagde MN, PetroΕ‘ V (2005) Fatigue properties of intact sandstone samples subjected to dynamic uniaxial cyclical loading. International Journal of Rock Mechanics and Mining Sciences 42(2):237–250 Baud P, Schubnel A, Heap M, Rolland A (2017) Inelastic compaction in high‐porosity limestone monitored using acoustic emissions. Journal of Geophysical Research: Solid Earth 122:9910–9989 184
ADE
Beck K, Al-Mukhtar M (2014) Cyclic wetting–drying ageing test and patina formation on tuffeau limestone. Environmental Earth Sciences 71(5):2361–2372 Burdine NT (1963) Rock Failure Under Dynamic Loading Conditions. Society of Petroleum Engineers Journal 3(1):1–8 Cardani G, Meda A (2004) Marble behaviour under monotonic and cyclic loading in tension. Construction and Building materials 18(6):419–424 Cerfontaine B, Collin F (2018) Cyclic and fatigue behaviour of rock materials: review, interpretation and research perspectives. Rock Mechanics and Rock Engineering 51(2):391–414 Cho SH, Ogata Y, Kaneko K (2003) Strain-rate dependency of the dynamic tensile strength of rock. International Journal of Rock Mechanics and Mining Sciences 40(5):763–777 Eberhardt E, Stead D, Stimpson B (1999) Quantifying progressive pre-peak brittle fracture damage in rock during uniaxial compression. International Journal of Rock Mechanics and Mining Sciences 36(3):361–380 Fairhurst CE, Hudson JA (1999) Draft ISRM suggested method for the complete stress-strain curve for intact rock in uniaxial compression. International Journal of Rock Mechanics and Mining Sciences 36(3):279–289 Ghamgosar M, Erarslan N (2016) Experimental and numerical studies on development of fracture process zone (FPZ) in rocks under cyclic and static loadings. Rock Mechanics and Rock Engineering 49(3):893–908 Guo H, Ji M, Zhang Y, Zhang M (2018) Study of mechanical property of rock under uniaxial cyclic loading and unloading. Advances in Civil Engineering Heap MJ, Faulkner DR, Meredith PG, Vinciguerra S (2010) Elastic moduli evolution and accompanying stress changes with increasing crack damage: implications for stress changes around fault zones and volcanoes during deformation. Geophysical Journal International 183(1):225–236 Hudson JA, Brown ET, Fairhurst C (1971) Optimising the control of rock failure in servo- controlled laboratory tests. Rock Mechanics 3:217–224 Li T, Pei X, Wang D, Huang R, Tang H (2019) Nonlinear behavior and damage model for 185
ADE
fractured rock under cyclic loading based on energy dissipation principle. Engineering Fracture Mechanics 206:330–341 Liu J, Xie H, Hou Z, Hou Z, Yang C, Chen L (2014) Damage evolution of rock salt under cyclic loading in unixial tests. Acta Geotechnica 9(1):153–160 Ma L, Liu X, Wang M, Xu H, Hua R, Fan P, Jiang S, Wang G, Yi Q (2013) Experimental investigation of the mechanical properties of rock salt under triaxial cyclic loading. International Journal of Rock Mechanics and Mining Sciences 34–41 Munoz H, Taheri A (2017a) Local damage and progressive localisation in porous sandstone during cyclic loading. Rock Mechanics and Rock Engineering 50(12):3253–3259 Munoz H, Taheri A (2017b) Specimen aspect ratio and progressive field strain development of sandstone under uniaxial compression by three-dimensional digital image correlation. Journal of Rock Mechanics and Geotechnical Engineering 9(4):599–610 Munoz H, Taheri A (2019) Postpeak deformability parameters of localised and nonlocalised damage zones of rocks under cyclic loading. Geotechnical Testing Journal 42(6):1663– 1684 Munoz H, Taheri A, Chanda EK (2016a) Rock Drilling Performance Evaluation by an Energy Dissipation Based Rock Brittleness Index. Rock Mechanics and Rock Engineering 49(8):3343–3355 Munoz H, Taheri A, Chanda EK (2016b) Pre-peak and post-peak rock strain characteristics during uniaxial compression by 3D digital image correlation. Rock Mechanics and Rock Engineering 49(7):2541–2554 Munoz H, Taheri A, Chanda EK (2016c) Fracture Energy-Based Brittleness Index Development and Brittleness Quantification by Pre-peak Strength Parameters in Rock Uniaxial Compression. Rock Mechanics and Rock Engineering 49(12):4587–4606 Peng K, Zhou J, Zou Q, Yan F (2019) Deformation characteristics of sandstones during cyclic loading and unloading with varying lower limits of stress under different confining pressures. International Journal of Fatigue 127:82–100 Singh SK (1989) Fatigue and strain hardening behaviour of graywacke from the flagstaff formation, New South Wales. Engineering Geology 26(2):171–179 186
ADE
Taheri A, Tatsuoka F (2012) Stress–strain relations of cement-mixed gravelly soil from multiple-step triaxial compression test results. Soils and Foundations 52(4):748–766 Taheri A, Tatsuoka F (2015) Small- and large-strain behaviour of a cement-treated soil during various loading histories and testing conditions. Acta Geotechnica 10(1):131–155 Taheri A, Royle A, Yang Z, Zhao Y (2016) Study on variations of peak strength of a sandstone during cyclic loading. Geomechanics and Geophysics for Geo-Energy and Geo-Resources 2(1):1–10 Taheri A, Hamzah N, Dai Q (2017) Degradation and improvement of mechanical properties of rock under triaxial compressive cyclic loading. Japanese Geotechnical Society Special Publication 5:71–78 Taheri A, Zhang Y, Munoz H (2020) Performance of rock crack stress thresholds determination criteria and investigating strength and confining pressure effects. Construction and Building Materials 243:118263 Wang S, Xu W, Sun M, Wang W (2019) Experimental investigation of the mechanical properties of fine-grained sandstone in the triaxial cyclic loading test. Environmental Earth Sciences 78(14):416 Wawersik WR, Fairhurst CH (1970) A study of brittle rock fracture in laboratory compression experiments. In: International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts. Elsevier, 7(5):561–575 Xiao J-Q, Ding D-X, Jiang F-L, Xu G (2010) Fatigue damage variable and evolution of rock subjected to cyclic loading. International Journal of Rock Mechanics and Mining Sciences 47:461–468 Xiao J-Q, Ding D-X, Xu G, Jiang F-L (2009) Inverted S-shaped model for nonlinear fatigue damage of rock. International Journal of Rock Mechanics and Mining Sciences 46(3):643–648 Xuefeng X, Linming D, Caiping L, Zhang Y (2010) Frequency spectrum analysis on micro- seismic signal of rock bursts induced by dynamic disturbance. Mining Science and Technology (China) 20(5):682–685 Zhang J, Deng H, Taheri A, Ke B, Liu C (2019a) Deterioration and strain energy development of sandstones under quasi-static and dynamic loading after freeze-thaw cycles. Cold 187
ADE
Chapter 7 Failure Behaviour of a Sandstone Subjected to the Systematic Cyclic Loading: Insights from the Double-Criteria Damage-Controlled Test Method Abstract The post-peak behaviour of rocks subjected to cyclic loading is very significant to appraise the long-term stability of underground excavations. However, an appropriate testing methodology is required to control the damage induced by the cyclic loading during the failure process. In this study, the post-failure behaviour of Gosford sandstone subjected to the systematic cyclic loading at different stress levels was investigated using the double-criteria damage-controlled testing methodology, and the complete stress-strain relations were captured successfully. The results showed that there exists a fatigue threshold stress in the range of 86-87.5% of the average monotonic strength in which when the cyclic loading stress is below this threshold, no failure occurred for a large number of cycles and in turn, the peak strength improved up to 8%. Also, the variation of the energy dissipation ratio, rock stiffness and acoustic emission hits for hardening tests showed that cyclic loading in the pre-peak regime creates no critical damage in the specimen, and a quasi-elastic behaviour dominates the damage evolution. The post-failure instability of such tests was similar to those obtained for monotonic tests. On the other hand, by exceeding the fatigue threshold stress, the brittleness of the specimens increased with an increase in the applied stress level, and class II behaviour prevailed over total post-peak behaviour. A loose-dense-loose behaviour with different extents was also observed in the post- peak regime of all fatigue cyclic loading tests. This was manifested then as a secondary inverted S-shaped damage behaviour by the variation of the cumulative irreversible axial and cumulative irreversible lateral strains with the post-peak cycle number. Furthermore, it was confirmed that the damage per cycle in the post-peak regime decreases exponentially with an increase in the applied stress level. Keywords: Pre-peak and post-peak behaviour, Systematic cyclic loading, Brittleness, Hardening, Fatigue, Damage evolution 190
ADE
List of Symbols 𝐸 Tangent Young’s modulus πœ€π‘–π‘Ÿπ‘Ÿ Irreversible axial strain π‘‘π‘Žπ‘› π‘Ž 𝜈 Poisson’s ratio πœ€π‘–π‘Ÿπ‘Ÿ Irreversible lateral strain 𝑙 𝜎 Major principal stress Ξ£πœ€π‘–π‘Ÿπ‘Ÿ Cumulative irreversible axial strain 1 π‘Ž 𝜎 Axial stress Ξ£πœ€π‘–π‘Ÿπ‘Ÿ Cumulative irreversible lateral strain π‘Ž 𝑙 𝜎 Indicator stress π‘ˆ Elastic energy at peak stress 𝑖 𝑒 𝜎 Axial peak stress π‘ˆπ‘– Elastic energy of cycle 𝑖 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ 𝑒 𝜎 Average monotonic strength π‘ˆπ‘– Dissipated energy of cycle 𝑖 π‘š 𝑑 𝜎 /𝜎 Applied stress level π‘ˆ Pre-peak dissipated energy π‘Ž π‘š π‘π‘Ÿπ‘’ 𝜎 /𝜎 Strength hardening ratio π‘ˆ Post-peak dissipated energy β„Ž π‘š π‘π‘œπ‘ π‘‘ 𝜎 /𝜎 Crack initiation stress ratio π‘ˆ Total fracture energy 𝑐𝑖 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ 𝑑 𝜎 /𝜎 Crack damage stress ratio π΄π‘šπ‘.(𝜎 ) Loading amplitude 𝑐𝑑 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ π‘Ž πœ€ Axial strain π΄π‘šπ‘.(πœ€ ) Lateral strain amplitude π‘Ž 𝑙 πœ€ Lateral strain 𝑛 Cycle number 𝑙 π‘‘πœ€ /𝑑𝑑 Lateral strain rate 𝑁 Total number of cycles 𝑙 π‘‘π‘œπ‘‘π‘Žπ‘™ πœ€ Axial strain at peak stress 𝑁 Number of cycles after failure point π‘Žβˆ’π‘π‘’π‘Žπ‘˜ π‘Žπ‘“π‘‘π‘’π‘Ÿ πœ€ Lateral strain at peak stress 𝐡𝐼 Brittleness index π‘™βˆ’π‘π‘’π‘Žπ‘˜ πœ€ Volumetric strain at peak stress 𝐷 Damage variable π‘£βˆ’π‘π‘’π‘Žπ‘˜ πœ€ Axial strain at the final cycle 𝑀 Post-peak modulus π‘Žβˆ’π‘“ 7.1. Introduction A high-complex stress state usually is created around deep-buried tunnels and caverns due to disturbances induced by different sources as displayed in Fig. 7.1. This stress state may affect mechanical rock properties and in turn, cause some specific failure phenomena such as slabbing/spalling, strainburst and zonal disintegration significantly different from those in shallow conditions (Gong et al. 2012; Shirani Faradonbeh and Taheri 2019). According to Martin and Chandler (1994) and Martin (1997), the surrounding rocks in underground excavations may experience load-and-deformation response to a different extent during operation, and rock may be exposed to cyclic loading. In particular, they argued that in remote to nearby excavation regions, rock may experience failure (i.e. the applied stress level exceeds the peak strength), damage (i.e. the applied stress is below the peak strength) or disturbance (i.e. different stress is applied due to the redistribution of the in-situ stresses) or the rock may remain undisturbed. From this viewpoint, the rock cyclic load-deformation response may take 191
ADE
place in the pre-peak or post-peak regime (Munoz and Taheri 2019). For instance, as depicted in Fig. 7.1, a pillar may experience cyclic loading due to blasting operation or other seismic activities beyond the limit in uniaxial conditions. Under such loading conditions, rock materials may still keep some loadings even in the post-failure regime. Therefore, the investigation of the pre-peak and post-peak behaviour of rocks is of paramount significance to understand more about the fracturing mechanism, resilient design and long-term stability assessment of the various rock engineering structures subjected to seismic disturbances. Experimental research on the influence of cyclic loading parameters on the damage evolution and rock strength and deformation parameters has a long tradition. These studies have been conducted under different loading histories and loading conditions such as uniaxial and triaxial compression tests (Heap and Faulkner 2008; Heap et al. 2009; Liu et al. 2018), indirect tensile tests (Erarslan et al. 2014; Wang et al. 2016), flexural tests (Cattaneo and Labuz 2001; Cardani and Meda 2004) and freeze-thaw tests (Liu et al. 2015; Zhang et al. 2019). A comprehensive review of the rock fatigue studies can be found in Cerfontaine and Collin (2018). The majority of prior rock fatigue studies have emphasised strength weakening of rocks due to incurring permanent deformations during cyclic loading (Haimson 1978; Fuenkajorn and Phueakphum 2010). However, very few studies have reported the strength improvement when the stress level that cyclic loading is applied is low enough to prevent failure (Singh 1989; Ma et al. 2013; Taheri et al. 2017). In prior studies, the process of damage evolution and the failure mechanism of rocks subjected to different cyclic loading histories have been investigated based on the measured stress-strain relations (Cerfontaine and Collin 2018). Indeed, the complete stress- strain relation of rocks (i.e. the pre-peak and the post-peak regimes) is considered as a prominent tool in rock engineering to describe strain energy evolution as well as for rock brittleness determination (Munoz et al. 2016a; Shirani Faradonbeh et al. 2020). According to Wawersik and Fairhurst (1970), the post-peak behaviour of rocks under quasi-static compression can be distinguished into two classes: a) class I which is characterised by the negative post-peak modulus (i.e. 𝑀 = π‘‘πœŽ/π‘‘πœ€ < 0) representing the gradual strength degradation of rock specimen and the need for extra energy and b) class II having a positive post-peak modulus represents the self-sustaining failure with strain recovery and release of excess elastic strain energy. The proper measurement of the complete stress-strain response of rocks significantly depends on the stiffness of the loading system, the applied load controlling technique throughout the test as well as rock brittleness (Wawersik and Fairhurst 1970; Munoz and Taheri 2019). 192
ADE
Shirani Faradonbeh et al. (2020) categorised the cyclic loading methods based on the loading histories and load control variables into two main groups of systematic cyclic loading (single- level or multi-level) (Figs. 7.2a and b) and damage-controlled cyclic loading (load-based or displacement-based) (Figs. 7.2c and d). Systematic cyclic loading can be conducted under load- controlled or displacement-controlled loading conditions. In both loading conditions, a sudden failure occurs during cyclic loading as a constant axial load amplitude, π΄π‘šπ‘.(𝜎 ), should be π‘Ž achieved during each loading cycle (e.g. Ma et al. 2013; Li et al. 2019). Similarly, in the load- based damage-controlled cyclic loading tests, as the specimen is forced to reach a prescribed stress level, it may experience an unexpected failure, and the post-peak behaviour cannot be captured (e.g. Heap et al. 2010; Guo et al. 2018). Regarding the displacement-based damage- controlled cyclic loading tests, as the post-peak behaviour of rocks in uniaxial compression is either class II or a combination of class I and class II (Munoz et al. 2016a), the post-peak response cannot be adequately captured by the axial displacement feedback signal (e.g. Wang et al. 2019). The lateral displacement, on the other hand, has been identified as an appropriate variable to control the amount of damage in the post-peak regime (Munoz and Taheri 2019). To our knowledge, no prior studies have examined the influence of systematic cyclic loading at different stress levels on the post-peak behaviour of rocks. This is due to the difficulties in controlling the axial load when a constant load amplitude should be achieved in every cycle in a systematic cyclic loading test. Also, if a prescribed lateral strain is considered to control the damage in a damage-controlled test, the axial load is reversed when a certain amount of lateral strain occurs, and therefore, the systematic cyclic loading cannot be conducted anymore in the pre-peak regime. However, as mentioned earlier, some mining and civil structures (e.g. mining pillars and bridge columns) may experience systematic cyclic loading at different fractions of their average peak strength. Under such loading conditions, the rocks may exhibit different behaviours in the post-peak regime. An appropriate experimental methodology is, therefore, required for measuring the post-peak behaviour of rocks subjected to systematic cyclic loading histories properly. As demonstrated in Fig. 7.2, a novel cyclic test method by combining the single-level systematic cyclic loading and lateral displacement-based damage-controlled cyclic loading is proposed in this study to control both the damage and the cyclic loading rate. Then, several systematic cyclic tests were conducted in uniaxial compression at different stress levels using the proposed test method. Based on the obtained complete stress-strain relations, the influence of systematic cyclic loading on both the pre-peak and the post-peak behaviours was evaluated comprehensively, and the results were discussed in detail. 193
