University
stringclasses
19 values
Text
stringlengths
458
20.7k
Colorado School of Mines
[55][56]. These programs use empirical relationships to determine a pillar stability factor which indicates the likelihood of unstable failure. The sophistication of these programs lies in the extensive collection of case histories of pillar failures in U.S. underground coal mines. The finite difference programs FLAC2D and FLAC3D, are highly versatile in definition of geometry, boundary conditions, and material properties [41, 42]. Each program includes numerous material models suitable for simulating a variety of geomaterials. Notable for rock is the strain softening plasticity model, which has been used to simulate yielding coal pillars in longwall mining [4]. The program has also been used to model stable and unstable failure modes in laboratory and in situ conditions [26]. 2.3.2 Simulating Unstable Failure in Underground Mining Specific to the problem of failure mode in underground mining, several sources can be cited that deal with unstable failure. In 1983, Zubelewicz and Mroz used a finite element model to study the violent failure of rock in various underground situations [95]. First, the static equilibrium is achieved, then the full equations of motion are solved explicitly after a disturbance is applied to the system. Kinetic energy is monitored, and if the energy increases drastically, the failure is considered to be unstable. Bardet created a finite element model to investigate surface buckling as a trigger for unstable failure [6]. These researchers claimed, by citing bifurcation theory, that instability can be detected when the stiffness matrix of the finite element grid becomes singular. The moment of instability was determined by finding the time step when eigenvalues of the stiffness matrix become negative. Muller followed up this study in 1991 with a comparison of explicit and implicit numerical methods in modeling unstable failure [59]. The author performed simulations in ANSYS, an explicit finite element program, and FLAC, an implicit finite difference program. Muller concluded that ANSYS was unable to represent the instability, but FLAC was successful by responding to instabilities with increases in local unbalanced force [2]. In their 1996 publication, Oelfke et al. presented a combined DEM-FEM code applicable to underground mine deformability [60]. The authors introduced the concept of mine instability as a function of local mine 20
Colorado School of Mines
stiffness and noted that the program could detect unstable failure as a divergence of the solution. Another group of researchers investigated the effect of a fractured rock mass as a loading system [14]. In this study, Chen et al. used a finite element model, called RFPA2D, to study the behavior of microseismicity during unstable failure [19]. They loaded a double rock sample in displacement control and monitored acoustic emission events. The authors claim that unstable failure can be detected as sudden changes in the microseismic rate, and demonstrated this with a realistic loading system with finite strength. In a 2009 study, Tan et al. suggested that unstable failure of pillars could be modeled using a discrete element model composed of two dimensional circular elements [81] These researchers used particle velocity to describe the intensity of failure. In a study using the program FLAC3D, Jiang et al. defined a term called the local energy release rate (LERR) that they claim can be used to describe the intensity of failure [46]. The LERR is the difference in stored strain energy in an element before and after failure. The authors compared LERR computed from simulations to known cases of unstable failure and showed that comparisons of magnitude of LERR were the same as the comparisons of intensity for the observed cases. Although, they stated that it is not possible to determine at what value of LERR an unstable failure occurs. A publication by Larson and Whyatt reviewed available stress analysis tools for under- ground coal mining [53]. They compared the use of three numerical models in simulating a deep western coal mine with strong, stiff overlying strata: ALPS, an empirical model, LaModel, a boundary element method program, and FLAC2D, a finite difference method program (FDM). In their study, they showed that FLAC2D was able to model the sudden collapse of the mine entry due to failure in the roof and floor while LaModel and ALPS do not possess this ability due to the assumption of elastic overlying strata. Following this, Esterhuizen et al. presented a method to determine the ground response curve in FLAC models, and showed that the span to depth ratio has a notable effect on the ground response curve [23]. 21
Colorado School of Mines
In some cases, strong consideration is given to how the structure of the rock mass is an integral part of its material behavior. Typically, the addition of joints, joint sets, and bedding planes brings an element of realism to the model. The computer programs UDEC and 3DEC allow the user to insert joints and joint sets with a variety of joint constitutive behaviors in 2d or 3d geometries respectively [45]. Barton 1995 presents an evaluation of the influence of joint properties on rock mass models that contain systems of joints [8]. 2.3.3 Discrete Element Modeling Techniques and Applications One other popular method of modeling rock is the discrete element method or DEM [18]. The two-dimensional method uses a collection of discs to simulate a granular material by detecting contact between the discs and calculating subsequent motion due to contact forces. Spheres can be substituted for discs to create a three-dimensional model, and if the elements are bonded together a solid material can be simulated [68]. DEM is powerful as a numerical modeling method because the user does not input a constitutive material law. Rather, the user specifies a set of micro-parameters that define stiffness and strength of the discs and bonds. Calibration of material behavior by selecting the appropriate micromechanical param- eters is an area that has widely been explored, but still needs much improvement. When calibrating a DEM model it is advantageous to understand the effects of changing micro- parameters on macro behavior. Potyondy and Cundall provided the initial guidance for choosing micro-parameters in their seminal paper [68]. Since then, several researchers have published papers on material calibration and sensitivity of macroscopic behavior on various micro-parameters. Kulatilake et al. demonstrated an iterative calibration scheme for rock behavior up to the strength at various levels of confinement [52]. Initial stiffness values were calculated from equations provided by the PFC3D manual and adjusted after testing the sample based on results for overall sample strength, elastic modulus, and Poisson’s ratio [44]. 22
Colorado School of Mines
A pervasive drawback of modeling rock failure with DEM is the underestimated angle of internal friction and the low compressive to tensile strength ratio. Fakhimi attempted to calibrate these behaviors by using a technique where the particle assembly is slightly overlapped at all contacts and then normal forces are zeroed [24]. It is thought that the increased contact frictional force in absence of the normal contact force would increase the overall internal friction angle. While the internal friction angle and compressive to tensile strength ratio both improved, the modified DEM still yielded unrealistic values. In 2007, Fakhimi and Villegas published a dimensional analysis of DEM micro-parameters that rein- forced the importance of the sample genesis pressure to the material failure envelope [25]. Koyama and Jing showed the effect of model scale and particle size on the macro behavior of the sample and outlined a method to determine the representative elementary volume for a given set of micro-parameters [51]. Cho et al. introduced the idea that by clumping particles together into irregular shapes, one can improve the simulation of failure behavior in terms of the failure envelope and tensile to strength ratio [15]. Yoon suggested that by selective design of experiment, Plackett-Burman in this case, one can optimize the micro-parameters using sensitivity analysis [91]. This method results in reliable parameter selection for rock materials within ranges not applicable to the study of coal and only up to the point of failure. Hsieh et al. demonstrated the effect that complex arrangements of various types of particles with particularly defined contact parameters can affect deformability and strength behavior [34]. Wang and Tonon produced a sensitivity analysis that developed equations relating micro-parameters to sample deformability and strength [87]. Shopfer et al. showed the effect of sample porosity and initial crack density on material behavior up to peak strength [78]. Up to this point no researchers had dealt with the calibration of the post-peak behavior until 2011 when Garvey and Ozbay offered a method to calibrate deformability, strength, and post-peak modulus [27]. Despite difficulties in calibration, it has been demonstrated in numerous papers that with proper micro-parameter calibration, realistic rock properties emerge [31, 58]. Some 23