ADE
Figure 7.2 Classification of cyclic loading tests, a single-level systematic cyclic loading path, b multilevel systematic cyclic loading path, c load-based damage controlled cyclic loading path and d displacement-based damage controlled cyclic loading path, Amp. (𝜎 ) refers to π‘Ž loading amplitude, Amp. (πœ€ ) refers to lateral strain amplitude, and * can be conducted either 𝐿 in axial or lateral displacement-controlled mode, modified from Shirani Faradonbeh et al. (2020) 7.2. Specimen Preparation and Experimental Set-Up The Gosford sandstone as a medium-grained (0.2-0.3 mm), poorly cemented, immature quartz sandstone containing 20-30% feldspar and clay minerals with the serrate connection between quartz grains (Sufian and Russell 2013) was used in this study for conducting the experimental tests. According to the X-ray computed tomography scans conducted by Sufian and Russell (2013), the total porosity of this sandstone is about 18%. A total of 23 cylindrical specimens having a constant aspect ratio of 2.4 (i.e. 42 mm diameter and 100 mm length) were all cored from the same rectangular block and in the same direction and prepared according to the ISRM suggested method (Fairhurst and Hudson 1999). In this study, all the experiments were performed in dry condition. To do so, the rock specimens were dried in the room temperature before conducting the tests. The average dry density of the specimens was approximately 2204.26 kg/m3. Rock monotonic strength should be determined before undertaking systematic cyclic loading tests at different stress levels (𝜎 /𝜎 ). To do so, six uniaxial compression tests π‘Ž π‘š were performed following the lateral strain-controlled loading method. An MTS close-looped servo-controlled hydraulic compressive system having the maximum loading capacity of 300 kN (see Fig. 7.3) was used to undertake the monotonic and cyclic loading tests. As stated earlier, the axial load-controlled and axial strain-controlled loading techniques cannot capture the post-peak behaviour of rocks, as rocks usually show a combination of class I and class II behaviour in the post-peak regime (Munoz et al. 2016b). Therefore, as depicted in Fig. 7.4a, a constant lateral strain rate (π‘‘πœ€ /𝑑 ) of 0.02Γ—10-4/s was utilised during the uniaxial compression 𝑙 𝑑 tests to control the axial load both in the pre-peak and the post-peak regimes. This strain rate provides a static to quasi-static loading conditions (Wawersik and Fairhurst 1970; Munoz et al. 2016b). Axial load and axial and lateral displacements were recorded in real-time, respectively using the load cell, a pair of LVDTs externally mounted between the loading platens and a direct- contact chain extensometer wrapped around the specimens (see Fig. 7.3). Due to the large- 195
ADE
strain behaviour of rocks in the post-peak regime, the local strain measurement tools such as strain gauges are not effective. To characterise the post-peak instability of rocks in terms of brittleness, the complete stress-strain curves of rocks are required, and therefore, external LVDTs were used to measure the large-strain properties. LVDTs measure the deformation between loading platens; therefore, the deformation of the loading system is not included in the measurement. Still, the strain data may not be precise due to well-known bedding error (Taheri and Tani 2008). The bedding error refers to the additional deformations measured by LVDTs due to crushing the irregularities/asperities at the end faces of the specimens before the specimen deforms as well as the poor fitting of the specimen to the loading platens. This error is minimised in this study by carefully and smoothly grinding the ends of the specimen following the ISRM standard (Fairhurst and Hudson 1999). Besides, since the focus of this study is complete stress-strain behaviour, this error is deemed negligible in large strain stress- strain properties. The acoustic emission (AE) technique, as a passive non-destructive monitoring technology, was also employed in this study to measure the real-time formation and growth of local micro- cracks throughout the specimen (internal damage) during cyclic loading (Lockner 1993; Bruning et al. 2018). For this aim, as depicted in Fig. 7.3, two miniature PICO sensors were attached to the specimens, and the recorded acoustic signals by these sensors were amplified using a pre-amplifier (type 2/4/6) set to 60 dB of gain. The AE recordings were carried out using the Express-8 data acquisition card with the sampling rate of 2 MSPS (million samples per second). To ensure that mechanical noises induced by the loading system are not recorded during the tests, the AE threshold amplitude was changed from 20 dB to 60 dB, and it was found that after 45 dB amplitude, no additional noises are recorded. Therefore, this value was set as the AE threshold. The stress-strain curves obtained from the conducted uniaxial compressive tests and their relevant mechanical properties can be found in Fig. 4b, and Table 7.1, respectively. In Table 7.1, the tangent Young’s modulus (𝐸 ) and Poisson’s ratio (𝜈) values were determined at π‘‘π‘Žπ‘› 50% of the axial peak stress (𝜎 ) for each monotonic test. The crack initiation stress (𝜎 ) π‘Žβˆ’π‘π‘’π‘Žπ‘˜ 𝑐𝑖 and crack damage stress (𝜎 ) thresholds were also determined using the methods explained in 𝑐𝑑 Taheri et al. (2020). According to Fig. 7.4b, the stress-strain curves for all compression tests show almost a similar behavioural trend both in the pre-peak and the post-peak regimes. In the pre-peak regime, as listed in Table 7.1, the deformation parameters of axial (πœ€ ), lateral π‘Žβˆ’π‘π‘’π‘Žπ‘˜ (πœ€ ) and volumetric strains (πœ€ ) at peak stress points, 𝐸 , 𝜈, crack initiation stress π‘™βˆ’π‘π‘’π‘Žπ‘˜ π‘£βˆ’π‘π‘’π‘Žπ‘˜ π‘‘π‘Žπ‘› 196
ADE
ratio (𝜎 /𝜎 ) and crack damage stress ratio (𝜎 /𝜎 ) are approximately similar, 𝑐𝑖 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ 𝑐𝑑 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ which indicates a small discreteness of the tested specimens. As such, in the post-failure regime, the sudden drops and recoveries of the load-bearing capacity can be observed for all specimens which can be associated with the shear strain localisation, grain interlocking in between the sides of the generated macrocracks (Jansen and Shah 1997; Vasconcelos et al. 2009) as well as the automatic adjustment of applied load by the testing machine upon damage extension. The post-peak regime of rocks under uniaxial compressive loading demonstrates a combined class I-II behaviour, which is consistent with the prior study conducted by Munoz et al. (2016b). As listed in Table 7.1, the monotonic compressive strength (𝜎 ) of the tested π‘Žβˆ’π‘π‘’π‘Žπ‘˜ Gosford sandstone specimens varied between 45.76 MPa and 49.89 MPa with an average value of 48.15 MPa. This average monotonic strength was utilised in the following to define the stress levels where the systematic cyclic loading tests should be commenced. Table 7.1 The results of uniaxial compressive tests for Gosford sandstone specimens Test No. 𝜎 𝐸 𝜈 Strains at the peak stress point 𝜎 /𝜎 𝜎 /𝜎 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ π‘‘π‘Žπ‘› 𝑐𝑖 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ 𝑐𝑑 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ (MPa) (GPa) πœ€ π‘Žβˆ’π‘π‘’π‘Žπ‘˜ πœ€ π‘™βˆ’π‘π‘’π‘Žπ‘˜ πœ€ π‘£βˆ’π‘π‘’π‘Žπ‘˜ (%) (%) (Γ—10-4) (Γ—10-4) (Γ—10-4) GS-1 48.05 13.30 0.15 54.17 -38.35 -22.54 29.65 58.27 GS-2 49.54 13.43 0.12 52.18 -36.84 -21.51 30.60 58.67 GS-3 47.35 13.42 0.13 52.66 -39.10 -25.55 27.00 55.57 GS-4 45.76 12.97 0.15 51.39 -38.56 -25.73 25.80 55.96 GS-5 49.89 13.15 0.14 53.00 -36.97 -20.95 27.71 57.92 GS-6 48.29 14.14 0.15 50.17 -34.11 -18.05 26.94 52.70 Average 48.15 13.40 0.14 52.26 -37.32 -22.39 27.95 56.51 SD 1.51 0.40 0.01 1.38 1.81 2.93 1.82 2.25 SD standard deviation 197
ADE
(b) GS-1 50 GS-2 ) a GS-3 P M 40 GS-4 ( GS-5 a GS-6 , s 30 s e r t s l a 20 i x A 10 0 0 20 40 60 80 Axial strain, e (Β΄10-4) a Figure 7.4 (Continued) 7.3. Systematic Cyclic Loading Tests As discussed earlier, the single-criterion load-based and displacement-based loading methods are not sufficient to control the axial load in the post-failure stage during the systematic cyclic loading tests, especially when rocks demonstrate self-sustained failure behaviour. In this study, to address this issue, a new testing method called β€œdouble-criteria damage-controlled test method” (Shirani Faradonbeh et al. 2020) was employed. As demonstrated in Fig. 7.2, this test method is a combination of single-level systematic cyclic loading and damage-controlled cyclic loading lateral displacement-controlled loading method. In this regard, the MTS servo- controlled testing machine was programmed so that the hydraulic system was allowed to be adjusted continuously, automatically and rapidly according to the received feedback signals from both chain extensometer and load cell during a closed-loop procedure. The testing procedure can be summarised into the following four stages: 1. Load the specimen monotonically (π‘‘πœ€ /𝑑𝑑 = 0.02Γ—10-4) until the pre-defined stress 𝑙 level (𝜎 /𝜎 ), and then, unload it at the same loading rate until 𝜎 = 0.07 MPa, ensuring π‘Ž π‘š π‘Ž the specimen is always in complete contact with the loading platens. 2. Reload the specimen under a constant lateral strain rate of 3Γ—10-4/s until one of the two following criteria is met during loading: 199 s
ADE
a) the pre-defined maximum axial stress level (𝜎 /𝜎 ) is reached; π‘Ž π‘š b) the pre-defined maximum lateral strain amplitude, π΄π‘šπ‘.(πœ€ )= 32Γ—10-4 is 𝑙 reached; 3. Reverse the axial load to 𝜎 = 0.07 MPa, and repeat steps 1 and 2 until 1500 loading π‘Ž and unloading cycles are completed. 4. If the specimen did no fail during 1500 cycles, apply a monotonic loading (π‘‘πœ€ /𝑑𝑑 = 𝑙 0.02Γ—10-4) until complete failure occurs. In this study, π΄π‘šπ‘.(πœ€ )= 32Γ—10-4 was determined based on the conducted monotonic tests and 𝑙 the measured lateral strain of the rocks at the failure point, πœ€ (see Table 7.1). As seen in π‘™βˆ’π‘π‘’π‘Žπ‘˜ Table 7.1, the average value of πœ€ for the tested specimens is -37.32Γ—10-4. Based on the π‘™βˆ’π‘π‘’π‘Žπ‘˜ conducted several trial tests, it was found that 32Γ—10-4 is an appropriate value for Gosford sandstone. By adopting this value, it was possible to avoid failing the sample in a single cycle while allowing the axial stress level to reach the pre-defined value to apply a systematic cyclic loading. Figs. 7.5a and b show two representative time histories of axial stress and lateral strain for Gosford sandstone specimens experiencing failure during cyclic loading and final monotonic loading. In Fig. 7.5a, the specimen was loaded monotonically (π‘‘πœ€ /𝑑𝑑= 0.02Γ—10- 𝑙 4/s) up to 85% of the average monotonic strength (𝜎 /𝜎 = 85%). Afterwards, the specimen π‘Ž π‘š was unloaded with the same rate, and then the systematic cyclic loading was initiated under the lateral strain rate of 3Γ—10-4/s. As shown in the inset figure, the cycles always met the first criterion (i.e. the maximum stress applied during a cycle remained constant) during the systematic cyclic loading and the π΄π‘šπ‘.(πœ€ ) was considerably lower than the pre-defined 𝑙 maximum amplitude for lateral strain (i.e. 32Γ—10-4) in each cycle. As during 1500 loading/unloading cycles, the π΄π‘šπ‘.(πœ€ ) did not exceed 32Γ—10-4, a monotonic 𝑙 loading was applied automatically to the specimen under the lateral strain rate of 0.02Γ—10-4/s until the specimen is completely failed. By doing so, the post-peak behaviour was captured successfully for further analyses. In Fig. 7.5b, the same cyclic loading procedure was applied to another specimen at a higher axial stress level (i.e. 𝜎 /𝜎 = 87.25%). In the pre-peak stage, π‘Ž π‘š the π΄π‘šπ‘.(πœ€ ) increased gradually by increasing the cycle number, while the stress level was 𝑙 kept constant, satisfying the first criterion. However, at the onset of the failure (where the axial stress begins to reduce), the π΄π‘šπ‘.(πœ€ ) reached the pre-defined value of 32Γ—10-4 (see the inset 𝑙 figure), and the second criterion was activated to control the cyclic loading. By transferring to the post-peak stage, and strength degradation, the subsequent cycles were carried out so that 200
ADE
7.4. Stress-Strain Relations In total, 17 single-level systematic cyclic loading tests (see Table 7.2) were carried out at different stress levels (𝜎 /𝜎 ) ranging from 80% to 96% of the average monotonic strength π‘Ž π‘š following the proposed double-criteria damage-controlled testing method. As listed in Table 7.1, the stable and unstable crack growths of rocks on average initiate at 𝜎 /𝜎 = 27.95% 𝑐𝑖 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ and 𝜎 /𝜎 = 56.51%, respectively. This, in other words, shows that the cyclic loading 𝑐𝑑 π‘Žβˆ’π‘π‘’π‘Žπ‘˜ tests have been conducted in the unstable crack propagation stage, beyond the elastic stress- strain behaviour. To evaluate the influence of cycle number on mechanical properties and post- peak behaviour, the specimens GS-8 and GS-9 were subjected to 5000 and 10000 cycles at 𝜎 /𝜎 =80% and GS-11 experienced 5000 cycles at the stress level of 𝜎 /𝜎 =85% before a π‘Ž π‘š π‘Ž π‘š monotonic loading. Otherwise, the samples experienced a maximum of 1500 cycles and then a post-monotonic loading should they did not fail during the cyclic loading. According to Beniawski (1967), to ensure fatigue failure of a rock specimen in a timely manner, the cyclic loading test should be conducted just before the onset of the unstable crack propagation stage within the range of 70-85% of the peak strength. A recent review conducted by Cerfontaine and Collin (2018) on rock fatigue studies reported that the rock fatigue threshold ranges from 0.75 to 0.9 of the average monotonic strength for one million loading and unloading cycles depending on rock type and loading conditions. However, in this study, due to test limitations, further cycles did not apply, and the results are valid in the range of 1500-10000 cycles. Based on the results presented in Table 7.2, it is hypothesised that there exists a threshold of 𝜎 /𝜎 π‘Ž π‘š which lies between 86% and 87.5 % that indicates the critical boundary of rock strength hardening and fatigue under cyclic loading. In this study, the cyclic loading tests which experienced the monotonic loading at the failure stage were named as hardening cyclic loading tests, while those which failed during cyclic loading at higher stress levels were named as fatigue cyclic loading tests. Figs. 7.6 and 7.7 show the typical stress-strain results for hardening and fatigue cyclic loading tests, respectively. In these figures, the total post-peak behaviour was highlighted by connecting the indicator stresses (𝜎, the maximum stress of each cycle). The πœ€π‘–π‘Ÿπ‘Ÿ and πœ€π‘–π‘Ÿπ‘Ÿ 𝑖 π‘Ž 𝑙 respectively, represent the irreversible axial strain and the irreversible lateral strain. The areas of interest (AOIs) shown in Figs. 7.6c and 7.7c illustrate the specific parts of the volumetric strain (πœ€ ) evolution which were enlarged in Figs. 7.6d and 7.7d, respectively. Figs. 7.6a and π‘£π‘œπ‘™ 7.7a show that the testing methodology was successful in capturing the complete stress-strain 202
ADE
curves of Gosford sandstone specimens subjected to the systematic cyclic loading. Furthermore, like the monotonic tests, a combined class I-II behaviour at different extents can be seen in the post-peak regime for both hardening and fatigue cyclic loading tests. Generally, the variation of hysteretic loops along with the axial strain (Figs. 7.6a and 7.7a), lateral strain (Figs. 7.6b and 7.7b) and volumetric strain (Figs. 7.6c and d and Figs. 7.7c and d) show that the rock specimens which fail during the cyclic loading significantly experience more irreversible strains in the pre-peak regime compared with hardening cyclic loading tests. Also, as shown in Fig. 7.7d, after a few cycles, the hysteretic loops for the fatigue cyclic loading tests switch rapidly from the compaction to dilation, and dilation continues until complete failure. Table 7.2 The results of the conducted systematic cyclic tests Test No. 𝜎 /𝜎 (%) 𝑁 𝑁 Hardening (H) or πœ€ πœ€ Peak strength π‘Ž π‘š π‘‘π‘œπ‘‘π‘Žπ‘™ π‘Žπ‘“π‘‘π‘’π‘Ÿ π‘Žβˆ’π‘“ π‘Žβˆ’π‘π‘’π‘Žπ‘˜ fatigue (F) test? (Γ—10-4) (Γ—10-4) increase (%) GS-7 80 1500 - H 45.80 53.56 0.53 GS-8 80 5000 - H 43.03 52.36 7.31 GS-9 80 10000 - H 48.94 55.98 0.05 GS-10 85 1500 - H 46.38 53.70 6.22 GS-11 85 5000 - H 48.93 54.29 2.17 GS-12 86 1500 - H 45.52 50.92 1.93 GS-13 87.50 1500 - H 47.72 55.04 7.82 GS-14 86.81 636 49 F - 56.15 - GS-15 87.23 49 26 F - 56.06 - GS-16 87.25 240 42 F - 54.78 - GS-17 89.65 40 28 F - 54.75 - GS-18 89.82 103 45 F - 53.12 - GS-19 91.76 145 97 F - 52.75 - GS-20 93 49 36 F - 54.37 - GS-21 93.65 280 260 F - 54.98 - GS-22 95 752 730 F - 54.46 - GS-23 96 474 318 F - 37.84 - 𝑁 total number of cycles, 𝑁 number of cycles after failure point, πœ€ axial strain at the peak of the π‘‘π‘œπ‘‘π‘Žπ‘™ π‘Žπ‘“π‘‘π‘’π‘Ÿ π‘Žβˆ’π‘“ final cycle, πœ€ axial strain at the failure point π‘Žβˆ’π‘π‘’π‘Žπ‘˜ 203