Colorado School of Mines
of these properties are elasticity, fracturing, anisotropy with accumulated damage, dilation, strength increase with increased confinement, post-peak softening and hysteresis. Modeling rock behavior with DEM has three limitations worth noting. The measured Mohr-Coulomb friction angle is roughly half of its expected value, the Mohr-Coulomb failure envelope is linear, and the compressive to tensile strength ratio of the material is lower than the real rock [21, 67, 68]. It has been proposed that by introducing non-circular elements, the effects of these limitations are greatly reduced [15]. Many strategies are available to tailor a DEM model to most efficiently achieve its goal. Within bonded particle models, user defined contact laws have a notable effect on the overall behavior of the model. Resulting macro behaviors include time-dependent stress corrosion and sliding along pre-existing joints [58, 66]. Also, heterogeneity in rock can be modeled by enclosing groups of similarly strong or stiff particles with smooth contacts to create a larger grain [69]. The grain-based model is well suited for modeling cases that involve spalling or for instances where inter-granular and intra-granular cracking are pertinent features of the failure process. The method of modeling a rock mass by embedding a system of joints can be achieved in DEM by replacing bonds with smooth contacts along predetermined joint planes. The synthetic rock mass (SRM) approach was developed by Mas Ivars et al [58]. Pierce et al. showed that it satisfactorily predicted rock mass brittleness by validating fracturing in a case study block cave mine Pierce et al. [65]. The SRM approach has also been used to address scaling issues associated with using DEM to model rock masses. When DEM material is calibrated to intact rock properties, the well-known effect of strength degradation due to increased scale does not occur. Deisman et al. and Esmaieli et al. showed that an SRM model can simulate the scale dependency of macroscopic behaviors in a coal bed methane reservoir and an underground metals mine in Canada [20][22]. DEM models are well suited to model micro-mechanical behavior of rock such as notching [68]. Although, DEM is notorious for high run times, so it is unreasonable to construct large models out of relatively small particles. One reason is that equilibrating such large a system 24
Colorado School of Mines
CHAPTER 3 EVALUATION OF TWO DEM MODELS FOR SIMULATING UNSTABLE FAILURE IN COMPRESSION In order to simulate unstable failure using a numerical model, the post-peak behavior of the simulated material must have a softening characteristic. In this chapter, two different discrete element models are compared to determine which is better suited for simulating unstable failure in compression. The two models are described in detail then subjected to a series of compression tests. In each test, the same geometry of specimen is brought to failure under compressive loading. The suite of tests was chosen to investigate the effect of key loading conditions imposed on the DEM by in situ models of later chapters. Four different test procedures are used to investigate the behavior of each DEM. The first test is used to establish the so-called characteristic material behavior, which refers to the rock specimen deforming incrementally under gradually increased loading. The uniaxial compressive strength test, UCS, is used to establish characteristic material behavior for the purpose of this research. The other three tests that are used to investigate the effect of three separate loading conditions include: triaxial compressive strength test (TCS), elastic platen compression test (EPC), and loading rate compression test (LRC). The UCS and TCS model tests use a constant velocity boundary condition to load the specimen where no strain energy is available from the loading system to affect the specimen’s failure mode. To investigate the performance of DEM and also the failure modes of rocks, elastic end platens are placed on top and bottom of the specimen. In this series, the platens can store strain energy, which can be released during specimen failure in a gradual or sudden manner depending on the rock’s failure mode. In a specific test series called elastic platen compression, EPC, the platen stiffness is gradually reduced to observe the failure mode changing from being stable to unstable. Lastly, a test is implemented to investigate the 26
Colorado School of Mines
effect of loading rate on the characteristic material behavior of the DEM code. A series of UCS tests is conducted in which the loading rate is changed to a different value for each test. By comparing the resulting stress-strain curves, the effect of loading rate on elastic behavior, strength and post-peak behavior is analyzed. 3.1 Particle Flow Code in Two Dimensions (PFC2D) In this study, the DEM modeling studies are performed by using the commercially avail- able discrete element code PFC2D [43]. These codes allow for model customization via an embedded programming language called FISH. PFC2D comes with a collection of FISH functions that allow the user to accomplish complicated tasks with a moderate amount of background knowledge in discrete element modeling. The authors of PFC2D call the prein- stalled collection of fish functions the Fishtank. The Fishtank contains a series of functions that generates a discrete element model of a bonded rock-like material and performs tests to determine material properties. Templates are provided in the Fishtank that link together steps in a study such as material generation, testing, then data display. Templates can be run as examples or by providing custom inputs to the template, user specific tasks can be easily performed. Fishtank version 1-115 was used in this research. It is often necessary to modify the provided test procedures and functions for user specific purposes. Template files used in material generation, testing or function definition will be referenced along with customized inputs. Files containing modified test procedures or custom functions will be provided in the appendices. 3.2 Material Generation and Calibration This study utilizes the material generation procedure described in the PFC2D manual within the section entitled PFC Fishtank. Circular elements, or particles, are created within avesselandelementradiiarevarieduntilatargetisotropicstressisachieved. Then“floating” particles (particles with a number of contacts below a predefined threshold) are deleted and theelementsarebondedusingeitherparallelorcontactbonds. Here, periodicboundariesare 27
Colorado School of Mines
used to create a square vessel, resulting in square blocks of material that can be connected seamlesslytocreatealargerassembly. Thissocalledpbrickmethod isdecried indetail inthe PFC2D manual under the topic of adaptive continuum/discontinuum (AC/DC) logic. For material generation and specimen assembly, the Fishtank template acdc-2d is used. The file acdc-pbr.dvr is used to generate the pbrick, and the file acdc-bv.dvr is used to assemble pbricks into the specimen. The behavior of the generated material is largely determined by behavior of particle con- tacts, called the contact model, and the type of bond used to attach the particles to one another. The combination of contact model and bonding scheme make up the constitutive model of the discrete element model. Both the contact model and the bond are defined by a set of microparameters governing stiffness and strength properties. For this study an appropriate constitutive model must be chosen and calibrated for the purpose of simulating unstable failure. It is necessary that the constitutive model is capable of simulating a soft- ening post-peak characteristic. Two constitutive models available to PFC2D users that are capable of simulating a softening post-peak characteristic are the parallel bonded particle model and the displacement softening model. Figure 3.1 shows the components of a discrete element model constitutive model. The contact model can contain a bond in the form of a contact bond and a parallel bond can be added. Elastic contact behavior in both constitutive models described below is the same. Figure 3.2 shows a schematic of the components of stiffness between two particles that are in contact. Each particle is assigned a value for normal stiffness, k , and shear stiffness, k . n s Force between the particles is calculated using Hooke’s law, Equation 3.1, where i represents values associated with either the normal or shear direction. The stiffness coefficient,K , is i calculated by assuming the element stiffnesses to act in series. Therefore, contact stiffness is calculated using Equation 3.2, where A and B represent the particles involved in the contact. F = K dx (3.1) i i i 28
Colorado School of Mines
also fail in direct tension or due to bending. After a bond breaks, newly detected contacts obey the laws of the contact model. Failure in the parallel bond is dependent upon the geometry of the contacting particles and the cross sectional area of the bond. According to beam theory, the maximum shear and tensile stresses that can exist in the bond are Equation 3.3 and Equation 3.4 respectively. F s τ = (3.3) max A F M R n σ￿ = − | | (3.4) max A − I F and F are the shear and normal forces on the bond, M is the moment, and A = 2R, the s n cross-sectional area of the bond with unit thickness and I is the bond moment of inertia. In this study, success of the calibration of the BPM relied on achieving a softening post- peak characteristic. Until recently, no information existed in the literature on calibration of post-peak behavior in bonded DEM models. Garvey & Ozbay [27] introduced an iterative calibration method that uses an elitist-selection, genetic algorithm and an unconfined com- pression test to discover a set microparameters that achieve a target stress-strain behavior. For this research, the method was modified to utilize a two dimensional specimen assembled using the pbrick method. Table 3.1 shows the resulting microparameters necessary to re- produce the BPM material used in this study. Parameters not listed are set to the default values, which can be found in the PFC manual. 3.2.2 Displacement Softening Model The displacement softening model (DSM) in PFC2D is a constitutive model composed of a bonded contact model without a parallel bond. Figure 3.3 shows the force versus displacement curve for the DSM. The DSM behaves as described above in the elastic region. When initial contact bond strength, Fn, is reached, the contact behavior begins to follow a c linear strength softening curve. If unloaded during softening, the bond can rebound along the elastic path. When the user defined plastic displacement limit is reached, U , the pmax 30
Colorado School of Mines
Figure 3.3: Displacement-softening constitutive model behavior bond is inactive. The contact yields in tension when the resultant contact force, Equation 3.5, is greater than the contact strength, Equation 3.6. Here, the plastic displacement is in the direction of the resultant force. F = Fn2 +Fs2 (3.5) ￿ 2α 2α F = 1 Fn + Fs (3.6) cmax − π · c π · c ￿ ￿ If the contact is in compression, failure can occur due to shear. The strength of the contact, Equation 3.7, is dependent upon the coefficient of friction and the normal force, and plastic displacement is in the shear direction. F = µ Fn +Fs (3.7) cmax | | c During yield, the elastic displacment increments in the normal and shear directions are a function of only the portion of resultant force up to the strength and the contact stffiness, Equation 3.8, where k = n,s, and the plastic displacement increments are given by Equation 3.9. Fk = Kk Uk (3.8) ￿ ￿ e Uk = Uk Uk (3.9) ￿ p ￿ −￿ e 32
Colorado School of Mines
The calibration of the DSM can be performed iteratively by the user in a short period of time due to the intuitive response of macro behavior due to changes in microparame- ters. Material generation is performed once. The resulting particle assembly is used to test various combinations of microparameters. The UCS test is used to determine the stress- strain behavior. First, desired elastic behavior is achieved by varying contact stiffness. Then the plastic displacement limit is varied to change the post-peak softening behavior of the specimen. By increasing U , the post-peak modulus decreases and the strength of the pmax specimen increases. To achieve an appropriate UCS and post-peak modulus, the tensile and shear strengths of the contact are decreased. Some iteration is necessary to achieve the de- sired behavior. The following section presents results for UCS tests on the calibrated BPM and DSM. 3.3 Unconfined Compressive Strength Test (UCS) The DEM material properties necessary for study of unstable failure in compression can be found using a simulation of compressive strength tests. The UCS is used in this research to calibrate the DEM by approximating a target set of characteristic material properties. The target characteristic behavior is representative of an in situ western United States coal. The UCS test specimen is a one meter wide by two meter high assembly of two blocks of either DSM or BPM material. The specimen is loaded passed the point of failure until an adequateassessmentofpost-peakbehaviorcanbedetermined. TheFishtanktemplate,direct tension test with reversed platen displacement is used to perform this test. It is necessary to use “grip particles” in order to compress the specimen due to the roughness of the specimen ends. Table 3.3 shows UCS test template filenames and necessary inputs. For the DSM, all contacts are assigned the displacement softening contact model with microproperties listed in Table 3.2 after the specimen is restored. Stress in the sample is calculated by summing particle forces in each grip respectively, dividing by the width of the specimen, and then averaging the two grip stresses. Strain is calculated by determining the change in specimen height using grip displacement. The test is terminated in the post-peak region when the 34
Colorado School of Mines
measured axial stress in the sample is lower than half of the UCS. After half of the peak stress has been dissipated due to failure, a sufficient range of post-peak softening can be observed in order to quantify the material post-peak modulus. Table 3.3: UCS test parameters Figure 3.4 shows the stress-strain curves from UCS testing on the BPM and DSM. Ta- ble 3.4 lists the elastic properties and post-peak modulus of each model and for a target material. The target material is set to approximate an in situ, western United States coal. Both curves reflect a post-peak softening characteristic that is approximately equal in mag- nitude to the pre-peak modulus. Each material is within an acceptable range from the target material properties. While it is beyond the scope of this thesis, improvements in the calibration of both the BPM and DSM warrant further investigation. This would in- volve an in-depth look at the genetic algorithm used to calibrate the BPM or comprehensive sensitivity analysis of DSM microparameters respectively. Table 3.4: DEM and target characteristic material properties 35
Colorado School of Mines
Figure 3.4: UCS stress-strain curves for the BPM and DSM 3.4 Biaxial Compressive Strength Test (BCS) Consider an axially loaded rock specimen shown in Figure 3.5 under a constant confine- ment stress σ￿ and deviatoric stress σ￿ . According to the Coulomb failure criterion, failure 3 1 will occur along a plane oriented at an angle β. In this failure criterion, the strength is linearly proportional to the normal stress on the failure plane. Equation 3.11 shows the relationship between the shear strength, τ, the cohesion, c, the normal stress, σ￿ , and the n internal friction angle, φ. τ = c+σ tanφ (3.11) n Similar to what is observed in the laboratory, when confinement stresses are applied to bonded discrete element models, strength of the material increases. Strength increase in the presence of a confining stress is an emergent property of the DEM that is not directly calibrated. By performing BCS tests on the specimens described above, the internal friction angle of the DEM material can be calculated and compared to that of real rock. 36
Colorado School of Mines
Figure 3.5: Coulomb shear failure plane and stresses A constant confining pressure is applied to the specimen using the so called spanning chain algorithm included in the Fishtank. The spanning chain algorithm detects particles that lie along a boundary and applies forces to the particles to simulate a constant boundary pressure. The advantage of using this technique is that it allows for displacements on the boundary preventing stress concentrations that would be caused by a rigid boundary such as a wall. Custom functions were introduced in order to adapt the algorithm to the UCS specimen. These custom functions are shown in Listing A.1. The test template used for the UCS tests above was modified to include the spanning chain and pressure functionality. This file is shown in Listing A.2. The test parameters listed in Table 3.3 are also used in this set of tests. Figure 3.6 shows a screen shot of the specimen. The yellow marks are the disk shaped elements, and the red circles attached by black lines make up the spanning chain used to apply the confining stress. BCS tests were performed using 1, 2, 4 and 6 MPa confining stress on both the BPM and DSM DEMs. The tests were terminated post-failure when the axial stress decreased to ninety percent of the strength. Figure 3.7 and Figure 3.8 show stress versus axial strain curves for the BPM and DSM. The dotted lines are the horizontal stresses and the solid lines are the axial stresses. The dotted lines show that the spanning chain confinement method 37
Colorado School of Mines
provided more consistent confinement stress during the BPM tests than in the DSM tests. Horizontal stress increased approximately 1 MPa throughout each of the DSM tests and Figure 3.7: BPM confined compression test stress versus strain curves only approximately 0.1 MPa for the BPM tests. Peak axial stress, σ￿ , and the prescribed 1 confinement stress, σ￿ , were used to calculate the shear and normal stresses on the failure 3 plan in order to determine the friction angle. Figure 3.9 shows a shear stress versus normal stress plot using each test result for both discrete element models. The plot also shows the internal friction angle for each model. The friction angle for the DSM is significantly higher than the BPM. The friction angle of real coal is approximately 30 degrees. So, the BPM will simulate the coal as being weaker under confined conditions than reality while the DSM will simulate the coal as being stronger under confined conditions. Considering the increase in horizontal stress during the DSM tests, a slightly lower friction angle can be attributed to the DSM material. By considering the horizontal stress at failure rather than the prescribed confining stress as the true confining stress the friction angle of the DSM material decreases slightly to 42 degrees. The effect of confining stress on strength will determine the strength of the rock during compressive 39