ADE
(c) (d) ) a 50 AOI 50 Dilation Compaction P ) M a P ( 40 M 40 a ( , a s s 30 30 e , r t s s s e l a 20 r t s 20 i x l A 10 a i x 10 A 0 0 -600 -400 -200 0 100 -100 -80 -60 -40 -20 0 20 100 Volumetric strain, e (Β΄10-4) Volumetric strain, e (Β΄10-4) vol vol Figure 7.7 (Continued) 7.5. Rock Behaviour During Hardening Cyclic Loading Tests 7.5.1. Damage Evolution in the Pre-Peak Regime In rock engineering applications, the rock deformation and failure processes are associated with the strain energy evolution (Li et al. 2019). The total inputted mechanical energy during a loading and unloading cycle is transformed into the stored elastic energy (π‘ˆπ‘–) and the dissipated 𝑒 energy (π‘ˆπ‘–) as shown schematically in Fig. 7.8a. The dissipated energy due to the irreversible 𝑑 deformations causes stiffness degradation and rock damage. In this study, the energy dissipation ratio (i.e. 𝐾 = π‘ˆ /π‘ˆ ) and tangent Young’s modulus (𝐸 ) were utilised to 𝑑 𝑒 π‘‘π‘Žπ‘› investigate progressive damage evolution in the pre-peak regime for hardening cyclic loading tests. Fig. 7.8b shows the representative results for specimen GS-10 at 𝜎 /𝜎 =85%. The other π‘Ž π‘š hardening cyclic loading tests conducted at different stress levels and with a different number of cycles also showed a similar trend. According to Fig. 7.8b, a two-stage damage evolution procedure can be identified for the hardening cyclic loading tests. In stage A, the 𝐸 increased π‘‘π‘Žπ‘› dramatically during initial cycles (approximately 21.94% compared with the average 𝐸 for π‘‘π‘Žπ‘› monotonic tests in Table 7.1), which can cause to specimen become stiffer. This behaviour can be relevant to the closure of existing defects . An increase of stiffness during initial loading cycle also has been reported by other researchers (Trippetta et al. 2013; Momeni et al. 2015; Taheri and Tatsuoka 2015; Taheri et al. 2016b). On the other hand, the energy dissipation ratio (𝐾) decreased suddenly in stage A, which contributes to the accumulation of elastic strain energy in rock specimen. In stage B, while it was expected to see stiffness degradation due to 205 s s
ADE
incurring irreversible deformations in the specimen by doing more cycles, 𝐸 and 𝐾 remained π‘‘π‘Žπ‘› fairly constant, and no considerable energy was dissipated until 1500 cycles were completed (i.e. a quasi-elastic behaviour). This quasi-elastic behaviour can be further investigated using AE results. Fig. 7.8c shows the typical time-history of AE hits recorded for the specimen GS-10. As shown in this figure, few AE hits are observed at the initial monotonic loading stage, which corresponds to seating, loading adjustment by the testing apparatus and the crack closure stage. However, in the second stage, almost no macrocrack (macro-damage) is generated throughout the specimen as a constant trend was observed for the cumulative AE hits during the 1500 cycles. In other words, at this stage, only small amounts of low amplitude AE hits (micro-damages) are generated (see Fig. 7.8c). During the final monotonic loading stage, new microcracks are generated and propagated throughout the specimen, and the cumulative AE hits increase gradually until the peak strength point. This is followed by the rapid rise of cumulative AE hits in the post-peak regime, where the microcracks coalesce, and the cohesive strength of the rock specimen degrades. On the other hand, according to Fig. 7.6, during hardening cyclic loading tests, the specimens do not experience large axial, lateral and volumetric irreversible deformations after 1500 cycles and the hysteretic loops for such tests are very dense. This clearly can be seen from the variation of volumetric strains in the area of interest (AOI) (see Fig. 7.6d). In Fig. 7.6d, it is observed that the slope of the hysteretic loops between the lowest points and the peak points is positive, implying that the current volume of the specimen is mostly at the compaction stage with slight dilation at the end of pre-peak cyclic loading. According to the evolution of damage parameters (i.e. 𝐸 and 𝐾), AE hits and the irreversible strains discussed above, the following π‘‘π‘Žπ‘› potential mechanism can be inferred for the observed quasi-elastic behaviour in this study: During cyclic loading below the fatigue threshold stress, but in the unstable crack propagation stage, some microcracks might be created within the specimens, which may result in grain size reduction and the creation of some pore spaces. The grain size reduction induced by cyclic loading also has been reported by Trippetta et al. (2013) based on the conducted microscopic analysis, although they used different loading history (i.e. damage-controlled cyclic loading tests). On the other hand, by performing additional loading and unloading cycles, the existing or newly generated defects which have been oriented horizontally are closed, and the rock specimen is compacted progressively. This is while the defects which have been oriented vertically are opened progressively. Therefore, it can be hypothesised that the observed quasi- elastic behaviour in this study can be due to the competition between two mechanisms of 206
ADE
(c) 2000 6000 (1) Initial monotonic loading stage (2) Systematic cyclic loading stage (3) Final monotonic loading until peak stress (4) Post-peak stage 1500 (5) Pre-peak stage s 5 t i ) 4000 h s e 10 14 E m 4 A s t i h Ei At ( 1000 1 stis he Em Ai () t 2468 246811 02 stih E A evitalum uC 3 2000 e v i t a l u m u 500 C 0 0 5000 10000 15000 20000 Time, t (s) 2 0 0 0 10000 20000 30000 Time, t (s) (d) 80 ) 4 60 -0 1 Β΄ ( a ,n 40 i a r t s l a i x 20 A Axial strain at the failure point, e a-peak Axial strain at the final loading cycle, e a-f 0 80 82 84 86 88 Applied stress level, s/s (%) a m Figure 7.8 a Energy components for a loading and unloading cycle, b typical evolution of the energy dissipation ratio and stiffness parameters for the specimen GS-10, c typical time- history of AE hits for the specimen GS-10, d the variation of axial strain at the final loading cycle and the failure point with stress level for hardening cyclic loading tests 7.5.2. Effect of Pre-Peak Cyclic Loading on the Post-Peak Monotonic Behaviour In Fig. 7.9, the results of hardening cyclic loading tests are compared with monotonic test results, as normalised axial stress-strain curves. As it may be seen in this figure, the overall post-peak behaviour of monotonic and hardening cyclic loading tests are almost similar. Also, the increase in cycle number at stress levels 𝜎 /𝜎 =80% (from 1500 to 10000 cycles) and π‘Ž π‘š 𝜎 /𝜎 =85% (from 1500 to 5000 cycles), has no significant influence on the general post-peak π‘Ž π‘š 208 e
ADE
behaviour. In other words, when the stress level that cyclic loading is applied is not high enough to fail the specimen during cyclic loading, the cyclic loading has a negligible effect on the post- failure behaviour. This can be further investigated based on the variation of rock brittleness. Although there is no consensus regarding the rock brittleness definition and its criterion, it is well-known that brittle rocks show small irreversible deformation before peak strength which is followed by a self-sustaining failure in the post-peak regime (Tarasov and Potvin 2013). From 1956 to date, many rock brittleness indices have been developed by different researchers; however, the strain energy-based indices perform relatively better than others (Zhang et al. 2016). The brittle vs. ductile behaviour of rock materials can be revealed in stress-strain curves during loading and failure. Thus, the rock brittleness indices, which consider the complete stress-strain behaviour of rocks may be more reliable. Munoz et al. (2016a) proposed three fracture energy-based brittleness indices considering both pre-peak and post-peak regimes of stress-strain curves for different rocks under uniaxial compressive tests. They reported that the proposed indices properly describe an unambiguous and monotonic scale of brittleness with increasing pre-peak strength parameters (i.e. 𝜎 , 𝐸 and 𝜎 ). Therefore, in this study, 𝑐𝑑 π‘‘π‘Žπ‘› π‘Žβˆ’π‘π‘’π‘Žπ‘˜ the following equations were used to measure the overall brittleness (𝐡𝐼) of the tested specimens under systematic cyclic loading. 𝐡𝐼 = π‘ˆπ‘’ = π‘ˆπ‘’ (7.1) π‘ˆπ‘‘ π‘ˆπ‘π‘Ÿπ‘’+π‘ˆπ‘π‘œπ‘ π‘‘ 𝜎2 π‘ˆ = π‘Žβˆ’π‘π‘’π‘Žπ‘˜ (7.2) 𝑒 2πΈπ‘‘π‘Žπ‘› where π‘ˆ , π‘ˆ , π‘ˆ and π‘ˆ are total fracture energy in the pre-peak and post-peak stages, 𝑑 𝑒 π‘π‘Ÿπ‘’ π‘π‘œπ‘ π‘‘ elastic energy at peak stress, the pre-peak dissipated energy and the post-peak dissipated energy, respectively. Figure 7.10a shows the different strain energy components defined above for rock brittleness determination under monotonic loading. For hardening cyclic loading tests (i.e. GS-7 to GS- 13), the final monotonic loading stress-strain curves were extracted from the stress-strain relations shown in Fig. 7.9. The strain energy components were calculated for all monotonic and hardening cyclic loading tests, and the corresponding 𝐡𝐼 values were determined. The results are listed in Table 7.3. Fig. 7.10b shows the variation of BI values for these tests. As may be seen in this figure, the 𝐡𝐼 values of the specimens tested under hardening cyclic loading are almost similar to those obtained under the monotonic loading conditions. Therefore, it can 209
ADE
for several cycles, a quasi-elastic behaviour dominated the damage evolution during the pre- peak cyclic loading. This behaviour was accompanied by the progressive rock compaction (see Fig. 7.6) and strength improvement up to 8%. It should be noted that rock strength improvement induced by cyclic loading also has been reported in several studies for porous Hawkesbury sandstone (up to 11%) (Taheri et al. 2016a, 2017), hard graywacke sandstone (up to 29%) (Singh 1989) and rock salt (up to 171%) (Ma et al. 2013). This shows that rocks depending on their intrinsic characteristics and the applied loading history and loading conditions, may show strength hardening behaviour at different extents. Taheri et al. (2017) argue that when the rock specimen is subjected to cyclic loading at a stress level lower than a threshold value, the weak bonding between the mesoscopic elements may be broken down, and the created fine materials, may fill up the internal voids, causing rock compaction and strength improvement. It should be mentioned that other potential mechanisms such as microcrack tip blunting and the interlocking of grains/asperities may involve in strength hardening. For instance, by considering the initial porosity of Gosford sandstone (i.e. 18%), due to the grain size reduction induced by cyclic loading during the quasi-elastic stage, some additional pore spaces might be generated within the specimens. When the cyclic loading-induced microcracks meet these pores, their tips may become blunt, resulting in a decrease in stress concentration at the crack tips and an increase in fracture toughness. This, on the other hand, may cause to stopping the microcrack propagation. This behaviour can also be accompanied by grain interlocking, closure of cracks, and finally, compaction of the specimens during cyclic loading. Further microscopic investigations will shed more light on cyclic loading induced hardening mechanism. (a) Monotonic loading tests ) 52 Hardening cyclic loading tests Upper limit a Average P M ( k a e 50 Upper limit p a- ,s s e r t 48 Lower limit s k a e p l 46 a i x Lower limit A 44 1 2 3 4 5 6 7 8 9 0 1 2 3 -S G -S G -S G -S G -S G -S G -S G -S G -S G 1 -S 1 -S 1 -S 1 -S G G G G Test number 213 s
ADE
(b) m 1.08 GS-13 / h GS-8 ,o GS-10 i 1.06 t a r g n i 1.04 n e d r a h h 1.02 GS-11 GS-12 t g n GS-7 e GS-9 r 1.00 t S 0.98 80 82 84 86 88 Applied stress level, s/s (%) a m Figure 7.11 a The variation of axial peak stress for all monotonic and hardening cyclic loading tests and b strength hardening ratio vs. applied stress level for hardening cyclic loading tests 7.6. Rock Behaviour During Fatigue Cyclic Loading Tests 6.1. Evaluation of Post-Peak Behaviour As discussed in section 7.5.2, the systematic cyclic loading has no notable effect on the post- peak behaviour of Gosford sandstone specimens if the cyclic stress level is below fatigue threshold stress. In this section, the influence of systematic cyclic loading beyond the fatigue threshold stress on the post-peak behaviour of Gosford sandstone specimens was evaluated. Figure 7.12 shows the normalised axial stress-strain curves for both monotonic tests and fatigue cyclic loading tests. The effect of cyclic loading history on the post-failure behaviour can be evaluated using the variation of rock brittleness index (𝐡𝐼) with the applied stress level. To do so, the envelope curve connecting the loci of the indicator stresses (𝜎) both in the pre-peak and 𝑖 the post-peak regimes were drawn, and the same procedure explained in section 7.5.2 was utilised to measure the overall brittleness index. Fig. 7.13a shows the extracted envelope curve for the typical test of GS-16. The strain energy components along with the 𝐡𝐼 values were determined for all fatigue cyclic loading tests, and the obtained values were tabulated in Table 7.3. Figure 7.13b displays the variation of 𝐡𝐼 values with the applied stress level. From this figure, it can be observed that the overall rock brittleness increases with an increase in the applied stress level. This means that rock may fail in a more brittle manner when it experiences cyclic loading at the stress levels close to its monotonic strength. In other words, in deep 214 s s
ADE
(b) 1.0 I B0.8 ,x e d n0.6 i s s e n e0.4 l t t i r B 0.2 0.0 86 88 90 92 94 96 Applied stress level, s /s (%) a m Figure 7.13 (Continued) 7.6.2. Damage Evolution in the Post-Peak Regime The irreversible deformations are not accumulated at a constant rate in the rock specimen during the pre-peak cyclic loading but follow an inverted S-shaped behaviour comprising three main phases of transient, steady and acceleration (Fig. 7.14a) (Royer-Carfagni and Salvatore 2000; Xiao et al. 2009; Fuenkajorn and Phueakphum 2010). These three phases are manifested as loose-dense-loose behaviour in the stress-strain curves of systematic cyclic loading based on the variation of hysteretic loops (Fig. 7.14b). According to Zoback and Byerlee (1975), the initial loose cycles correspond to the energy consumption for crack growth, that stabilises after several cycles. In the second phase that hysteric loops are closed and dense, the frictional work is more dominant, and the micro-cracks are opened and closed constantly without any significant extension. However, when the rock specimen is close to the failure point (i.e. the acceleration phase), the crack growth dominates, and hysteresis of the cycles increases. At higher stress levels, due to rapid accumulation of damage, the steady phase will not be visible. On the other hand, at lower stress levels (as discussed in section 7.5.1), after the initial phase, a steady-state dominates the whole test for a long time (Xiao et al. 2009). According to the stress-strain curves obtained for the fatigue cyclic loading tests in this study (Fig. 7.11), the loose-dense-loose behaviour with different extents can be identified for hysteretic loops not only in the pre-peak regime but also in the post-peak regime. For instance, Fig. 7.14c and d shows the typical results for specimen GS-23 in which the loose-dense-loose behaviours are evident. As shown in the inset figure of the axial stress-strain curve, in the pre- peak regime, the hysteretic loops follow a loose-dense-loose behaviour according to the 217
ADE