Colorado School of Mines
Figure 3.8: DSM confined compression test stress versus strain curves failurebutitwillnoteffectthefailuremodeoftherock. So, whileastudyonthemechanisms underlying unstable failure will not be directly affected by the friction angle the BCS test provides supplementary information on the accuracy of DEM simulation of rock behavior. 3.5 Elastic Platen Compression Test (EPC) Byvaryingthestiffnessoftheloadingsystem, theDEMspecimencanbefailedinastable or unstable mode of failure. A mechanical coupling method is used to fail the DEM models under elastic platens simulated using the finite-difference, continuum code FLAC2D. Here, first the coupling method will be explained and then results from the compression tests will be presented and discussed. The mechanical coupling of FLAC2D and PFC2D relies on the exchange of gridpoint velocities and particle forces at the coupling boundary via a socket connection like that used in TCP/IP transmission over the internet. The coupling boundary consists of a layer of discs in the PFC model which overlaps the FLAC grid and the gridpoints associated with the FLAC grid on that boundary. The particles on the boundary are called control particles and the gridpoints that are on the FLAC coupling boundary are called control gridpoints. 40
Colorado School of Mines
Figure 3.9: Shear stress versus normal stress results from the confined compression tests Figure 3.10 shows a diagram of the coupled model calculation cycle. The red arrows indicate communication between FLAC2D and PFC2D. PFC2D uses Fish functions provided in the Fishtank to update boundary conditions before every cycle. With these functions, PFC2D uses control gridpoint velocities to calculate and then apply velocities to control particles. Following a calculation cycle in PFC2D, updated forces on the control particles are used to calculate and then send control gridpoint forces to FLAC2D. Following a calculation cycling in FLAC2D, updated control gridpoint velocities are sent back to PFC and the coupled model cycle repeats. Each control particle is associated with a FLAC2D segment, which is defined by two control gridpoints Figure 3.11 shows a diagram of two control gridpoints, 0 and 1, and one control particle, P. In order to apply the velocity boundary condition to PFC2D, control particle velocity is determined by linear interpolation of control gridpoint velocities. The relationship between control particle velocity, v, and control gridpoint velocities, v and v 0 1 is shown using Equation 3.12 v(ξ) = v +ξ(v v ) (3.12) 0 1 0 − where, 41
Colorado School of Mines
file, [Listing A.11]. PFC2D acts as the master program and controls coupling by issuing commands to cycle the two codes one calculation step at a time. This is done in the main driver file, along with loading default functions for control of the coupling boundary and custom functions used to initialize the system, [Listing A.12 and Listing A.13]. Functions used for measuring model state variables and recording data are also called from the driver file, [Listing A.14]. Figure 3.14 shows the test geometry and boundary conditions for the coupled simulation. The upper and lower platens are moved inward at the velocity used for calibration. It was expected that the specimen would fail in a stable manner when the loading system modulus washigherthanthanthespecimenpost-peakmodulusandunstablewhentheloadingsystem modulus is lower than the specimen’s post-peak modulus. A series of tests for each DEM were run in which only the platen elastic modulus was varied. A range of moduli were chosen for each model so as to clearly depict the transition from stable to unstable failure as platen moduli decreases. Figure 3.14: Mechanically coupled compression geometry and boundary conditions 45
Colorado School of Mines
Figure 3.15 and Figure 3.16 show stress strain curves for the EPC simulations using the BPM and DSM respectively. BPM simulations were conducted using ten different platen moduli: 0.5, 1, 2, 3, 5, 10, 15, 20, 35, and 50 GPa. DSM simulations were conducted using seven different moduli: 1, 1.5, 2.5, 5, 10, 20, and 35 GPa. In both plots, the unconfined compressive strength stress-strain curve is included to provide a reference to the calibrated characteristic material behavior. Both plots show that for tests with stiff platens as com- pared to the material post-peak modulus the post-peak behavior follows the slope of the characteristic material curve from the UCS test. Tests with soft platens show a deviation in specimen post-peak behavior from the characteristic material post-peak behavior. Figure 3.15: BPM coupled simulation stress-strain curves A determination of failure stability is possible by comparing the assigned loading system modulus and the specimen modulus determined from observed post-peak behavior. The specimenpost-peakmodulusisdeterminedbymeasuringthepost-peakslopeofthespecimen stress-strain curve. During stable failure, the specimen fails according to its characteristic material properties. So, the post-peak modulus equals the calibrated characteristic post- peak modulus, E . During unstable failure, all load bearing capacity of the specimen is pp lost. Without the resistance of the specimen, the platens rebound according to the elastic 46
Colorado School of Mines
Figure 3.16: DSM coupled simulation stress-strain curves properties. So,itappearsthatthespecimenpost-peakmoduluschangestoequalthemodulus of the loading system. Table 3.5 and Table 3.6 show the specimen post-peak modulus measured from each test and the strength of each specimen for the BPM and DSM respectively. The characteristic material behavior is also included, labeled as “UCS Grip”. The shaded area indicates tests that resulted in unstable failures. Figure 3.17 shows a scatter plot of E and E from pp plat each EPC test, the measured specimen post-peak modulus and the assigned loading system modulus respectively. The vertical lines are E from the “Grip UCS” tests, which are the pp calibrated specimen post-peak moduli for the BPM and DSM. Table3.5andTable3.6shownumericallythatwhenE becomeslargerthanthecharac- plat teristicE theEPCE beginstodeviatefromE . ThisisshowngraphicallyinFigure3.17 pp pp plat as a asymptotic trend to the right of the vertical lines marking characteristic E . Ideally, pp each model should transition immediately to stable failure as loading system modulus in- creases beyond this value. The trend in the DSM post-peak modulus values indicate that a fairly sharp transition from unstable to stable failure occurs. This is reflected in the consis- tency of values in Table 3.6 that are not shaded. On the other hand, the post-peak behavior 47
Colorado School of Mines
Figure 3.17: Loading system and specimen post-peak moduli in EPC tests of stable BPM failure has a range of values. This is shown in Figure 3.17 as a slow transition toward stability in terms of EPC E and inconsistent values for high moduli tests. The pp stability transition behavior for the BPM indicates that other factors are in effect. This so called quasi-stable behavior in the BPM tests could be caused by micro-mechanical behavior that is not visible in this type of analysis. For example variation in failure progression could lead to a change in the post-peak characteristic during the process of the failure, leading to a E different from the characteristic E or E . This effect is not noticeable in the DSM pp pp plat results. Additionally, the DSM has consistent E for each stable test while the BPM E for pp pp the high modulus tests (more likely to be stable than the quasi-stable moduli tests) varies slightly and remains below characteristic E . pp In the EPC tests, the effect of loading system stiffness on the mode of failure of two DEMs was investigated by changing the modulus of elasticity of the loading platens between different tests. The stability transition behavior of the two DEM models is important to the current research in so far as we are able to reliably detect unstable failure. According 49
Colorado School of Mines
to the stability transition behavior presented above, in a situation where the loading system stiffness is similar to the characteristic material post-peak stiffness, the DSM more clearly distinguishes between stable and unstable failure. The BPM behaves in a quasi stable manner when the loading system stiffness is similar to the characteristic material post-peak stiffness. The quasi-stable behavior is more difficult to discern as stable or unstable because the post-peak behavior is equal to neither the characteristic post-peak behavior nor the loading system stiffness. A dffierence arrises between the DSM and BPM model behavior during the EPC test when examining the strength. Figure 3.16 show a change in specimen strength for the three softest DSM EPC tests while in the strength in the BPM EPC tests remains fairly consistent. Table 3.5 and Table 3.6 show the strength of each specimen for each test. These results indicate that the strength of the DSM material decreases when subjected to unstable loading conditions. The BPM model does not experience this change. A possible reason for this reduction in strength could be the sudden onset of localized, unstable crack growth. Contact bond failure associated with material yielding could be prematurely accelerated if a large amount of stored strain energy is available. In the case of soft, elastic loading systems, this is possible. Additional work would be needed to confirm this hypothesis. Local measurements of material stiffness could provided evidence for this micromechanical process. Regardless, when using this DEM, a reduction in strength of the DSM material should be expected when loading system stiffness is less than material post-peak stiffness. 3.6 Loading Rate Compression Test (LRC) In the final test of DSM and BPM behavior in compression, the UCS test is revisited with different loading rates, referred to as LRC. Four loading rates are chosen, including the loading rate used for the previous tests. As in the UCS test, grip particles are moved inward to load the specimen. Vertical stress and strain measurements are taken using the grip particles and the stress-strain curves for each test are compared. 50
Colorado School of Mines
Figure 3.18 shows the stress strain curves for four LRC tests of the BPM. Figure 3.19 shows the stress-strain curves for the same tests using the DSM. ?? and Table 3.8 show the loading velocity (v), the post-peak modulus (E ) and the strength (σ ) for the BPM pp c and DSM tests. Due to the shape of the DSM curves in the post-peak region, some liberty had to be taken to choose a representative section of each curve in Figure 3.19 to describe the post-peak behavior as linear. The straight portions of the curves ranging from 50% to approximately 80% of peak strength were used to calculate E . pp Table 3.7: LRC BPM loading velocity, post-peak modulus and strength values Table 3.8: LRC DSM loading velocity, post-peak modulus and strength values Figure 3.18 and Figure 3.19 show that there is no noticeable effect of loading rate on the elastic region of the stress-strain curve. As loading rate increases, the strength of both DEMsincreases. TheeffectismoreprominentintheDSMmodelthanintheBPMmodel. In the post-peak region, a change in material behavior emerges as the loading rate is changed. Figure 3.18 and Table 3.5 show that the E of the BPM increases as the loading rate is pp decreased. Thechangeinpost-peakstiffnessissodrastic,thatthematerialentirelyloosesthe calibrated post-peak softening characteristic. Figure 3.19 shows that the change in loading rate has an effect on DSM post-peak stiffness. As loading rate decreases the stress-strain curve reveals abrupt changes in vertical stress as the material fails. Despite changes to the 51
Colorado School of Mines
BCS tests showed a stark difference in the effect of confinement on strength between the two models. The DSM has an internal friction angle that is higher than the approximate internal friction angle of coal and the BPM’s internal friction angle is too low. The low internal friction angle of the BPM is a known shortcoming of the model [68]. The higher than desired internal friction angle of the DSM is an unexpected result. EPC test results showed that failure stability in the DEMs could be determined by comparing loading system and material post-peak stiffnesses. Using this comparison as an indicator of unstable failure, when platen modulus increased beyond the material post-peak stiffness the DSM transitioned more quickly than the BPM from unstable to stable failure. LRC tests revealed changes in post-peak behavior in both models. In the BPM, as loading rate decreased the post-peak stiffness increased drastically. While in the DSM, the slowest loading rate resulted in abrupt changes in stress during failure, but overall the material retained its post-peak softening characteristic. The results from the EPC and the LRC tests provide important model behavior charac- teristics that suggest that the DSM is more appropriate for the studies of unstable failure in underground coal mining. The DSM’s sudden transition from unstable to stable failure seen in Figure 3.17 indicate that the expression of unstable failure is more ambiguous in models using the BPM. It was important in this chapter to show that the chosen DEMs could satisfactorily simulate unstable failure, but in work in later chapters it will be shown that detecting instances of unstable failure in larger models is crucial to studying the effects of local mine conditions on instability. Inalaterchapter, anundergroundminemodelwillbeintroducedthatusesinsitustresses to load the PFC2D coal. Rather than applying a consistent loading velocity, gradual mining steps will redistribute stresses in the model resulting in increased load on material near the excavation. The velocity by which the model applies the load is not controlled by the user and will vary from the velocity used to calibrate the PFC2D material. Figure 3.18 showed drastic changes in the post-peak softening characteristic of the BPM material as loading 54
Colorado School of Mines
CHAPTER 4 INDICATORS OF UNSTABLE COMPRESSIVE FAILURE IN DEM COAL STRENGTH TESTS This chapter is concerned with characterizing the expression of unstable compressive fail- ure in the displacement softening model (DSM). Various measurements of DEM behaviors can be used to indicate whether failure is unstable or stable and give a measure of the degree of failure instability. These calculated values are called stability indicators. Nine stability indicators are explained and employed here. They are damping work, maximum instanta- neous kinetic energy, cumulative kinetic energy, maximum instantaneous mean unbalanced force, cumulative mean unbalanced force, maximum instantaneous maximum unbalanced force, cumulative maximum unbalanced force, contact softening, and the number of broken contacts. The nine indicators are first applied to a simulation of a laboratory test, the elastic platen strength (EPC) test from the previous chapter. Since both stable and unstable failure occurred within the series of EPC tests, the behavior of each indicator during unstable and stable failure is observed. The indicators are then compared to one another to determine suitability for tracking unstable failure. It is useful to apply the stability indicators to DSM models of various sizes so that in-situ geometries can be investigated. Therefore, the nine stability indicators are applied to a series of slender pillar compressive strength (SPCS) tests. During these tests, pillars of various sizes are failed by loading systems with different stiffness to encourage stable and unstable failures. Failure stability is determined using a comparison between loading system stiffness and post-peak behavior similar to that used on the EPC tests. The effect of model size on stability indicator performance is observed and the indicators are once again compared for suitability in tracking unstable failure. 56
Colorado School of Mines
Due to the change of model size it is also beneficial to observe the spatial distribution of damage and damage intensity in the pillar. Additional analysis is performed on the SPCS tests to observe the spatial distribution of contact softening and the damping work due to failure. A grid based measurement technique used to track the two indicators is explained and then the correlations between model damage and failure stability are discussed. 4.1 Description and Calculation of Stability Indicators in DEM Compressive Failure Each of the stability indicators are calculated using PFC2D particle and contact state information. This section provides details on how each indicator is calculated and references custom FISH functions that facilitate the calculations. 4.1.1 Damping Work The PFC2D model uses a damping mechanism to dissipate kinetic energy, so that a steady state solution may be arrived at within a reasonable number of calculation steps. The damping mechanism applies force to particles undergoing acceleration in the direction opposite that of the particle’s motion. Equation 4.1 shows the damping force applied to each particle: →−F = α →−F (sign v ) vˆ (4.1) d unbal →− ·− · ￿ ￿ ￿ ￿ where →−F d is the damping force, α is a￿n dime￿nsionless coefficient, →−F unbal is the unbalanced force on the particle, and v is the particle velocity. The coefficient, α, is used to define the →− level of damping. A value of 0.7 is used in all of the simulations in this thesis. This is the value recommended by the authors of PFC2D for quasi-static conditions. During failure, the damping mechanism applies larger forces to the model in order to stabilize the failure process. Over a calculation timestep, dt, the damping forces perform a quantifiable amount of work that can be summed over the entire model. Damping work is summed over each degree of freedom, i, over all particles, N, from timestep t to t . The i f work is summed over the interval of failure. Equation 4.2 is the work done for translational 57