mechanism explained above. The loose behaviour at the end of the pre-peak systematic cyclic loading extends to the post-peak regime and then accelerates. In Fig. 7.14e and f the cumulative irreversible axial (βˆ‘πœ€π‘–π‘Ÿπ‘Ÿ) and cumulative irreversible lateral strains (βˆ‘πœ€π‘–π‘Ÿπ‘Ÿ) measured after full π‘Ž 𝑙 unloading of each loading cycle in the post-peak regime of specimen GS-23 are plotted against the axial stress ratio (𝜎 /𝜎 ). According to these figures, when the specimen loses its π‘Ž π‘Žβˆ’π‘π‘’π‘Žπ‘˜ load-bearing capacity until 𝜎 /𝜎 = 0.69, due to quick dissipation of strain energy, the π‘Ž π‘Žβˆ’π‘π‘’π‘Žπ‘˜ cumulative irreversible strains increases rapidly, which provided the loose hysteretic loops. Then, interestingly, the hysteretic loops are closed and experience a dense behaviour for a large number of cycles in the post-peak regime until 𝜎 /𝜎 = 0.38. Finally, by the creation of π‘Ž π‘Žβˆ’π‘π‘’π‘Žπ‘˜ large axial and lateral deformations within the specimen, the cumulative irreversible strains increased dramatically until complete failure occurred. This, in turn, provided the final loose hysteretic loops. The observed loose-dense-loose behaviour in the post-peak regime for this specimen can be summarised as a secondary inverted S-shaped damage behaviour, as shown in Fig. 7.14g. Depending on the number of cycles that the specimens have experienced after failure point, similar damage evolution trends with different extents also were observed for other fatigue cyclic loading tests. According to Table 7.2 and as shown in Fig. 7.14h, it can be observed that with the increase of applied stress level (𝜎 /𝜎 ), the number of cycles after π‘Ž π‘š failure point increases exponentially, which is consistent with the formation of the secondary three-stage inverted S-shaped behaviour in the post-peak regime. In other words, it can be found out that the damage per loading/unloading cycle in the post-peak regime of the fatigue cyclic loading tests decreases with the increase of the applied stress level. (a) (b) Loose Dense Loose a a , n Transient , i s a phase s r e t r s t s l a l i a x i A Acceleration x A phase Steady phase Cycle number, n Axial strain, e a Figure 7.14 a, b Typical inverted S-shaped damage behaviours in the pre-peak regime (Modified from Guo et al. 2012), c, d the loose-dense-loose behaviour in the post-peak 218 e s
ADE
7.7. Conclusions In this study, a series of systematic cyclic loading tests were conducted on Gosford sandstone specimens using an innovative double-criteria damage-controlled testing method. A comprehensive evaluation was carried out on the experimental results in terms of damage evolution, post-peak instability and strength hardening behaviour. The following conclusions can be drawn: 1. It was found that there exists a threshold of 𝜎 /𝜎 , which lies between 86-87.5%. For π‘Ž π‘š 𝜎 /𝜎 lower than this range, the specimens did not fail after experiencing a large π‘Ž π‘š number of cycles. The evaluation of the energy dissipation ratio, tangent Young’s modulus and AE hits for hardening cyclic loading tests showed that the rock specimens follow a two-stage damage evolution law dominated by a quasi-elastic behaviour in the pre-peak regime. This quasi-elastic behaviour can be attributed to a balance between two mechanisms of dilatant microcracking and rock compaction during cyclic loading below the fatigue threshold stress. Moreover, the damage evolution in the pre-peak regime of the hardening cyclic loading tests was found to be independent of the number of cycles, as no significant influence on damage and/or hardening behaviour was observed by increasing the cycle number from 1500 to 10000 cycles. 2. A similar pre-peak and post-peak behaviour was observed for monotonic tests and hardening cyclic loading tests when they were compared as the normalised axial stress- strain relations. Also, according to the variation of an energy-based brittleness index (𝐡𝐼), it was found that the pre-peak systematic cyclic loading has negligible influence on the post-failure instability, when the applied stress level is not high enough to fail the specimen during cyclic loading. 3. For the specimens subjected to the systematic cyclic loading below the fatigue threshold stress, the peak strength increased up to 8% after applying the monotonic loading. This strength enhancement might be due to rock compaction and porosity reduction mechanism induced by cyclic loading. On the other hand, the fatigue failure was observed for the specimens cyclically loaded beyond the fatigue threshold stress. For such tests, a rapid accumulation of lateral and volumetric strains was observed in the pre-peak regime. 4. For the systematic cyclic loading tests conducted beyond the fatigue threshold stress, it was observed that with the increase of the applied stress level, the rock specimens tend to behave as self-sustained in the post-failure stage. This was confirmed by the increase 220
ADE
of brittleness index (𝐡𝐼) with 𝜎 /𝜎 for the fatigue cyclic loading tests. Therefore, π‘Ž π‘š rocks may behave in a more brittle/violent manner when the cyclic loading is applied at stress levels close to their monotonic strength. 5. The evolution of hysteretic loops for fatigue cyclic loading tests showed that the rock specimens follow a loose-dense-loose behaviour in the pre-peak regime. However, the loose behaviour before the failure point is extended to the post-peak stage for several cycles. These loose hysteretic loops are followed by a dense behaviour for a large number of cycles until the complete failure of the specimen occurs, demonstrating another loose behaviour. This generally can be manifested as a secondary inverted non- linear S-shaped damage behaviour when the cumulative axial and cumulative lateral irreversible strains are plotted against the post-peak cycle number. It was observed that damage per cycle decreases exponentially with the increase of the applied stress level, and the three phases of the inverted S-shaped damage behaviour become more visible in the post-peak regime. Acknowledgements The first author acknowledges the University of Adelaide for providing the research fund (Beacon of Enlightenment PhD Scholarship) to conduct this study. The authors would like to thank the laboratory staff, in particular, Simon Golding and Dale Hodson, for their aids in conducting the tests. References Beniawski ZT (1967) Mechanism of brittle fracture of rock. International Journal of Rock Mechanics and Mining Sciences 4(4):395-406 Bruning T, Karakus M, Nguyen GD, Goodchild D (2018) Experimental Study on the Damage Evolution of Brittle Rock Under Triaxial Confinement with Full Circumferential Strain Control. Rock Mechanics and Rock Engineering 51(11):3321–3341 Cardani G, Meda A (2004) Marble behaviour under monotonic and cyclic loading in tension. Construction and Building materials 18(6):419–424 Cattaneo S, Labuz JF (2001) Damage of marble from cyclic loading. Journal of materials in civil engineering 13(6):459–465 Cerfontaine B, Collin F (2018) Cyclic and fatigue behaviour of rock materials: review, 221
ADE
interpretation and research perspectives. Rock Mechanics and Rock Engineering 51(2):391–414 Erarslan N, Alehossein H, Williams DJ (2014) Tensile Fracture Strength of Brisbane Tuff by Static and Cyclic Loading Tests. Rock Mechanics and Rock Engineering 47(4):1135– 1151 Fairhurst CE, Hudson JA (1999) Draft ISRM suggested method for the complete stress-strain curve for intact rock in uniaxial compression. International Journal of Rock Mechanics and Mining Sciences 36(3):279–289 Fuenkajorn K, Phueakphum D (2010) Effects of cyclic loading on mechanical properties of Maha Sarakham salt. Engineering Geology 112(1-4):43–52 Gong QM, Yin LJ, Wu SY, et al (2012) Rock burst and slabbing failure and its influence on TBM excavation at headrace tunnels in Jinping II hydropower station. Engineering Geology 124:98–108 Guo H, Ji M, Zhang Y, Zhang M (2018) Study of mechanical property of rock under uniaxial cyclic loading and unloading. Advances in Civil Engineering 2018, pp 1-6 Haimson BC (1978) Effect of Cyclic Loading on Rock. In: Silver ML, Tiedemann D (eds) Dynamic Geotechnical Testing. ASTM International, West Conshohocken, PA, pp 228– 245 Heap MJ, Faulkner DR (2008) Quantifying the evolution of static elastic properties as crystalline rock approaches failure. International Journal of Rock Mechanics and Mining Sciences 45(4):564–573 Heap MJ, Faulkner DR, Meredith PG, Vinciguerra S (2010) Elastic moduli evolution and accompanying stress changes with increasing crack damage: implications for stress changes around fault zones and volcanoes during deformation. Geophysical Journal International 183(1):225–236 Heap MJ, Vinciguerra S, Meredith PG (2009) The evolution of elastic moduli with increasing crack damage during cyclic stressing of a basalt from Mt. Etna volcano. Tectonophysics 471(1):153–160 Jansen D, Shah S (1997) Effect of Length on Compressive Strain Softening of Concrete. Journal of Engineering Mechanics 123(1):25–35 222
ADE
Li T, Pei X, Wang D, et al (2019) Nonlinear behavior and damage model for fractured rock under cyclic loading based on energy dissipation principle. Engineering Fracture Mechanics 206:330–341 Liu J, Xie H, Hou Z, et al (2014) Damage evolution of rock salt under cyclic loading in unixial tests. Acta Geotechnica 9(1):153–160 Liu Q, Huang S, Kang Y, Liu X (2015) A prediction model for uniaxial compressive strength of deteriorated rocks due to freeze-thaw. Cold Regions Science and Technology 120:96– 107 Liu X, Liang Z, Zhang Y, et al (2018) Experimental study on the monitoring of rockburst in tunnels under dry and saturated conditions using AE and infrared monitoring. Tunnelling and Underground Space Technology 82:517–528 Lockner D (1993) The role of acoustic emission in the study of rock fracture. International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts 30(7):883– 899 Ma L, Liu X, Wang M, et al (2013) Experimental investigation of the mechanical properties of rock salt under triaxial cyclic loading. International Journal of Rock Mechanics and Mining Sciences 34–41 Martin CD, Chandler N (1994) The Progressive Fracture of Lac du Bonnet Granite. International Journal of Rock Mechanics and Mining Sciences and Geomechanics Abstracts 31(6):643–659 Martin CD (1997) Seventeenth Canadian Geotechnical Colloquium: The Effect of Cohesion Loss and Stress Path on Brittle Rock Strength. Canadian Geotechnical Journal 34(5):698–725 Momeni A, Karakus M, Khanlari G, Heidari M (2015) Effects of cyclic loading on the mechanical properties of a granite. International Journal of Rock Mechanics and Mining Sciences 100(77):89-96 Munoz H, Taheri A (2019) Postpeak deformability parameters of localized and nonlocalized damage zones of rocks under cyclic loading. Geotechnical Testing Journal 42(6):1663– 1684 Munoz H, Taheri A, Chanda EK (2016a) Fracture Energy-Based Brittleness Index 223
ADE
Development and Brittleness Quantification by Pre-peak Strength Parameters in Rock Uniaxial Compression. Rock Mechanics and Rock Engineering 49(12):4587–4606 Munoz H, Taheri A, Chanda EK (2016b) Pre-peak and post-peak rock strain characteristics during uniaxial compression by 3D digital image correlation. Rock Mechanics and Rock Engineering 49(7):2541–2554 Royer-Carfagni G, Salvatore W (2000) The characterization of marble by cyclic compression loading: experimental results. Mechanics of Cohesive-frictional Materials 5(7):535–563 Shirani Faradonbeh R, Taheri A (2019) Long-term prediction of rockburst hazard in deep underground openings using three robust data mining techniques. Engineering with Computers 35(2):659–675 Shirani Faradonbeh R, Taheri A, Karakus M (2021) Post-peak behaviour of rocks under cyclic loading using a double-criteria damage-controlled test method. Bulletin of Engineering Geology and the Environment 80(2):1713–1727 Singh SK (1989) Fatigue and strain hardening behaviour of graywacke from the flagstaff formation. New South Wales. Engineering Geology 26(2):171–179 Sufian A, Russell AR (2013) Microstructural pore changes and energy dissipation in Gosford sandstone during pre-failure loading using X-ray CT. International Journal of Rock Mechanics and Mining Sciences 57:119–131 Taheri A, Hamzah N, Dai Q (2017) Degradation and improvement of mechanical properties of rock under triaxial compressive cyclic loading. Japanese Geotechnical Society Special Publication 5(2):71–78 Taheri A, Tatsuoka F (2015) Small- and large-strain behaviour of a cement-treated soil during various loading histories and testing conditions. Acta Geotechnica 10(1):131–155 Taheri A, Royle A, Yang Z, Zhao Y (2016a) Study on variations of peak strength of a sandstone during cyclic loading. Geomechanics and Geophysics for Geo-Energy and Geo-Resources 2(1):1–10 Taheri A, Yfantidis N, L. Olivares C, et al (2016b) Experimental Study on Degradation of Mechanical Properties of Sandstone Under Different Cyclic Loadings. Geotechnical Testing Journal 39(4):673-687 224
ADE
Taheri A, Zhang Y, Munoz H (2020) Performance of rock crack stress thresholds determination criteria and investigating strength and confining pressure effects. Construction and Building Materials 243:118263 Tarasov B, Potvin Y (2013) Universal criteria for rock brittleness estimation under triaxial compression. International Journal of Rock Mechanics and Mining Sciences 59:57–69 Trippetta F, Collettini C, Meredith PG, Vinciguerra S (2013) Evolution of the elastic moduli of seismogenic Triassic Evaporites subjected to cyclic stressing. Tectonophysics 592:67– 79 Vasconcelos G, LourenΓ§o P, Alves C, Pamplona J (2009) Compressive Behavior of Granite: Experimental Approach. Journal of Materials in Civil Engineering 21(9):502–511 Wang S, Xu W, Sun M, Wang W (2019) Experimental investigation of the mechanical properties of fine-grained sandstone in the triaxial cyclic loading test. Environmental earth sciences 78(14):416 Wang W, Wang M, Liu X (2016) Study on Mechanical Features of Brazilian Splitting Fatigue Tests of Salt Rock. Advances in Civil Engineering Wawersik WR, Fairhurst CH (1970) A study of brittle rock fracture in laboratory compression experiments. In: International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts. Elsevier, pp 561–575 Xiao J-Q, Ding D-X, Jiang F-L, Xu G (2010) Fatigue damage variable and evolution of rock subjected to cyclic loading. International Journal of Rock Mechanics and Mining Sciences 47(3):461–468 Zhang D, Ranjith PG, Perera MSA (2016) The brittleness indices used in rock mechanics and their application in shale hydraulic fracturing: A review. Journal of Petroleum Science and Engineering 143:158–170 Zhang J, Deng H, Taheri A, et al (2019) Deterioration and strain energy development of sandstones under quasi-static and dynamic loading after freeze-thaw cycles. Cold Regions Science and Technology 160:252–264 Zoback MD, Byerlee JD (1975) The effect of cyclic differential stress on dilatancy in westerly granite under uniaxial and triaxial conditions. Journal of Geophysical Research (1896- 1977) 80(11):1526–1530 225
ADE
Statement of Authorship Title of Paper Fatigue Failure Characteristics of Sandstone Under Different Confining Pressures Publication Status Published Accepted for Publication Submitted for Publication U npublished and Unsubmitted work written in manuscript style Publication Details Shirani Faradonbeh R, Taheri A, Karakus M (2021) Fatigue Failure Characteristics of Sandstone Under Different Confining Pressures. Rock Mechanics and Rock Engineering x(x):x–x Note: Under review [the paper submitted on 22/05/2021] Principal Author Name of Principal Author (Candidate) Roohollah Shirani Faradonbeh Contribution to the Paper Conducting laboratory tests, analysis of the results, and preparation of the manuscript Overall percentage (%) 80% Certification: This paper reports on original research I conducted during the period of my Higher Degree by Research candidature and is not subject to any obligations or contractual agreements with a third party that would constrain its inclusion in this thesis. I am the primary author of this paper. Signature Date 17 June 2021 Co-Author Contributions By signing the Statement of Authorship, each author certifies that: i. the candidate’s stated contribution to the publication is accurate (as detailed above); ii. permission is granted for the candidate in include the publication in the thesis; and iii. the sum of all co-author contributions is equal to 100% less the candidate’s stated contribution. Name of Co-Author Abbas Taheri Contribution to the Paper Research supervision, review and revision of the manuscript Signature Date 21 June 2021 Name of Co-Author Murat Karakus Contribution to the Paper Review and revision of the manuscript Signature Date 21 June 2021 226
ADE
Chapter 8 Fatigue Failure Characteristics of Sandstone Under Different Confining Pressures Abstract Rock fatigue behaviour including the fatigue threshold stress (FTS), post-peak instability and strength weakening/hardening during cyclic loading, is of paramount significance in terms of safety and stability assessment of underground openings. In this study, the evolution of the foregoing parameters for Gosford sandstone subjected to systematic cyclic loading, in the pre- peak and the post-peak regimes at different stress levels and under seven confinement levels (𝜎 /π‘ˆπΆπ‘† ) was evaluated comprehensively. The results showed that the FTS of rocks 3 π‘Žπ‘£π‘” decreases exponentially from 97% to 80%, when 𝜎 /π‘ˆπΆπ‘† increases from 10% to 100%. 3 π‘Žπ‘£π‘” The brittleness of rocks under monotonic and cyclic loading conditions increases with an increase in 𝜎 /π‘ˆπΆπ‘† when 𝜎 /π‘ˆπΆπ‘† ranging between 10-65% (known as the transition 3 π‘Žπ‘£π‘” 3 π‘Žπ‘£π‘” point). For higher confinements, however, the brittleness of rock transits from self-sustaining behaviour into ductile behaviour. The evolution of fatigue damage parameters for hardening tests showed that no critical damage happens within the specimens during cyclic loading; rather, they experience more compaction. This is while for weakening cyclic loading tests, continuous damage along with stiffness degradation was dominant. Furthermore, the variation of axial strain at failure point (πœ€ ) shows that for lower confinement levels, the applied stress π‘Žπ‘“ level does not affect the pre-peak irreversible deformation; its effect, however, becomes significant when confining pressure is high. For the specimens that did not fail in cycles, cyclic loading resulted in peak strength weakening or hardening depending on the applied stress level. Weakening effect was observed in higher confining pressures, which was mainly due to a higher amount of irreversible deformation accumulation in rocks in the pre-peak cyclic loading. An empirical model was proposed using classification and regression tree (CART) algorithm to estimate the peak strength variation of Gosford sandstone based on 𝜎 /π‘ˆπΆπ‘† and the 3 π‘Žπ‘£π‘” applied stress level. Keywords: Triaxial loading, Systematic cyclic loading, Confinement level, Brittleness, Fatigue threshold stress, Strength hardening/weakening 227
ADE
List of Symbols 𝑀 Post-peak modulus π‘ž Peak deviator stress π‘š 𝐸 Pre-peak modulus π‘ž Residual deviator stress π‘Ÿπ‘’π‘  𝑁 Number of cycles before failure π‘ž /π‘ž Deviator stress level 𝑒𝑛 π‘šβˆ’π‘Žπ‘£π‘” 𝑅 Strain gauge resistance π‘ž /π‘ž Fatigue threshold stress 𝑓 π‘šβˆ’π‘Žπ‘£π‘” π‘ž Deviator stress 𝜎 /π‘ˆπΆπ‘† Confinement level 3 π‘Žπ‘£π‘” 𝐡𝐼 Brittleness index πœ€ Axial strain at failure π‘Žπ‘“ 𝐺𝐹 Strain gauge factor πœ€ Lateral strain at failure 𝑙𝑓 βˆ†π‘… Change in resistance πœ€π‘–π‘Ÿπ‘Ÿ Irreversible axial strain π‘Ž 𝐴𝐸 Acoustic emission π‘‘πœ€ /𝑑𝑑 Lateral strain rate 𝑙 𝐹𝑇𝑆 Fatigue threshold stress π‘‘πœ€ /𝑑𝑑 Axial strain rate π‘Ž 𝐢𝐴𝑅𝑇 Classification and regression tree π‘‘π‘ˆ Shear rupture energy π‘Ÿ 𝑉 Output voltage π‘‘π‘ˆ Withdrawn elastic energy π‘œ 𝑒 𝑉 Excitation voltage π‘‘π‘ˆ Residual elastic energy 𝑒π‘₯ π‘’π‘Ÿ πœ€ Mechanical strain π‘‘π‘ˆ Additional energy π‘Ž 𝐸 Tangent Young’s modulus πœ”π‘–π‘Ÿπ‘Ÿ Cumulative irreversible axial strain π‘‘π‘Žπ‘› π‘Ž π‘ˆπΆπ‘† Uniaxial compressive strength βˆ†πœ€π‘–π‘Ÿπ‘Ÿ Differential irreversible axial strain π‘Ž π‘ˆ Total elastic energy 𝜎 Major principal stress 𝑒 1 π΄π‘šπ‘.(πœ€ ) Lateral strain amplitude 𝜎 Confining pressure 𝑙 3 8.1. Introduction Depending on the depth, the geometry of the structures and the human- and/or environmental- induced seismic activities, rock masses in underground mining and geotechnical projects are usually subjected to a complex stress state, which may result in continuous damage and failure at different extents (Yang et al. 2017; Wang et al. 2021). Systematic cyclic loading induced by the rock breakage operation, mechanical excavation, and truck haulage vibrations is a common dynamic disturbance in underground openings that complicate the deformation and failure characteristics of rocks. Rock materials under such loading conditions are more prone to severe failure phenomena such as strain bursting and large-scale collapses (Bagde and PetroΕ‘ 2005; Munoz and Taheri 2019, Shirani Faradonbeh et al. 2021a; Meng et al. 2021). Therefore, there is a remarkable theoretical significance and engineering value to deeply understand the cyclic loading effect on the damage mechanism and, more importantly, the post-failure behaviour of rocks in terms of safety and long-term stability of the excavations. During the last decades, 228
ADE
different researchers have made many attempts to unveil the rock fatigue mechanism under different loading conditions using laboratory experiments (Cerfontaine and Collin 2018). In other words, the damage evolution mechanism in rocks can be characterised more efficiently using cyclic loading tests as it is straightforward to distinguish the elastic and plastic strains during each loading and unloading cycle (Zhou et al. 2019; Tian et al. 2021). According to the holistic classification proposed by Shirani Faradonbeh et al. (2021a), rock fatigue studies can be classified into two main groups of systematic cyclic loading tests and damage-controlled cyclic loading tests. Each of these groups can be performed either under load-controlled or displacement-controlled loading conditions. These loading techniques and their limitations have been discussed in more detail by Shirani Faradonbeh et al. (2021a). Generally, the rock fatigue studies can be discussed from two viewpoints: the pre-peak and post-peak domain analysis. From the viewpoint of the pre-peak-domain analysis, the literature review shows that cyclic loading depending on loading methods, loading conditions and intrinsic rock properties (e.g. porosity and mineral compositions) can either degrade (Wang et al. 2013; Erarslan et al. 2014; Yang et al. 2015; Taheri et al. 2016a) or improve (Burdine 1963; Singh 1989; Ma et al. 2013; Shirani Faradonbeh et al. 2021b) the peak strength of rocks. For instance, Ma et al. (2013) reported a 171.1% increase in triaxial compressive strength of rock salt subjected to systematic cyclic loading. Similarly, Taheri et al. (2016b) observed an 11% peak strength improvement for the porous Hawkesbury sandstone, and they also pointed out that rock strength increases respectively with applied stress level and the number of cycles before failure following linear and exponential functions. On the other hand, most of the fatigue cyclic loading studies have reported peak strength and stiffness degradation due to the accumulation of permanent deformations within the rock specimens following a non-linear S- shaped damage model (e.g. Xiao et al. 2009). Fatigue threshold stress (FTS = π‘ž /π‘ž ), the 𝑓 π‘šβˆ’π‘Žπ‘£π‘” maximum stress level at which rock specimen does not fail during cyclic loading under a constant amplitude, is a significant parameter for long-term stability assessment of underground openings subjected to seismic disturbances. In other words, rock materials never fail (after a few thousand cycles) if the cyclic loading is applied equal or below this threshold. According to Cerfontaine and Collin (2018), different values of FTS can be obtained depending on the tested material. However, FTS is also dependent on other factors, such as loading conditions and confining pressure (Burdine 1963). Therefore, more investigations are needed to unveil the effect of confining pressure on fatigue threshold stress. 229
ADE
From the viewpoint of the post-peak domain, due to difficulties in capturing the complete stress-strain relations of rocks under cyclic loading, especially for brittle rocks which show a class II post-peak behaviour (Wawersik and Fairhurst 1970), very few studies have investigated the influence of the pre-peak cyclic loading on post-failure behaviour. In most prior studies, the damage-controlled cyclic loading tests (with the incremental loading amplitude) have been used under axial displacement-controlled loading conditions to evaluate the post-peak behaviour (e.g. Yang et al. 2015, 2017; Zhou et al. 2019; Meng et al. 2021). These studies, however, were not sufficient to adequately measure the post-peak response of rocks. This is because, during each loading cycle, the axial load is reversed when a certain amount of displacement is achieved, and after the failure point, since most of the rocks show class II or a combination of class I and class II behaviours, rock failure occurs in an uncontrolled manner. However, Munoz and Taheri (2017) showed that lateral displacement control throughout the test is a promising technique in studying the failure behaviour of rocks subjected to the post- peak cyclic loading. Recently, Shirani Faradonbeh et al. (2021a and b) developed a novel testing methodology based on the lateral strain feedback signal to measure the complete pre- peak and post-peak behaviour of rocks under uniaxial systematic cyclic loading. Although many studies have been undertaken by different researchers on the evolution of rock fatigue damage and deformability parameters under different loading histories and loading conditions, no significant progress has been made regarding the effect of systematic cyclic loading on the cyclic loading-induced strength hardening, fatigue threshold stress and the post- peak instability of rocks under different confining pressures. This is while in underground rock engineering projects, rock materials are usually subjected to triaxial loading conditions with different levels of confinement accompanied by the systematic cyclic loading induced by different dynamic sources. Therefore, having in-depth knowledge concerning the foregoing parameters plays a critical role in stability assessment and reinforcement design. This study, for the first time, investigates the effect of systematic cyclic loading history on pre-peak and post-peak characteristics of rocks under different confinement levels. Some empirical equations are then proposed to manifest the evolution of peak strength, fatigue threshold stress and rock brittleness parameters. The obtained results are expected to provide a better understanding of the mechanical response of rocks to systematic cyclic loading under various confining pressures. 230
ADE
8.2. Experimental Profile 8.2.1. Gosford Sandstone In this study, Gosford sandstone (Fig. 8.1a) extracted from the massive Triassic Hawkesbury sandstone of the Sydney Basin, New South Wales, Australia, was chosen as the testing material (Ord et al. 1991; Masoumi et al. 2017). X-ray powder diffraction (XRD) analysis of this medium-grained (0.2-0.3 mm) sandstone revealed that quartz (86%) is the dominant mineral and illite (7%), kaolinite (6%) and anatase (1%) are other forming mineral composition. Fig. 8.1b displays the SEM analysis result of this sandstone. Sufian and Russell (2013) reported that Gosford sandstone has a total porosity of about 18%, and the density distribution of the pre- existing micro-cracks within its matrix is homogenous. This type of sandstone is usually known as a uniform or very uniform sandstone (Hoskins 1969; Vaneghi et al. 2018). Cylindrical specimens (Fig. 8.1a) having 42 mm diameter and 100 mm length were extracted from a single rock block and prepared following the ISRM recommended standards (Fairhurst and Hudson 1999). The specimens were air-dried before conducting the static and cyclic loading tests, and the average dry density of this rock type was approximately about 2.215 g/cm3. (a) 42 mm m m 0 0 1 (b) Kaolinite Quartz Illite Quartz 231
ADE
Figure 8.1 Gosford sandstone used in this study: a prepared specimens and b SEM photograph 8.2.2. Testing Equipment A fully digital closed-loop servo-controlled hydraulic compressive machine, i.e. Instron-1282 with the maximum loading capacity of 1000 kN, was employed to conduct the triaxial monotonic and cyclic loading tests. The testing machine can be programmed and equipped to perform different loading schemes using either the load-controlled or displacement-controlled loading techniques. As shown in Fig. 8.2a, a Hoek cell with a maximum capacity of 65 MPa was used to apply confining pressure. Also, a pair of LVDTs were installed between the loading platens to measure the axial displacement of the specimens during loading. Strain gauges are commonly used to measure the axial and/or lateral deformations of rocks in triaxial conditions. However, the strain gauges are only effective for local small-strain measurement, and they usually break after the peak stress when the specimen experiences large deformations (Munoz et al. 2016a; Bruning et al. 2018). A modified test arrangement is made to overcome this problem; four strain gauges were attached immediately alongside one another around the centre line of the Hoek cell membrane, as displayed in Fig. 8.2b. Then, the strain gauges were connected to form a Wheatstone bridge (half-bridge circuit). Any deformation in specimen changes the resistance and, therefore, facilitates a unique output voltage (𝑉) as a lateral strain π‘œ feedback signal. In the Wheatstone bridge shown in Fig. 8.2b, 𝑅 and 𝑅 represent the total 1 3 resistance values provided by the pairs of strain gauges (each gauge has 120Ξ© resistance) which are connected in series. To balance the bridge and achieve zero voltage when the specimen is unstrained, two 240 Ξ© precision resistors (i.e. 𝑅 and 𝑅 ) were used in this circuit. The feedback 2 4 signal, indeed, is the average of the lateral strain (πœ€ ) values measured by the strain gauges, 𝑙 which is calculated as follows: 𝑉 = 𝑉𝑒π‘₯(βˆ†π‘…1+βˆ†π‘…3) = 𝑉𝑒π‘₯.𝐺𝐹.(πœ€ +πœ€ ) (8.1) π‘œ 1 3 4 𝑅1 𝑅3 4 βˆ†π‘…/𝑅 GF = (8.2) πœ€ where 𝑅 is the resistance of the undeformed strain gauge, βˆ†π‘… is the change in resistance caused by strain, πœ€ is the mechanical strain, 𝐺𝐹 is the strain gauge factor and 𝑉 is the bridge excitation 𝑒π‘₯ voltage. 232
ADE
Through a high-pressure wire and a feed-through connector fitted to the Hoek cell, the feedback signal is sent to the control unit of the testing machine to adjust the loading rate. By doing so, the membrane gauges are protected from damage during loading, and finally, the complete lateral deformation of rocks can be recorded in both pre-peak and post-peak regimes. Moreover, two miniature AE sensors (type PICO, from the American Physical Acoustics Corp.) were attached to the spherical seats, which have a direct connection to the specimen in the Hoek cell, to record the microcracking process during loading. The pre-amplifier was set to 60 dB of gain (Type 2/4/6) to amplify the acoustic emission (AE) signals during loading. To ensure that mechanical noises induced by the loading system are not recorded during the tests, the AE threshold amplitude was changed from 20 dB to 60 dB, and it was found that after 40 dB amplitude, no additional noises are recorded. Therefore, this value was set as the AE threshold. The axial load, axial and lateral displacements, and the AE outputs were recorded simultaneously by running the tests. (a) Loading platen AE sensors LVDT1 LVDT2 Hoek cell Hydraulic pressure inlet High-pressure wire feed-through Figure 8.2 Experimental set-up, a overview of the experiment and b strain gauged membrane 233
ADE
(b) High-pressure wire Strain gauges R R 1 2 + Wheatstone Bridge - + V V ex o - R 4 R 3 Figure 8.2 (Continued) 8.3. Test Scheme and Conditions 8.3.1. Uniaxial and Triaxial Monotonic Loading Tests Before conducting the triaxial monotonic and cyclic loading tests at different confining pressures, the uniaxial compressive strength (π‘ˆπΆπ‘†) of Gosford sandstone should be determined. Shirani Faradonbeh et al. (2021b) performed a series of uniaxial monotonic tests on this rock type under a constant lateral strain rate (π‘‘πœ€ /𝑑𝑑) of 2Γ—10-6/s. In their study, the 𝑙 axial strain was measured using a pair of external LVDTs, and the lateral strain feedback signal was measured using a direct-contact chain extensometer. Fig. 8.3a shows the normalised stress- strain relations of the performed uniaxial monotonic tests. As it is shown in this figure, the rock specimens are quite uniform and demonstrate almost similar pre-peak and post-peak stress- strain relations. Gosford sandstone has an average uniaxial peak strength (π‘ˆπΆπ‘† ) and tangent π‘Žπ‘£π‘” Young’s modulus (𝐸 ) values of 48.15 MPa and 13.4 GPa, respectively. π‘‘π‘Žπ‘›βˆ’π‘Žπ‘£π‘” Based on the determined π‘ˆπΆπ‘† , seven different confinement levels, i.e. 𝜎 /π‘ˆπΆπ‘† = 10%, π‘Žπ‘£π‘” 3 π‘Žπ‘£π‘” 20%, 35%, 50%, 65%, 80% and 100%, were adopted for triaxial monotonic and cyclic compression tests. For each confinement level, three triaxial monotonic tests were carried out. 234
ADE
The tests were conducted in a way that the axial load and confining pressure were applied simultaneously to the rock specimen under a constant axial strain rate of π‘‘πœ€ /𝑑𝑑= 0.03 mm/min π‘Ž until the desired confining pressure level is achieved. Thereafter, the confining pressure and axial load were kept constant for five minutes to ensure the stress was distributed uniformly (pre-consolidation stage). Then, while the confining pressure was maintained constant, the deviator stress (i.e. π‘ž = 𝜎 βˆ’πœŽ ) was applied under a constant lateral strain rate (π‘‘πœ€ /𝑑𝑑) of 1 3 𝑙 2Γ—10-6/s until the complete failure occurs. The lateral strain rate was adjusted during the test based on the feedback signal received from the four strain gauges mounted on the Hoek cell membrane. Fig. 8.3b shows a typical time history of stress and strains during a triaxial compression test at 𝜎 /π‘ˆπΆπ‘† =10%. Table 8.1 presents a summary of results for all conducted 3 π‘Žπ‘£π‘” triaxial monotonic tests. Fig 3c, shows the representative stress-strain relations for the triaxial monotonic tests. According to Table 8.1 and Fig. 8.3c, the increase in 𝜎 /π‘ˆπΆπ‘† , affected 3 π‘Žπ‘£π‘” both the pre-peak and the post-peak characteristics of rocks. Generally, with an increase in confining pressure, the axial strain at the failure point (πœ€ ) increases. Also, as shown in Fig. π‘Žπ‘“ 8.3d, the average peak deviator stress (π‘ž ) of Gosford sandstone increased by confining π‘šβˆ’π‘Žπ‘£π‘” pressure following a quadratic trend. Section 5 discusses the triaxial compression test results in more detail. (a) 50 ) a P 40 M ( a 30 ,s s e r ts 20 la ix A 10 0 0 20 40 60 80 Axial strain, e (Β΄10-4) a Figure 8.3 a Normalised stress-strain relations for uniaxial monotonic tests, modified from Shirani Faradonbeh et al. (2021b), b typical time-history of stress and strains for a triaxial monotonic test at 10% confinement level, c representative stress-strain relations for triaxial monotonic tests at different confinement levels and d the variation of peak deviator stress with confinement level 235 s