Colorado School of Mines
motion where dx is the incremental translational displacement and Equation 4.3 is work →−i done for rotational motion where −M→ is the damping moment and d r is the incremental d →− rotation. Translational damping work and rotational damping work are summed to obtain the total damping work. The functions responsible for calculating the total damping work are given Listing C.15. The function called param loop bp loops through all of the particles in the assembly and pfc wd calculates the damping work on each particle. tf N 2 W = →−F dx (4.2) dtrans ￿t=ti n ￿p=1 ￿i=1 ￿￿ di ￿·| →−i | ￿np,t ￿ ￿ tf N ￿ ￿ W = −M→ d r (4.3) drot ￿t=ti n ￿p=1￿￿ d ￿·| →− | ￿np,t ￿ ￿ ￿ ￿ 4.1.2 Maximum Instantaneous Kinetic Energy During the simulation, the kinetic energy of the model is calculated by summing the rotational and translational kinetic energies of all the particles for a single timestep. Equa- tion 4.4 and Equation 4.5 are the equations for rotational and translational kinetic energy respectively, and the total kinetic energy is given by Equation 4.6: 1 KE = Iω2 (4.4) rot 2 2 1 KE = mv2 (4.5) trans 2 i i=1 ￿ KE = KE +KE (4.6) rot trans where, I = 1/2mr2, and ω is the rotational velocity. KE is calculated every step as an instantaneous value. The failure stability and intensity of failure should be reflected in the velocity of particles. Therefore, the maximum value of instantaneous kinetic energy during failure is used as a stability indicator because it provides information on the velocity of particles. 58
Colorado School of Mines
4.1.3 Cumulative Kinetic Energy The cumulative kinetic energy, KE , can be determined from the record of instantaneous c kinetic energy by summing the instantaneous kinetic energy over the time interval of failure, from timestep t to t as in the case of damping work, Equation 4.7. The kinetic energy i f during the entire duration of failure reflects the stability and intensity of the failure in its entirety in terms of particle velocity. tf KE = KE (4.7) c t ￿t=ti 4.1.4 Maximum Instantaneous Mean Unbalanced Force The instantaneous mean unbalanced force, F , is the average of the absolute values of µ the out of balance force components for each particle and is calculated using Equation 4.8. The maximum instantaneous mean unbalanced force is the largest value of mean unbalanced force of all the timesteps during the failure interval. The mean unbalanced force provides a measure of the level of instability because unbalanced forces are lowest when the model is near static equilibrium. By taking the mean of all unbalanced forces the effect of outlying values is minimized. F F = (4.8) µ N N 3 F = ( |F unbali|) np n ￿p=1 ￿i=1 4.1.5 Cumulative Mean Unbalanced Force The cumulative mean unbalanced force is calculated in the same way as the cumulative unbalanced force. Equation 4.9 is used to calculate the cumulative mean unbalanced force. tf F = F (4.9) µc µt ￿t=ti 59
Colorado School of Mines
4.1.6 Maximum Instantaneous Maximum Unbalanced Force The maximum unbalanced force is the unbalanced force of greatest magnitude during a timestep. This value is determined by a PFC2D intrinsic FISH function instantaneously and during each step. The maximum instantaneous maximum unbalanced force gives a measure of the intensity of failure using the element furthest from static equilibrium. F = max ( F ) : i = 1,2,3 & n = 1,...,N (4.10) max | unbali| np p p ￿ ￿ 4.1.7 Cumulative Maximum Unbalanced Force The cumulative maximum unbalanced force is determined similarly to the cumulative mean unbalanced force. This indicator provides a measure of failure intensity by finding the degree of freedom with the largest amount of applied force each step. The maximum unbalanced force should be larger the more unstable the failure is. tf F = F (4.11) maxc maxt ￿t=ti 4.1.8 Contact Softening The DSM is a softening contact model, as explained in detail in the previous chapter. The contact begins to soften once the initial strength of the contact is reached (Figure 3.3). The contact bond is inactive once the softening limit is reached. Before the plastic displace- ment limit is reached the amount of softening, U , can be observed using an intrinsic FISH p command to access a contact state variable called the contact softening ratio, U . The rat contact softening ratio is the amount of contact softening divided by the plastic displacement limit, Equation 4.12. 60
Colorado School of Mines
U = U /U (4.12) rat p pmax The value of U becomes unity at maximum softening. By summing U over all of the rat rat contacts in the model, a measure is made of the contact softening due to compressive failure. The sum of softening ratios is calculated incrementally. The functions responsible for cal- culating the sum of softening ratios are located in Listing C.15, where param loop cp loops through all of the contacts in the PFC2D assembly and pfc sof retrieves contact state in- formation. The amount of contact softening, the indicator used in the following analysis, is determined by multiplying the softening ratio sum by the plastic displacement limit. This yields a value for contact softening in units of meters. Equation 4.13 is the amount of contact softening, where C is the total number of contacts. C U = U U (4.13) rat pmax · ￿ ￿ n ￿c=1 4.1.9 Number of Broken Contacts In a DSM contact, the contact is deemed broken once the plastic displacement limit, U , is achieved. One way of assessing damage in the DEM assembly is by tracking the pmax number of contacts that have broken. The number of broken contacts is determined using the variable sof numbroke shown in Listing C.15. 4.2 Stability Indicator Results in EPC Tests Each of the indicators explained above are used here to describe the failure in elastic platen strength (EPC) tests presented in the previous chapter. The trends of indicators are shown by means of scatter plots of the indicator values versus time step. Values of indicators are determined from line plots of the indicator, so each point on the scatter plots represent indicator magnitude for an individual test. Figure 4.1 shows the instantaneous and cumulative kinetic energy during the EPC test with 5 GPa platens. During the loading phase of the test, the cumulative kinetic energy 61
Colorado School of Mines
Figure 4.2: Accumulated damping work during the failure interval in EPC tests small increase with 2.5 GPa platens the damping work begins a sharp increase for the two softest platens. Figure 4.3 shows the maximum instantaneous kinetic energy during failure on a semi-log plot. While the trend seen in damping work of increasing indicator with decreasing stability of failure is present, the kinetic energy shows an even more drastic increase for the most unstable cases. The three highest platen moduli exhibit consistent maximum instantaneous kinetic energy, suggesting that this indicator may be particularly useful in identifying stable failure. The maximum instantaneous mean unbalanced force is shown in Figure 4.4 on a semi-log plot. The results for this indicator are similar to the maximum instantaneous kinetic energy in that the most unstable failure has a significantly higher value than the next softest test, and the values for the most stable tests are very consistent. The maximum instantaneous maximum unbalanced force is shown in Figure 4.5. The trend in maximum instantaneous maximum unbalanced force is similar to kinetic energy and mean unbalanced force although there exists some irregularity in the value for the stable failures. 63
Colorado School of Mines
Figure 4.5: Maximum instantaneous maximum unbalanced force in EPC tests Like damping work, cumulative values for kinetic energy, mean unbalanced force, and maximum unbalanced force were normalized with respect to the stress drop during failure. Figure 4.6 shows the cumulative kinetic energy. The cumulative kinetic energy is consistent for stable failures and increases as failure stability decreases. In Figure 4.7, the cumulative mean unbalanced force generally shows the expected trends for stable versus unstable failure. Although, outlying results for the unstable failures using 2.5 and 5 GPa platens indicate that variability in mean unbalanced force in unstable failures can occur and caution should be exercised when using this indicator. Figure 4.8 shows the cumulative maximum unbalanced force. The cumulative maximum unbalanced force is fairly consistent for all tests with the exception of the most unstable failure. Therefore, it does not reliably distinguish between stable and unstable failures. ContactsofteningduringthefailureintervalforeachEPCtestisshowninFigure4.9. The amount of softening is normalized with respect to the stress drop during the failure interval for each test respectively. The amount of contact softening remains consistent for the most stable failures. For unstable failures the amount of contact softening exhibits no particular trend as the two most unstable failure result in the most extreme cases of softening. 65
Colorado School of Mines