ADE
8.4. Confining Pressure Effect on Fatigue Threshold Stress As mentioned earlier, fatigue threshold stress (FTS) is a critical parameter, that can be used as an effective compressive strength of the intact rock subjected to static, dynamic and cyclic loads. Depending on the rock type, testing method and loading history, various range of values for FTS were reported by different researchers. Table 8.3 reviews these studies and lists the used materials and testing methods along with the determined FTSs. Table 8.3 shows that most of the existing studies have been conducted in uniaxial loading condition. Taheri et al. (2016b) performed the systematic cyclic loading tests on Hawkesbury sandstone under a single confining pressure of 𝜎 = 4 MPa. In an earlier study, Burdine (1963) performed a series of 3 triaxial dynamic loading tests under three confining pressures (i.e. 𝜎 = 0.21 MPa, 1.38 MPa 3 and 5.17 MPa) on Berea sandstone. The study showed that with an increase in confining pressure from 0 to 5.17 MPa, the fatigue threshold stress increases from 74% to 93% of the monotonic strength. In the current study, a more comprehensive range of confining pressure was considered to evaluate the variation of FTS under systematic cyclic loading for Gosford sandstone. According to Table 8.2, for each confinement level, a fatigue threshold stress (π‘ž /π‘ž ) can 𝑓 π‘šβˆ’π‘Žπ‘£π‘” be derived. Fig. 8.7 plots the variation of the determined FTS values against the confinement level. As can be seen in this figure, with an increase in 𝜎 /π‘ˆπΆπ‘† from 10% to 100%, 3 π‘Žπ‘£π‘” π‘ž /π‘ž decreases constantly, which shows the weakening/negative influence of confining 𝑓 π‘šβˆ’π‘Žπ‘£π‘” pressure on the fatigue life of the rock under cyclic loading. These results, show that in underground projects, with the increase of depth, rock materials may fail at a stress level lower than the determined monotonic strength. The behavioural trend observed for FTS in this study is in contrast to that reported by Burdine (1963). According to Fig. 8.7, the FTS can be predicted using the following logarithmic function with high accuracy: 𝐹𝑇𝑆 = π‘ž 𝑓 = βˆ’0.074𝐿𝑛( 𝜎3 )+0.806 ; 𝑅2 = 0.982 (8.3) π‘žπ‘šβˆ’π‘Žπ‘£π‘” π‘ˆπΆπ‘†π‘Žπ‘£π‘” Also, based on the proposed Eq. 3, a binary condition can be defined to classify the failure status of the rock specimens, i.e. occurrence (1) or non-occurrence (0), under a specific stress level and confining pressure as follows: 1 π‘ž /π‘ž > 𝐹𝑇𝑆 𝑒𝑛 π‘šβˆ’π‘Žπ‘£π‘” Failure status= { (8.4) 0 π‘ž /π‘ž ≀ 𝐹𝑇𝑆 𝑒𝑛 π‘šβˆ’π‘Žπ‘£π‘” 242
ADE
8.5. Confining Pressure Effect on Post-Peak Instability As mentioned earlier, the post-peak instability of rocks can be characterised as class I and class II, representing the stable and unstable rock fracturing process under a specific loading history, respectively. Brittleness is an appropriate intact rock property that can be employed to quantify the post-peak instability. Many rock brittleness indices can be found in the literature (Meng et al. 2020). However, as the evolution of strain energy accompanies the process of rock deformation and failure, the energy balance-based indices can better reflect the post-peak instability and the potential of severe failures (Li et al. 2019). Therefore, in this study, the following strain energy-based brittleness indices (𝐡𝐼s) proposed by Tarasov and Potvin (2013) were used to evaluate the post-peak instability of rocks: 𝐡𝐼 = π‘‘π‘ˆπ‘Ÿ = π‘€βˆ’πΈ (8.5) 1 π‘‘π‘ˆπ‘’ 𝑀 𝐡𝐼 = π‘‘π‘ˆπ‘Ž = 𝐸 (8.6) 2 π‘‘π‘ˆπ‘’ 𝑀 π‘‘π‘ˆ = π‘ž 𝐡2βˆ’π‘ž 𝐴2 𝑒 2𝐸 π‘ž2βˆ’π‘ž2 (8.7) π‘‘π‘ˆ = 𝐡 𝐴 π‘Ž 2𝑀 π‘‘π‘ˆ = π‘‘π‘ˆ βˆ’π‘‘π‘ˆ { π‘Ÿ 𝑒 π‘Ž where π‘‘π‘ˆ , π‘‘π‘ˆ and π‘‘π‘ˆ are, respectively, the withdrawn elastic energy, the additional/excess 𝑒 π‘Ž π‘Ÿ energy and the shear rupture energy in the post-peak regime (see Fig. 8.8). The π‘ž and π‘ž are 𝐴 𝐡 the deviator stresses corresponding to points A and B, respectively, and 𝐸 and 𝑀 are, respectively, the pre-peak and the post-peak modulus. To evaluate the effect of both confining pressure and loading history on rock brittleness, 𝐡𝐼 1 and 𝐡𝐼 were calculated for all monotonic and the cyclic loading tests (the tests that 2 experienced the final monotonic loading). The evolution of the average 𝐡𝐼 values was plotted against 𝜎 /π‘ˆπΆπ‘† in Fig. 8.9. Shirani Faradonbeh et al. (2021b) performed a series of uniaxial 3 π‘Žπ‘£π‘” systematic cyclic loading tests on Gosford sandstone at different stress levels and found that below the fatigue threshold stress, the rock brittleness values are similar to those obtained in monotonic loading conditions. In this study, the 𝐡𝐼 values were calculated again for all uniaxial monotonic and cyclic loading tests using Eqs. 8.5 and 8.6. According to Fig. 8.9, similar 𝐡𝐼 values were obtained for these two types of tests in uniaxial conditions. Also, as can be seen in Fig. 8.9, with an increase in 𝜎 /π‘ˆπΆπ‘† from 0% to 65%, the rock brittleness for both 3 π‘Žπ‘£π‘” monotonic and cyclic loading tests changed similarly from an almost transitional state (i.e. 244
ADE
𝐡𝐼 β‰ˆ 1 and 𝐡𝐼 β‰ˆ 0) to more class II/brittle behaviour. By increasing the confining pressure 1 2 to a certain amount (i.e. 𝜎 /π‘ˆπΆπ‘† =50%), the maximum rock brittleness was achieved, and 3 π‘Žπ‘£π‘” then, the 𝐡𝐼 values showed a decremental trend. A drastic drop in 𝐡𝐼 was observed for 𝜎 /π‘ˆπΆπ‘† > 65%, specifically for cyclic loading tests, where the rock specimens transferred 3 π‘Žπ‘£π‘” from the class II region (green area) to the class I region (yellow area). Indeed, there is more opposition against the self-sustaining failure at high confinement levels, and more energy should be added axially by the loading system to yield the specimen completely. Therefore, a transition point at 65% confinement level can be estimated for Gosford sandstone, as the rock specimens transfer from a brittle to ductile failure behaviour. The evolutionary trend observed in Fig. 8.9 is also consistent with the stress-strain curves of rocks shown in Fig. 8.3c. Similar unconventional trends for 𝐡𝐼 also have been reported in a few studies, (i.e. Tarasov and Potvin 2013 and Ai et al. 2016), for stronger rocks such as quartzite and black shale. According to these studies, the increase in brittleness of rocks with confining pressure can be attributed to the energy-efficient fan-head mode shear failure. Indeed, during Class II failure behaviour, a domino structure of blocks is created by tensile cracks along the future failure plane. Due to the fracture propagation, these blocks are rotated without collapse behaving as hinges and create a fan-shaped structure in the fracture tip. This, in turn, provides an active force (negative shear resistance) that is beneficial for maintaining the crack propagation and is responsible for the self-sustaining failure behaviour of rocks. Therefore, the increase in confining pressure for these rock types seems to provide a higher amount of active forces and consequently increases rock brittleness. By considering the decremental trend of fatigue threshold stress with confinement level, discussed in the previous section, as well as the incremental trend of rock brittleness with confinement for a specific extent, it can be inferred that with an increase in depth in rock engineering projects, the propensity of rock structures to violent/brittle failures such as strain bursting at stress levels lower than the determined average peak strength can be aggravated. The brittleness reduction at high confinement levels can be attributed to the more plastic deformation accumulation induced by the loading and unloading cycles within the specimens, which result in more energy dissipation in the pre-peak regime. This, in turn, provides less amount of elastic strain energy (the source for self-sustaining behaviour) at the failure point, leading to more ductile post-peak behaviour. This behaviour is more evident for cyclic loading tests than monotonic ones due to the more weakening effect of loading and unloading cycles at higher confinement levels. The damage evolution of rocks under different confinement levels is evaluated in more detail in section 8.6. 245
ADE
8.6. Confining Pressure Effect on Fatigue Damage Evolution 8.6.1. Hardening and Weakening Cyclic Loading Tests Rock specimens usually experience deformation under external forces, and a part of this deformation can be recovered by withdrawing the applied force, representing elastic characteristics. However, owing to intrinsic material properties, e.g., porosity and microcracks, and loading-induced damage, the complete deformation recovery after unloading is not possible. Therefore, a certain amount of irreversible/plastic deformation is retained in the specimens (Taheri and Tatsuoka, 2015; Peng et al. 2019). The irreversible strain is accumulated incrementally by applying more cycles, which is accompanied by rock stiffness degradation. Cumulative strain can be utilised to manifest the non-visible damage incurred in the specimen during the systematic cyclic loading tests (Taheri et al. 2016b). According to Table 8.2, for the specimens that did not fail during 1000 loading and unloading cycles, two types of tests can be distinguished based on peak strength variation: strength weakening tests (i.e., final monotonic loading strength is less than π‘ˆπΆπ‘† ) and strength hardening tests (i.e., final monotonic loading π‘Žπ‘£π‘” strength is more than π‘ˆπΆπ‘† ). As seen in Table 8.2, the strength weakening is evident for the π‘Žπ‘£π‘” tests undertaken under 𝜎 /π‘ˆπΆπ‘† β‰₯ 80%. To appraise the rock damage evolution in both 3 π‘Žπ‘£π‘” conditions, the cumulative irreversible axial strain (πœ”π‘–π‘Ÿπ‘Ÿ) and tangent Young’s modulus (𝐸 ) π‘Ž π‘‘π‘Žπ‘› were determined for two representative tests. Fig. 8.10 shows the variation of πœ”π‘–π‘Ÿπ‘Ÿ and 𝐸 π‘Ž π‘‘π‘Žπ‘› for specimens GS-C-15 (with 4.38% strength hardening) and GS-C-31 (with -3.96% strength weakening) at 35% and 100% confinement levels, respectively. The other weakening and hardening cyclic loading tests also showed similar behaviour. According to Fig. 8.10, for both specimens, the elastic modulus increased notably for initial cycles, making the specimens stiffer and more difficult to deform. This can be related to the closure of pre-existing defects and yield surface expansion during cyclic loading (Taheri and Tatsuoka 2015; Peng et al. 2019). However, for specimen GS-C-15 (i.e., hardening test), by performing further cycles, the stiffness of the specimen decreased slightly and then remained almost constant until 1000 cycles were completed, which is consistent with the trend observed by Ma et al. (2013) triaxial systematic cyclic loading tests. On the other hand, during the initial cycles for specimen GS-C-15, πœ”π‘–π‘Ÿπ‘Ÿ evolved slightly to a certain amount due to the primary π‘Ž loose hysteretic loops, and then like 𝐸 , retained almost constant, which shows that no more π‘‘π‘Žπ‘› damage is cumulated within the specimen. As stated by Shirani Faradonbeh et al. (2021b), this quasi-elastic behaviour can be due to the competition between the mechanisms of grain-size 247
ADE
reduction and rock compaction under consecutive loading and unloading cycles. For specimen GS-C-31 (i.e., weakening test), although no failure was recorded during the cycles, a different trend for variations of πœ”π‘–π‘Ÿπ‘Ÿ was observed (see Fig. 8.10). For the weakening test, πœ”π‘–π‘Ÿπ‘Ÿ increased π‘Ž π‘Ž rapidly, first for several cycles (i.e., initial hysteretic loops), and then by experiencing the dense hysteretic loops, shows a linear increase. At the end of cyclic loading, the increase of πœ”π‘–π‘Ÿπ‘Ÿ π‘Ž becomes more pronounced which may indicate that the specimen could have failed during cyclic loading should the test be continued. These results are consistent with 𝐸 variations for π‘‘π‘Žπ‘› the weakening test, shown in Fig. 8.10. As can be seen in this figure, unlike the hardening test, the damage evolution for weakening test was accompanied by the progressive stiffness degradation of rock during the whole cyclic loading test. Therefore, it can be stated that the strength weakening observed in Table 8.2 for systematic cyclic loading tests can be relevant to the progressive damage evolution/stiffness degradation of rocks in the pre-peak regime, which is aggravated when confining pressure exceeds the transition point (i.e. 𝜎 /π‘ˆπΆπ‘† >65%). 3 π‘Žπ‘£π‘” This is while for lower confinement levels, when cyclic stress level is low enough, cyclic loading has no considerable effect on damage evolution; rather, improves peak strength. The above observations are further investigated using AE results. E wirr tan a 20 ) 4 -0 19 1 ) a Β΄ P G rri( a ( ,s u luE n a t 18 ( h a G rdS e- nC in- g1 5 test) 15 ,n ia r ts la d ix o m s GS-C-31 ( W e aG kS e- nC in-3 g1 test) 10 a e lb 'g n 17 (Weakening test) is r u e o v Y GS-C-15 e tn e g (hardening test) 5 r r i e v n a 16 ita T lu m 0 u C 0.0 0.2 0.4 0.6 0.8 1.0 Relative cycle number, n/N Figure 8.10 Typical evolution of πœ”π‘–π‘Ÿπ‘Ÿ and 𝐸 for hardening and weakening cyclic loading π‘Ž π‘‘π‘Žπ‘› tests 248 w
ADE
8.6.1.1. Acoustic Emission Characteristics Acoustic emission (AE) is a well-known non-destructive technique that can monitor the micro and macrocrack evolution in rocks during loading in real-time. Due to the local micro-scale deformations, small fracturing events corresponding to the immediate release of strain energy are created in the form of elastic waves within the specimens. Recording and analysing these elastic waves during the tests can directly measure internal damage (Cox and Meredith 1993; Lockner 1993). Therefore, the AE technique was utilised to elucidate the cracking procedure during the hardening and weakening cyclic loading tests better. In this regard, the evolution of AE hits, representing the number of generated cracks, and its cumulation throughout the representative hardening and weakening tests GS-C-15 and GS-C-31 were respectively depicted in Figs. 8.11a and b. To better unveil the damage mechanism under different confining pressures, the AE results of specimen GS-C-29 (𝜎 /π‘ˆπΆπ‘† =80%) which showed the greatest 3 π‘Žπ‘£π‘” peak strength decrease (i.e. -13.18% strength weakening) were also displayed in Fig. 8.11c. As shown in Fig. 8.11, the evolution of AE hits for the specimens can be investigated throughout three main loading phases: initial monotonic loading (phase A), systematic cyclic loading (phase B) and final monotonic loading (phase C). For all three specimens, during the seating of loading platens on the specimens and the closure of pre-existing defects, few AE hits were recorded in stage A and cumulative AE hits increased slightly. For specimen GS-C-15 (𝜎 /π‘ˆπΆπ‘† =35% and π‘ž /π‘ž = 80%), as shown in Fig. 8.11a, the cumulative AE hits 3 π‘Žπ‘£π‘” 𝑒𝑛 π‘šβˆ’π‘Žπ‘£π‘” then remained almost constant (i.e. quasi-elastic behaviour) during loading and unloading cycles. The zoomed-in figure also shows only small amounts of low-amplitude AE hits during phase B. The cumulated AE hits at the end of stage B is almost 1.77% of the total damage experienced by the specimen during the test. This shows that no considerable cyclic loading induced damage is generated should the specimens be loaded below the fatigue threshold stress and at confinement levels lower than the transition point. This behaviour also is consistent with the variation of πœ”π‘–π‘Ÿπ‘Ÿ discussed in the previous section. The majority of rock damage for π‘Ž specimen GS-C-15 occurred in phase C, where the final monotonic loading was applied to the specimen. In this phase, due to opening the compacted microcracks, the generation of new ones and their coalescence close to and after peak strength point, the cohesive strength of rock is gradually substituted by the frictional resistance, which was accompanied by a higher amount of AE hits. Unlike specimen GS-C-15 which showed a quasi-elastic behaviour during the systematic cyclic loading, a different AE evolution behaviour was observed for specimen GS-C-31 249
ADE
(𝜎 /π‘ˆπΆπ‘† =100% and π‘ž /π‘ž =80%) in phase B. According to Fig. 8.11b, after a slight 3 π‘Žπ‘£π‘” 𝑒𝑛 π‘šβˆ’π‘Žπ‘£π‘” increase in AE hits during the initial monotonic loading, the microcracking increased with a higher rate by increasing loading and unloading cycles in phase B, which is manifested by a higher number of AE hits. The cumulated AE hits at the end of phase B is almost 27.09% of the total damage incurred in the specimen throughout the test, which is relatively higher than that observed for specimen GS-C-15. As discussed earlier, this microcracking induced by cyclic loading results in stiffness degradation (see Fig. 8.10) and more ductile behaviour in the pre-peak regime. The generated damage was not enough to fail the specimen, however, it resulted in strength weakening of -3.96% during the final monotonic loading. For specimen GS-C-29 which experienced a -13.18% decrease in peak strength at 80% confinement level, as seen in Fig. 8.11c, by applying systematic cyclic loading, the AE hits began to grow first with a lower rate until about 500 cycles were completed. Then by performing further cycles, the rate of AE hits cumulation increased dramatically, representing the continuous generation of macrocracks within the specimen. According to Fig. 8.11c, about 93.90% of the total rock damage happened at the end of phase B, which is far greater than those observed for specimens GS-C-15 and GS-C-31. Based on the above observations for AE outputs, it can be stated that for confinement levels beyond the transition point (𝜎 /π‘ˆπΆπ‘† = 65%), although cyclic loading 3 π‘Žπ‘£π‘” below the fatigue threshold stress does not lead to fatigue failure during 1000 loading cycles, it creates significant damage, which results in a considerable strength weakening during final monotonic loading. (a) 3500 A: Initial monotonic loading pahse 9000 B: Systematic cyclic loading phase 3000 C: Final monotonic loading phase 7500 2500 6 150 s t s t i h E 2000 stih E A 24 140stih E A e v ita lu 46 50 00 00 i h E A e v i A 1500 130m u C C t a l u 0 3000 m 1000 40 80 120 160 u Time, t C 500 A B 1500 0 0 0 50 100 150 200 Time, t (min ) Figure 8.11 Representative AE results for cyclic loading tests: a hardening test (GS-C-15), b weakening test (GS-C-31) and c weakening test (GS-C-29) 250
ADE
(b) 4000 35000 A: Initial monotonic loading pahse 3500 B C: : S Fy ins ate l m ma ot nic o c toy nc il cic l olo aa dd inin gg p p hh aa ss ee 30000 3000 s 25000 t i h 2500 E s t i h 2000 A B C 20000 A e v E A 15000 i t a 1500 l u m 10000 u 1000 C 500 5000 0 0 0 100 200 300 Time, t (min ) (c) 3000 30000 A: Initial monotonic loading pahse B: Systematic cyclic loading phase C 2500 C: Final monotonic loading phase 25000 s t 2000 20000 i h E s A t i h 1500 15000 e E A A B v i t a l u 1000 10000 m u C 500 5000 0 0 0 100 200 300 Time, t (min ) Figure 8.11 (Continued) 8.6.2. Damage Cyclic Loading Tests In this section, the effect of confining pressure is evaluated on the Gosford sandstone specimens which failed during loading and unloading cycles, i.e., damage cyclic loading tests. Fig. 8.12a displays the variation of πœ”π‘–π‘Ÿπ‘Ÿ for damage cyclic loading tests under different confinement π‘Ž levels. To prevent Fig. 8.12a be crowded, only one damage test was considered for each confinement level. Generally, the irreversible strain increased quickly at the beginning of the tests. Then, a relatively uniform accumulation in strain followed by a rapid strain increase as the rock specimens head toward failure. As is clear from Fig. 8.12a, the damage accumulation rate increased by an increase in 𝜎 /π‘ˆπΆπ‘† from 10 to 100%. This damage evolution, however, 3 π‘Žπ‘£π‘” is more significant for the tests undertaken under high confining pressures (i.e. over the transition point) where the irreversible/plastic deformations largely incurred in the pre-peak 251
ADE
8.6.3. Applied Stress Level Effect on Damage Evolution As stated earlier, systematic cyclic loading was applied to the specimens at different stress levels (π‘ž /π‘ž ). To evaluate the effect of the applied stress level on damage evolution of 𝑒𝑛 π‘šβˆ’π‘Žπ‘£π‘” rocks under different confining pressures, the axial strain at the failure point (πœ€ ) was π‘Žπ‘“ determined for all monotonic and cyclic loading tests. The results were listed in Tables 8.1 and 8.2. For uniaxial monotonic and cyclic loading conditions, πœ€ values were adapted from π‘Žπ‘“ Shirani Faradonbeh et al. (2021b). Fig. 8.14 represents the variation of πœ€ for monotonic, π‘Žπ‘“ hardening, weakening and damage cyclic loading tests with π‘ž /π‘ž . It can be seen from 𝑒𝑛 π‘šβˆ’π‘Žπ‘£π‘” Fig. 8.14 that under a specific confinement level (i.e. 35%), cyclic loading at various stress levels has no significant influence on πœ€ and their values are almost similar to those obtained π‘Žπ‘“ for monotonic loading tests. However, for higher confinements, larger values of πœ€ is observed π‘Žπ‘“ at the stress levels equal to or greater than the fatigue threshold stresses, due to the accumulation of irreversible strain in the sample during the pre-peak regime before the failure. The above behaviour is more evident in Fig. 8.15, where the variation of average axial strain at failure point (πœ€ ) for different stress levels was depicted against 𝜎 /π‘ˆπΆπ‘† . As seen π‘Žπ‘“βˆ’π‘Žπ‘£π‘” 3 π‘Žπ‘£π‘” in this figure, for monotonic loading tests, πœ€ evolved linearly with the increase of π‘Žπ‘“βˆ’π‘Žπ‘£π‘” 𝜎 /π‘ˆπΆπ‘† ; this is while, for hardening/weakening and damage cyclic loading tests, this 3 π‘Žπ‘£π‘” evolution occurred exponentially. According to Fig. 8.15, for 𝜎 /π‘ˆπΆπ‘† ≀35%, the 3 π‘Žπ‘£π‘” monotonic and cyclic loading tests have almost similar πœ€ values, which means that π‘Žπ‘“βˆ’π‘Žπ‘£π‘” loading and unloading cycles below and beyond the fatigue threshold stress have no striking influence on pre-peak behaviour, and damage evolution under cyclic loading is similar to monotonic loading conditions. However, for higher confinement levels, πœ€ increased first π‘Žπ‘“βˆ’π‘Žπ‘£π‘” gradually until 𝜎 /π‘ˆπΆπ‘† = 65% representing more accumulation of plastic deformations 3 π‘Žπ‘£π‘” within the specimens in the pre-peak regime compared with the monotonic loading conditions. The evolutionary trend of πœ€ , then, was aggravated for confinement levels of 80 and π‘Žπ‘“βˆ’π‘Žπ‘£π‘” 100%, where a sharp increase in πœ€ was observed for weakening and damage cyclic π‘Žπ‘“βˆ’π‘Žπ‘£π‘” loading tests. 255
ADE
Figure 8.14 Variation of axial strain at failure point for monotonic and cyclic loading tests under different confinement levels: a 0%, b 10%, c 20%, d 35%, e 50%, f 65%, g 80% and h 100% ) 4 -0 Damage cyclic loading tests 1 (Β΄ 480 Hardening/weakening cyclic loading tests g v Monotonic loading tests a -a 400 ,e y=57.405e0.018x r u l i 320 R2= 0.904 a f t a y=63.822e0.013x n i 240 R2= 0.801 a r t s l 160 a i x a e 80 g y=0.994x+62.440 a r e R2= 0.956 v 0 A 0 20 40 60 80 100 Confinement level, s/UCS (%) 3 avg Figure 8.15 Average axial strain at failure for monotonic and cyclic loading tests 8.7. Confining Pressure Effect on Strength Hardening/Weakening 8.7.1. Peak Strength Variation As seen in Table 8.2, depending on the stress level that cyclic loading is applied as well as the confinement level, rock specimens have experienced different values of increase/decrease in peak strength during final monotonic loadings. As discussed in sections 8.6.1 and 8.6.1.1, when the stress level during cyclic loading is low enough (i.e. lower than the estimated FTS), cyclic loading at lower confinement levels did not create macro-damage in the specimens, and a quasi- elastic behaviour dominated the rock damage evolution. This, in turn, resulted in a hardening behaviour under loading and unloading cycles, and consequently, strength improvement which is observed during final monotonic loading. The rock compaction due to cyclic loading in the hardening region is evident in Fig. 8.5 for the representative test GS-C-13 (with 4.06% hardening), where the specimen did not experience large axial, lateral and volumetric irreversible strains, and the rock volume was entirely in the compaction stage during cyclic loading. This is while for rocks that failed during cycles (see Fig. 8.6), relatively higher strain values were recorded, and rocks were mainly in the dilation-dominated stage. The strength hardening induced by cyclic loading also has been reported by other researchers for different 257 e
ADE
rock types under various loading conditions, such as Gosford sandstone (up to 7.82% increase) under uniaxial systematic cyclic loading (Shirani Faradonbeh et al. 2021b), Tuffeau limestone under uniaxial multi-level systematic cyclic loading (up to 28.55% increase) (Shirani Faradonbeh et al. 2021a), hard graywacke sandstone under uniaxial systematic cyclic loading (up to 29% increase) (Singh 1989), Hawkesbury sandstone under triaxial systematic cyclic loading (up to 11% increase) (Taheri et al. 2016b) and rock salt under triaxial systematic cyclic loading (up to 171% increase) (Ma et al. 2013). Fig. 8.16a represents variation in peak strength with confinement level (𝜎 /π‘ˆπΆπ‘† ). The 3 π‘Žπ‘£π‘’ results of hardening tests under uniaxial condition (𝜎 =0) were extracted from Shirani 3 Faradonbeh et al. (2021b). According to Fig. 8.16a and Table 8.2, the peak strength parameter varies between two distinct zones, i.e. hardening zone and damage zone. Also, the maximum increase and decrease in peak strength values of Gosford sandstone specimens are 7.82% and -13.18%, respectively. Generally, with an increase in 𝜎 /π‘ˆπΆπ‘† , the amount of strength 3 π‘Žπ‘£π‘” hardening induced by cyclic loading decreased and when 𝜎 /π‘ˆπΆπ‘† > 65% (i.e. transition 3 π‘Žπ‘£π‘” point), rock specimens demonstrate strength weakening behaviour (see Fig. 8.16a). To better reflect the mechanism behind the rock moving from hardening into weakening, a parameter is proposed as below: βˆ†πœ€π‘–π‘Ÿπ‘Ÿ = (πœ€π‘–π‘Ÿπ‘Ÿ) βˆ’(πœ€π‘–π‘Ÿπ‘Ÿ) (8.8) π‘Ž π‘Ž 𝑓 π‘Ž 𝑖 where βˆ†πœ€π‘–π‘Ÿπ‘Ÿ is the differential irreversible axial strain (measured between valley points), and π‘Ž (πœ€π‘–π‘Ÿπ‘Ÿ) and (πœ€π‘–π‘Ÿπ‘Ÿ) are, respectively, the irreversible axial strains measured for final and initial π‘Ž 𝑓 π‘Ž 𝑖 loading cycles. Fig. 8.16b demonstrates the variation of βˆ†πœ€π‘–π‘Ÿπ‘Ÿ for cyclic loading tests at different stress levels π‘Ž with 𝜎 /π‘ˆπΆπ‘† . As can be seen in this figure, the range of variation for βˆ†πœ€π‘–π‘Ÿπ‘Ÿ increased 3 π‘Žπ‘£π‘” π‘Ž continuously with an increase in confining pressure, and this is more significant for 𝜎 /π‘ˆπΆπ‘† > 65%, where a high amount of irreversible deformation was experienced by the 3 π‘Žπ‘£π‘” specimens. The incremental trend of βˆ†πœ€π‘–π‘Ÿπ‘Ÿ with confinement results in more plastic behaviour π‘Ž and, therefore, pre-peak damage even when cycles don’t result in a failure. This, finally, resulted in a decremental trend of the maximum peak strength variation at each confinement level under cyclic loading, as shown in Fig. 8.16c. 258
ADE
8.7.2. An Empirical Model for Strength Prediction As discussed above, the study on strength variation of rocks under the coupled influence of cyclic loading and confining pressure is rare and limited to some specific confining pressures. Therefore, no empirical model can be found in the literature to predict strength variation after loading cycles. The classification and regression tree (CART) algorithm was employed in this study to predict the amount of strength hardening/weakening in Gosford sandstone after cyclic loading history. The CART algorithm, developed by Breiman et al. (1984), is a computational- statistical algorithm that can predict the target variable in the form of a decision tree. The CART tree is created by the binary splitting of the datasets from the root node into two sub-nodes using all predictor variables. The best predictor usually is chosen based on impurity or diversity measures (e.g. Gini, twoing and least squared deviation). The aim is to create subsets of the data which are as homogeneous as possible concerning the output variable. For each split, each input parameter (predictor) is evaluated to find the best groupings of categories (for nominal and ordinal predictors) or cut point (for continuous predictors) according to the improving score or reduction in impurity. Thereafter, the predictors are compared, and the predictor with the greatest improvement is selected for the split. This process is repeated until one of the stopping criteria (e.g. the maximum tree depth) is met (Salimi et al. 2016; Liang et al. 2016; Khandelwal et al. 2017). A detailed description of the CART algorithm can be found in (Breiman et al. 1984). In this study, the applied stress level (π‘ž /π‘ž ) and confinement level (𝜎 /π‘ˆπΆπ‘† ) were 𝑒𝑛 π‘šβˆ’π‘Žπ‘£π‘” 3 π‘Žπ‘£π‘” defined as input variables to predict the percentage of strength hardening/weakening as output variable. Based on the results presented in Table 8.2 and the conducted cyclic loading tests in uniaxial conditions by Shirani Faradonbeh et al. (2021b), a database containing 28 tests that experienced a monotonic loading after a cyclic loading history was compiled. The test GS-C- 29, which showed -13.18% strength weakening was identified as an outlier (in terms of statistics) and excluded from the modelling procedure. The CART parameters, including the maximum tree depth, impurity index and the minimum size of parent and child nodes (i.e. the minimum number of objects that a node must contain to be split) were changed for different runs to obtain a predictive model with high accuracy and low complexity. Finally, the best model was achieved according to the settings listed in Table 8.4. The modelling procedure was carried out in the MatLab environment. Fig. 8.17 represents the obtained regression tree for the best model. As shown in this figure, the developed regression tree provides a practical tool to estimate the percentage variation of the peak strength straightforwardly. Fig. 8.18 compares 260
ADE
) % ( 8 n o i t a i r a 4 v h t g n e 0 r t s k a e p -4 y=0.8996x+0.2625 d e R2= 0.90 t c i d e -8 r P -8 -6 -4 -2 0 2 4 6 8 Measured peak strength variation (%) Fig. 8.18 The comparison of the measured and predicted values of peak strength variation 8.8. Conclusions Triaxial monotonic and cyclic loading tests were undertaken in this study on Gosford sandstone at different confinement levels to scrutinise the effect of both systematic cyclic loading history and confining pressure on the evolution of rock fatigue characteristics. For this aim, a modified triaxial testing procedure was employed to control the axial load during the tests using a constant lateral strain feedback signal. Based on the experimental results, the following conclusions were drawn: 1. The confining pressure displayed a significant effect on fatigue threshold stress (FTS). It was found that with an increase in 𝜎 /π‘ˆπΆπ‘† from 10% to 100%, FTS decreases 3 π‘Žπ‘£π‘” from 97% to 80%. This indicates that rocks in great depth experience failure due to cyclic loading at stress levels much lower than the determined monotonic strength. 2. According to the obtained stress-strain relations, the post-peak behaviour of rocks followed an unconventional trend with the increase in confining pressure so that for lower 𝜎 /π‘ˆπΆπ‘† , rock specimens showed a self-sustaining (brittle) failure behaviour, 3 π‘Žπ‘£π‘” while for higher 𝜎 /π‘ˆπΆπ‘† , the ductile behaviour was dominant. The post-peak 3 π‘Žπ‘£π‘” instability of rocks was quantified using strain energy-based brittleness indices (𝐡𝐼𝑠), and a transition point at 𝜎 /π‘ˆπΆπ‘† = 65% was identified, where the rocks transited 3 π‘Žπ‘£π‘” from the brittle failure behaviour to ductile one. The results also showed that cyclic loading at confinement levels lower than the transition point has no notable effect on 263
ADE