Figure 4.8: Cumulative maximum unbalanced force in EPC tests The number of broken contacts for each tests is shown in Figure 4.10. The number of broken contacts is also normalized against the stress drop in each test. The number of broken contacts also suggests that stable failures have lower, consistent values while the as the failure becomes more unstable the value increases. Although, the number of broken contacts increases slightly as the elastic modulus increases. 4.2.1 EPC Indicator Results Discussion With the exception of contact softening, each of the indicators utilized for the analysis of failure stability in EPC tests exhibit similar trends for the EPC tests. Each of the indicatorsshowshighvaluesforunstablefailuresanddecreaseasstabilityoffailureincreases. Although, the number of broken contacts exhibits an increase with increasing platen elastic modulus. The damping work, maximum instantaneous mean unbalanced force, maximum instantaneous kinetic energy and cumulative kinetic energy appear to be suitable indicators for tracking unstable failures. For each of these indicators, consistent values are measured for stable failures and as failure stability decreased, the indicator likewise increased to provide a qualitative measure of failure stability. 67
Colorado School of Mines
Trends in cumulative values for kinetic energy, mean unbalanced force and maximum unbalanced force are helpful in describing failures in that the cumulative value contains information for the duration of the failure rather than a single calculation step. Each of these cumulative values performed with various levels of success in distinguishing between stable and unstable failures. The cumulative kinetic energy performs well in distinguishing between stable and unstable failure while the cumulative maximum unbalanced force looses the expression of instability for all unstable failures except for the 1 GPa test. The mean unbalanced force should be used with caution as there is some variability in the values of unstable failures. Although, additional work into methods of normalization may reveal the expected trend. Contact softening does not increase with decreasing failure stability but the variability in the value increases. Additional analysis could possibly reveal a trend similar to the other indicators, but is not pursued further in this study. Rather, since contact softening is a result of failure in contact bonds, it can be used to identify the locations and extent of damage in the model. This application will be applied in the following pillar tests. 4.3 Slender Pillar Compressive Strength (SPCS) Test Description In underground mining, both the material properties and the dimensions of the mine affect the loading system stiffness and consequently the failure mode. In this section, a series of slender coal pillars are failed under an elastic loading system. The stiffness of the loading system is varied by changing the modulus of elasticity of the loading system and also the size of the pillar. A total of nine tests are conducted, failing three pillar sizes under three separate loading systems of various elastic moduli. The pillar height is kept constant and the width is changed to produce three different sized pillars. The pillars are described by the ratio of pillar width to pillar height. Pillars are constructed of width to height ratios one, two, and three. Each pillar size is failed with a 5 GPa, 20 GPa, and 35 GPa loading system. Failure stability is determined by comparing the loading system stiffness to the pillar post-peak stiffness. Then, the performance of the nine stability indicators is assessed 69
Colorado School of Mines
for stable and unstable pillar failures. 4.3.1 SPCS Geometry and Boundary Conditions Figure 4.11 shows a schematic depicting the geometry and boundary conditions of the slenderpillartests. TheFLAC2Dgridiscomprisedofafineinnergridandacoarseoutergrid in order to capture the forces and stresses at the resolution of the PFC2D model and to save memory. The grid input file for the width to height ratio one pillar is shown in Listing C.16. The grid is expanded for the larger pillar tests in the horizontal direction by adding a proportional amount of elements. The same FISH functions are used to facilitate coupling as with the EPC tests, only the coupling boundary segment list as seen in cpf EPC.fis must be changed. The model has a symmetric boundary condition along the vertical edges, simulating an infinite series of identical pillars. The width of the excavation is kept constant for each pillar size, andthewidthofthemodelischangedinaccordanceonlytothepillarwidth. Themodel is first equilibrated with the entire FLAC2D nulled region filled with PFC2D elements. Then the entries are ‘excavated’ by deleting the elements within three meters of the left and right boundaries of the model. After a subsequent equilibration stage, the coal pillar, modeled in PFC2D, is loaded under an increasing compressive load by a constant velocity boundary condition applied to the upper and lower most boundaries of the model. 4.3.2 Local Mine Stiffness Calculation Each pillar failure is determined to be stable or unstable based on a comparison of the local mine stiffness during failure and the post-peak pillar stiffness. Following from the laboratory tests above, if the pillar post-peak stiffness is equal to the unloading local mine stiffness then the failure is considered unstable. The stiffness of the loading system is measured using only the tributary area above and below the pillar width. From these calculations, a local mine stiffness measurement is made for each test by assessing average pillar vertical reaction force on the surrounding mine and average pillar-mine boundary 70
Colorado School of Mines
Figure 4.11: Slender pillar test geometry and boundary conditions vertical displacement. The stiffness of the roof and floor can be determined individually using the the equation for force on a spring, Equation 4.14. k =∆ F /∆D (4.14) P ∆F is the change in force exerted on the roof or floor by the pillar and ∆D is the change P in displacement in the roof and floor respectively. ∆D is defined as the compression of the tributary area averaged along the width of the pillar. The pillar reaction force is calculated using average pillar stress and the cross sectional area of the pillar (P x 1m). W Figure 4.12 is a conceptual illustration of typical pillar behavior trends versus calculation step. The average pillar stress and the average loading system displacement exhibit similar trends, therefore they are both illustrated as the narrow width line. The bold line represents average pillar strain. The step interval, dT, denotes the time of failure and is defined as beginning at the onset of pillar softening through the occurrence of residual stress. The local stiffness is calculated using ∆F and ∆D during the interval dT. Then, by considering the P 71
Colorado School of Mines
for observing damage is necessary. Also, observing damping work on a localized basis could indicate whether the damage is due to a stable or unstable failure and give a measure of the intensity of failure. A grid based measurement technique is implemented to observe the behavior of the contact softening and damping work on a local level. To track contact softening and damping work spatially in the model, a fictitious grid that is comprised of square pixels is superimposed onto the PFC2D assembly. The square pixels are 0.1 m on a side and grid resolution is kept constant in each model as model size changes. Each particle and contact is permanently assigned to a pixel at the beginning of the simulation, thereby ignoring effects of pixel-to-pixel movement. Irregular values at the model’s boundaries, due to empty space in the pixels, have been found to be irrelevant to model behavior and can also been ignored. Listing C.15 shows the FISH code used to execute the grid based measurement technique. The functions included in this algorithm compute both grid based values and totals for damping work and contact softening. Figure 4.13 is a flow chart depicting the process of calculation. The algorithm can be described as having two parts, the initialization and the calculation cycle. The initialization defines the necessary functions, grid, and memory arrays for data processing and histories. The grid based calculation is executed at the beginning of every PFC2D calculation step. During the grid based calculation, the function loops through each particle and then each contact in the DEM in order to calculate values of desired parameters. Then the data array is updated and the data histories are recorded. Although values of the indicators are accumulating every step, histories are only recorded once every 5000 calculation steps in order to reduce memory usage. 4.4 SPCS Test Results The results of SPCS test results are presented in the following section using line plots to show the stress versus strain behavior of each pillar and the loading system displacement during each test. Stability indicator results are presented and then analyzed in the context of whetherfailureofthepillarisstableorunstable. Thenthegridbasedindicatormeasurements 73
Colorado School of Mines