rock brittleness, while for 𝜎 /π‘ˆπΆπ‘† = 80% and 100%, the weakening effect of 3 π‘Žπ‘£π‘” systematic cyclic loading history on rock brittleness was more significant. 3. Fatigue damage evaluation of rocks using different parameters (i.e. 𝐸 , πœ”π‘–π‘Ÿπ‘Ÿand AE π‘‘π‘Žπ‘› π‘Ž hits) showed that for hardening cyclic loading tests, no macro-damage is observed within the specimens, and the stiffness of the rocks remain almost constant during a large number of cycles, representing a quasi-elastic behaviour. However, for weakening cyclic loading tests, although no failure was observed during cycles, 𝐸 and πœ”π‘–π‘Ÿπ‘Ÿ π‘‘π‘Žπ‘› π‘Ž increased and decreased, respectively, with cycle loading. Compared to the hardening cyclic loading tests, the AE activities (micro-cracking) was more evident for specimens that showed a higher amount of strength degradation. On the other hand, for damage cyclic loading tests, it was found that damage is accumulated with a higher rate and extent with an increase in confining pressure. 4. Looking at the variation of axial strain at the failure point (πœ€ ) for monotonic, π‘Žπ‘“ hardening/weakening and damage cyclic loading tests, it was found that under confinement levels below the transition point, the applied stress level has no notable effect on the cumulation of irreversible deformations in the pre-peak regime and the values of πœ€ are similar to those in monotonic loading conditions. However, for higher π‘Žπ‘“ confinements, cyclic loading resulted in larger irreversible strain values before the failure point. 5. After a cyclic loading history, the peak deviator stress of Gosford sandstone varied between -13.18% and 7.82%. According to the evolution of damage parameters, the observed quasi-elastic behaviour during cyclic loading and the variation of plastic axial, lateral and volumetric strains for hardening cyclic loading tests, the strength hardening can be related to the rock compaction induced by cyclic loading. It was observed that the increase in confining pressure decreases the amount of strength hardening due to the accumulation of irreversible strains in the rock specimens. An empirical regression tree-based model was proposed to estimate peak strength variation of Gosford sandstone based on the applied stress level and confining pressure. The results showed the high accuracy of the model. Acknowledgements The first author acknowledges the University of Adelaide for providing the research fund (Beacon of Enlightenment PhD Scholarship) to conduct this study. The authors would like to thank the laboratory technicians particularly Simon Golding and Dale Hodson, for their aids in 264
ADE
Sciences 42(2):237–250 Breiman L, Friedman J, Stone CJ, Olshen RA (1984) Classification and Regression Trees. CRC press Bruning T, Karakus M, Nguyen GD, Goodchild D (2018) Experimental Study on the Damage Evolution of Brittle Rock Under Triaxial Confinement with Full Circumferential Strain Control. Rock Mechanics and Rock Engineering 51(11):3321–3341 Burdine NT (1963) Rock Failure Under Dynamic Loading Conditions. Society of Petroleum Engineers Journal 3(1):1–8 Cerfontaine B, Collin F (2018) Cyclic and fatigue behaviour of rock materials: review, interpretation and research perspectives. Rock Mechanics and Rock Engineering 51(2):391–414 Cox SJD, Meredith PG (1993) Microcrack formation and material softening in rock measured by monitoring acoustic emissions. In: International journal of rock mechanics and mining sciences & geomechanics abstracts. Elsevier 30(1):11–24 Erarslan N, Alehossein H, Williams DJ (2014) Tensile Fracture Strength of Brisbane Tuff by Static and Cyclic Loading Tests. Rock Mechanics and Rock Engineering 47(4):1135– 1151 Erarslan N, Williams DJ (2012) Investigating the effect of cyclic loading on the indirect tensile strength of rocks. Rock Mechanics and Rock Engineering 45(3):327–340 Grover HJ, Dehlinger P, McClure GM (1950) Investigation of fatigue characteristics of rocks. Rep Battelle Mem Inst Drill Res Inc Guo Y, Yang C, Mao H (2012) Mechanical properties of Jintan mine rock salt under complex stress paths. International Journal of Rock Mechanics and Mining Sciences 56:54–61 Haimson BC, Kim C (1971) Mechanical behaviour of rock under cyclic fatigue. Proceedings of the 13th symposium on rock mechanics. ASCE, New York, pp 845–863 Hoskins ER (1969) The failure of thick-walled hollow cylinders of isotropic rock. In: International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts. Elsevier 6(1):99–125 Jamali Zavareh S, Baghbanan A, Hashemolhosseini H, Haghgouei H (2017) Effect of 269
ADE
microstructure on fatigue behavior of intact rocks under completely reversed loading. Analytical and Numerical Methods in Mining Engineering 6:55–62 Khandelwal M, Armaghani DJ, Faradonbeh RS, et al (2017) Classification and regression tree technique in estimating peak particle velocity caused by blasting. Engineering with Computers 33(1):45–53 Li N, Zou Y, Zhang S, et al (2019) Rock brittleness evaluation based on energy dissipation under triaxial compression. Journal of Petroleum Science and Engineering 183:106349 Liang M, Mohamad ET, Faradonbeh RS, et al (2016) Rock strength assessment based on regression tree technique. Engineering with Computers 32(2):343–354 Lockner D (1993) The role of acoustic emission in the study of rock fracture. International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts 30:883–899 Ma L, Liu X, Wang M, et al (2013) Experimental investigation of the mechanical properties of rock salt under triaxial cyclic loading. International Journal of Rock Mechanics and Mining Sciences 34–41 Masoumi H, Horne J, Timms W (2017) Establishing empirical relationships for the effects of water content on the mechanical behavior of Gosford sandstone. Rock Mechanics and Rock Engineering 50(8):2235–2242 Meng F, Wong LNY, Zhou H (2020) Rock brittleness indices and their applications to different fields of rock engineering: A review. Journal of Rock Mechanics and Geotechnical Engineering Meng Q, Liu J-F, Ren L, et al (2021) Experimental Study on Rock Strength and Deformation Characteristics Under Triaxial Cyclic Loading and Unloading Conditions. Rock Mechanics and Rock Engineering 54(2):777-797 Munoz H, Taheri A (2017) Local damage and progressive localisation in porous sandstone during cyclic loading. Rock Mechanics and Rock Engineering 50(12):3253–3259 Munoz H, Taheri A, Chanda EK (2016) Pre-peak and post-peak rock strain characteristics during uniaxial compression by 3D digital image correlation. Rock Mechanics and Rock Engineering 49(7):2541–2554 Munoz H, Taheri A (2019) Post-peak deformability parameters of localised and non-localised 270
ADE
damage zones of rocks under cyclic loading. Geotechnical Testing Journal 42(6), 1663- 1684 Nejati HR, Ghazvinian A (2014) Brittleness effect on rock fatigue damage evolution. Rock Mechanics and Rock Engineering 47(5):1839–1848 Ord A, Vardoulakis I, Kajewski R (1991) Shear band formation in Gosford sandstone. In: International journal of rock mechanics and mining sciences & geomechanics abstracts. Elsevier 28(5):397–409 Peng K, Zhou J, Zou Q, Yan F (2019) Deformation characteristics of sandstones during cyclic loading and unloading with varying lower limits of stress under different confining pressures. International Journal of Fatigue 127:82–100 Salimi A, Faradonbeh RS, Monjezi M, Moormann C (2016) TBM performance estimation using a classification and regression tree (CART) technique. Bulletin of Engineering Geology and the Environment 77(1):429-440 Shirani Faradonbeh R, Taheri A, Karakus M (2021a) Post-peak behaviour of rocks under cyclic loading using a double-criteria damage-controlled test method. Bulletin of Engineering Geology and the Environment 80(2):1713-1727 Shirani Faradonbeh R, Taheri A, Karakus M (2021b) Failure Behaviour of a Sandstone Subjected to the Systematic Cyclic Loading: Insights from the Double-Criteria Damage- Controlled Test Method. Rock Mechanics and Rock Engineering [In Press] Singh SK (1989) Fatigue and strain hardening behaviour of graywacke from the flagstaff formation, New South Wales. Engineering Geology 26(2):171–179 Sufian A, Russell AR (2013) Microstructural pore changes and energy dissipation in Gosford sandstone during pre-failure loading using X-ray CT. International Journal of Rock Mechanics and Mining Sciences 57:119–131 Taheri A, Yfantidis N, L. Olivares C, et al. (2016a) Experimental study on degradation of mechanical properties of sandstone under different cyclic loadings. Geotechnical Testing Journal 39(4):673-687 Taheri A, Royle A, Yang Z, Zhao Y (2016b) Study on variations of peak strength of a sandstone during cyclic loading. Geomechanics and Geophysics for Geo-Energy and Geo-Resources 2(1):1–10 271
ADE
Taheri A, Tatsuoka F (2015) Small- and large-strain behaviour of a cement-treated soil during various loading histories and testing conditions. Acta Geotechnica 10(1):131–155 Tarasov B, Potvin Y (2013) Universal criteria for rock brittleness estimation under triaxial compression. International Journal of Rock Mechanics and Mining Sciences 59:57–69 Tian W-L, Yang S-Q, Wang J-G, Dong J-P (2021) Failure behavior of the thermal treated granite under triaxial cyclic loading–unloading compression. Geomechanics and Geophysics for Geo-Energy and Geo-Resources 7(1):1–21 Vaneghi RG, Ferdosi B, Okoth AD, Kuek B (2018) Strength degradation of sandstone and granodiorite under uniaxial cyclic loading. Journal of Rock Mechanics and Geotechnical Engineering 10(1):117–126 Wang Y, Feng WK, Hu RL, Li CH (2021) Fracture evolution and energy characteristics during marble failure under triaxial fatigue cyclic and confining pressure unloading (FC-CPU) conditions. Rock Mechanics and Rock Engineering 54(2):799–818 Wang Z, Li S, Qiao L, Zhao J (2013) Fatigue behavior of granite subjected to cyclic loading under triaxial compression condition. Rock Mechanics and Rock Engineering 46(6):1603– 1615 Wawersik WR, Fairhurst CH (1970) A study of brittle rock fracture in laboratory compression experiments. In: International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts. Elsevier, pp 561–575 Xiao J-Q, Ding D-X, Xu G, Jiang F-L (2009) Inverted S-shaped model for nonlinear fatigue damage of rock. International Journal of Rock Mechanics and Mining Sciences 3(46):643–648 Yamashita S, Sugimoto F, Imai T, et al (1999) The relationship between the failure process of the creep or fatigue test and of the conventional compression test on rock. In: 9th ISRM Congress. International Society for Rock Mechanics and Rock Engineering Yang S-Q, Ranjith PG, Huang Y-H, et al (2015) Experimental investigation on mechanical damage characteristics of sandstone under triaxial cyclic loading. Geophysical Journal International 201(2):662–682 Yang SQ, Tian WL, Ranjith PG (2017) Experimental Investigation on Deformation Failure Characteristics of Crystalline Marble Under Triaxial Cyclic Loading. Rock Mechanics 272
ADE
Chapter 9 Conclusions and Recommendations 9.1. Conclusions In this thesis, state-of-the-art methodologies comprising machine learning (ML)- and experimental-based approaches were employed to investigate the rockburst phenomenon in detail. The significant findings and major contributions of the conducted research project can be outlined as follows: β€’ The statistical analysis techniques, including the box-plot, principal component analysis (PCA) and agglomerative hierarchical clustering (AHC) were identified as robust tools to visually represent the distribution of data points, analyse the interrelationship of the parameters, detect the outliers and natural groups in the datasets and finally, prepare a homogeneous database. [see Chapters 2, 4 and 5] β€’ The three ML algorithms of gene expression programming (GEP), genetic algorithm- based emotional neural network (GA-ENN) and the decision tree-based C4.5 algorithm showed the high performance in predicting the occurrence or non-occurrence of rockburst hazard as a binary classification problem (i.e. the prediction accuracy was higher than 80%). [see Chapter 2] β€’ The hybrid GA-ENN algorithm overcame the limitations of the prior ANNs (e.g., getting trapped in local minima) and provided a global solution for the problem. The C4.5, as a white-box ML algorithm, provided a visual simple tree structure for determining the rockburst status straightforwardly based on the specific range of values defined by the algorithm for different input parameters. The GEP algorithm, unlike the other ML techniques, through its inherent capability of function finding, successfully detected the latent complex non-linear relationship between the input parameters and the corresponding output. The GEP algorithm can open the black-box nature of the common ML algorithms and by providing the explicit models, facilitates the in-depth investigation of mining and geotechnical hazards. [see Chapter 2] 274
ADE
β€’ The results of the sensitivity analysis conducted on the developed GEP-based binary model for rockburst status prediction revealed that the input parameters of maximum tangential stress (𝜎 ), elastic energy index (π‘Š ), uniaxial tensile strength (𝜎 ) and uniaxial πœƒ 𝑒𝑑 𝑑 compressive strength (𝜎 ) have the highest influence on rockbursting in deep underground 𝑐 mines, respectively. Due to the significant role of 𝜎 in rockburst occurrence, more πœƒ considerations should be taken into account during the design stage of the underground projects to control this parameter (i.e. by optimisation of the mining layout). [see Chapter 2] β€’ The comparison of the five conventional rockburst criteria, i.e., Russeness criterion, Hoek criterion, stress coefficient criterion, brittleness index criterion and elastic energy index (EEI) criterion, with the proposed ML-based models, showed that except for EEI criterion, the other conventional criteria have the prediction accuracy lower than 80% and cannot provide reliable estimations in practice. This can be attributed to the case study-based nature of the conventional criteria and considering few input parameters in their equations. [see Chapter 2] β€’ The complex relationship between different strength/stress- and energy-based parameters with the rockburst risk levels (i.e. the intensities of β€œnone”, β€œlight”, β€œmoderate” and β€œstrong”) was recognised with high accuracy using the unsupervised learning algorithm of self-organising map (SOM). This algorithm, through an intelligent procedure, categorised the rockburst events having similar conditions in distinct clusters. [see Chapter 3] β€’ The determined weighted distances between the clusters by the SOM algorithm were also consistent with the rockburst intensities defined by the engineers. This demonstrated the high capability of this technique in adapting to mining-related problems, specifically for rockburst risk level investigation as a multi-class problem. [see Chapter 3] β€’ The evaluation of the weights of input variables in each cluster revealed that the maximum tangential stress of the surrounding rock mass (𝜎 ) has the strongest influence on πœƒ rockbursting, which is consistent with the results of the binary classification of rockburst status reported in Chapter 2. [see Chapter 3] β€’ The SOM algorithm with the value of 100% for the five performance indices of accuracy rate, precision, recall, F1 score and Kappa, proved its superiority over fuzzy c-mean (FCM) algorithm and the rockburst conventional criteria in clustering the rockburst risk levels. [see Chapter 3] β€’ The intact rock properties (i.e., uniaxial compressive strength, tensile strength, elastic 275
ADE
modulus, and Poisson’s ratio) represented a significant effect on the failure mechanism (i.e., squeezing, slabbing, and strain burst) of the competent overs-stressed rock masses. The initial assessment of the compiled database from different underground mining projects showed that the failure mechanisms cannot be predicted solely by a single indicator. [see Chapter 4] β€’ Although the GEP algorithm can provide a mathematical equation to estimate the output parameter, it cannot be used solely to solve multi-class classification problems such as failure mechanism detection. It was found that the combination of the GEP algorithm with the logistic regression (LR) is an efficient methodology to overcome this difficulty. The GEP score calculated for each binary model of the failure mechanisms can be fed into the logistic regression as the independent variable to determine the occurrence probability of each failure mechanism. The failure mechanism having the highest probability value is selected as the final prediction. [see Chapter 4] β€’ According to the results of the confusion matrices and the receiver operating (ROC) curves, the developed GEP-based binary models in this research project were able to predict the status (occurrence or non-occurrence) of each failure mechanism, respectively, with 100% (AUC=1), 100% (AUC=1), and 97.14% (AUC=0.964) accuracy for squeezing, slabbing and strain bursting failure. However, the developed multi-class classifier of GEP- LR predicted the final class of failure based on the given intact rock properties with 100% accuracy. [see Chapter 4] β€’ The further validation of the GEP-LR model with nine unseen/new datasets also proved the high capability of this model in predicting the failure mechanisms accurately. Therefore, the developed GEP-LR model can be used as a practical tool by engineers and researchers to measure the propensity of the competent over-stressed rock masses to different failure mechanisms at the preliminary stages of the projects. [see Chapter 4] β€’ It was found that the maximum rockburst stress (𝜎 ), i.e., the stress level that bursting 𝑅𝐡 occurs and the rockburst risk level (𝐼 ) inferred from the conducted comprehensive true- 𝑅𝐡 triaxial unloading tests are appropriate and reliable indices to investigate the rockburst phenomenon. [see Chapter 5] β€’ The correlation analysis and the stepwise selection and elimination (SSE) procedure were identified as efficient tools for dimension reduction (i.e., recognition of the most influential parameters), removing the multicollinearity among the independent parameters, and 276