are presented and discussed. 4.4.1 Pillar Stress-Strain Behavior and Loading System Displacement The stress-strain results from the nine pillar tests are organized into three plots. Each plot contains three tests, showing stress-strain behavior of one width to height ratio tested with 5, 20 and 35 GPa elastic modulus loading systems. Figure 4.14 shows stress-strain behavior for the width to height ratio one pillar, Figure 4.15 shows stress-strain behavior for the width to height ratio two pillar, and Figure 4.16 shows stress-strain behavior for the width to height ratio three pillar. Figure 4.14: Stress-strain curves for width to height ratio one pillar tests Eachcurveshowshowpillarstressincreasesduringtheloadingphaseofthetestsandthen as the pillar fails, stress is dissipated. Pillar strength is dependent upon the pillar size and the loading system stiffness. As pillar size increases the strength of the pillar increases, and as loading system stiffness decreases the pillar strength decreases. In the post-peak region, the post-peak modulus is dependent upon both the pillar size and loading system stiffness. As pillar size increases, the post-peak modulus decreases and as loading system stiffness increases the post-peak modulus increases. A significant change in post-peak behavior is 75
Colorado School of Mines
apparent in each of the 5 GPa tests where the change in post-peak modulus is more gradual from 35 GPa to 20 GPa loading system elastic modulus. The elastic displacement of the loading system in each of the nine tests are shown in plots organized in the same way as the stress-strain plots. Figure 4.17 shows average loading system displacement for the width to height ratio one pillar tests, Figure 4.18 shows average loading system displacement for the width to height ratio two pillar tests, and Figure 4.19 shows average loading system displacement for the width to height ratio three pillar tests. Loading system displacement increases during the loading phase of the tests and then de- creases as the pillar fails. Displacement at the point of failure is higher when elastic modulus of the loading system is low and increases as the pillar size increases. In the post peak region, the 5 GPa tests show a fast decrease in loading system displacement, while tests with 20 and 35 GPa loading system exhibit a more gradual decrease in loading system displacement. Figure 4.17: Loading system displacements for width to height ratio one pillar tests Using data from the stress-strain and displacement plots, stability of the pillar failure can be assessed. Similar to the EPC tests, sudden rebound of the loading system indicates unstable failure. A sudden rebound of the loading system can be detected by comparing the measured post-peak stiffness of the pillar to the loading system stiffness during failure. If 77
Colorado School of Mines
these two values are similar, unstable failure is assumed to have occurred. Figure 4.20 shows measurements of loading system stiffness and pillar post-peak stiffness during the failure interval for each test. The data are color coded according to loading system elastic modulus. Calculated loading system stiffness is represented by exes and pillar post-peak stiffness is represented by triangles. Generally, as loading system elastic modulus increases the pillar post-peak stiffness and loading system stiffness increases. Stability of the failure is determined by comparing the loading stiffness and post-peak stiffness for each test. The 20 GPa and 35 GPa tests show consistent difference between pillar behavior and loading system stiffness measurements indicating stable pillar failure for all six tests. The 5 GPa tests show coincident values, indicating unstable failure for all three pillar sizes. Based on these results, further analysis of indicators will assume stable failure for the 20 and 35 GPa tests and unstable failure for the 5 GPa tests. Figure 4.20: Pillar post-peak stiffness and loading system stiffness measurements 4.4.2 SPCS Test Indicator Results Results are presented here for each of the nine indicators during the nine SPCS tests. Figure 4.21 shows cumulative and instantaneous mean unbalanced force versus calculation 79
Colorado School of Mines
Figure 4.22 shows damping work for the pillar strength tests. When unstable failure occurs the damping work is markedly increased, while for both the stable failures of each width to height ratio, the damping work is similar. Despite the normalization with respect to both stress drop and pillar size, the damping work is higher for larger pillars. As with the EPC tests, a larger amount of damping work suggests that the failures are more unstable or in other words, more violent. Figure 4.22: Damping work in pillar strength tests Figure 4.23 shows the maximum instantaneous kinetic energy model during failure. Fig- ure 4.24 and Figure 4.25 show the maximum instantaneous mean and maximum unbalanced forces in the pillar strength tests, respectively. For each of these indicators the unstable failures generally have higher values. Although the difference between stable and unstable cases is not as pronounced as with the damping work. The maximum instantaneous max- imum unbalanced force for the 20 GPa width to height ratio two tests is an outlier in this trend. The maximum mean unbalanced force decreases for larger pillars. This likely is the result of averaging over a larger number of particles. Each of these indicators only contains model information for one calculation step, and therefore should be used with caution and in conjunction with other indicators. 81
Colorado School of Mines
Figure 4.25: Maximum instantaneous maximum unbalanced force in SPCS tests Cumulative values for kinetic energy, mean unbalanced force and maximum unbalanced force are presented in Figure 4.26 through Figure 4.28. Cumulative values may express the failurestabilityofthemodelbetterbecauseinformationiscontainedfromtheentireduration of failure. The cumulative kinetic energy describes the total amount of energy translated into motion that was initially stored in the specimen and loading system as strain energy. Figure 4.26 shows that the kinetic energy increases drastically for unstable failures while val- ues for stable failures are grouped at a noticeably lower magnitude. The cumulative mean unbalanced force in Figure 4.27 shows a similar behavior only the stable values are grouped more closely. However, as model size increases the number of elements over which the un- balanced force is averaged increases. Since unbalanced force is higher in the areas of damage many elements have low unbalanced force, therefore the mean unbalanced force decreases as model size increases. The values for stable failures also decrease slightly as the model size increases. The cumulative maximum unbalanced force in Figure 4.28 exhibits similar behav- ior, but values for stable cases increase for larger pillars. This trend suggests that for larger assemblies the maximum unbalanced force may not be able to clearly distinguish stable and unstable failure, but additional testing of larger pillars would need to be performed to verify 83
Colorado School of Mines
this. Figure 4.26: Cumulative kinetic energy in SPCS tests Contact softening is shown in Figure 4.29. The cumulative amount of contact softening describes the plastic displacement of the model on the contact level. Contact softening distinguishesbetweenstableandunstablefailuresinthatstablefailuresforsimilargeometries exhibit similar amounts of contact softening while the unstable failures display a larger amount of contact softening. Although, as pillar size increases for unstable failures the amount of contact softening ceases to increase. Once again, additional tests on larger pillars may reveal additional features to the trend. The number of broken contacts normalized with respect to stress drop and pillar size is shown in Figure 4.30. The number of broken contacts is the number of contacts that have reached the softening limit. Broken contacts are typically thought of as cracks in DEM models, but due to the softening component this definition is debatable. Regardless of the definition of crack location, the location of a broken contact identifies a location of significant damage in the model. The normalized number of broken contacts is consistently higher for unstable failures and there exists a large gap between closely grouped stable failures and the unstable failures. As with mean unbalanced force and softening indicators, the number of 84
Colorado School of Mines
Figure 4.29: Contact softening in pillar strength tests broken contacts levels off as pillar size increases for unstable failures. 4.4.3 Grid Based Instability Indicator Results The damping work and contact softening were measured using the grid based measure- menttechniqueshowninFigure4.13. Figure4.31throughFigure4.34showcontactsoftening and damping work for each of the pillar strength tests. Each image is produced from the data at the last step of the failure interval used for the previous indicator analysis. A shaded bar is provided with each image to indicate the range of values present. The image value range is determined by setting the maximum value to the maximum value detected in the grid. This way a comparison between tests can be made using both the local maximum magnitude of the indicator and the pattern of the indicator. Figure 4.31, Figure 4.32, and Figure 4.33 each show that the maximum value of contact softening increases as the elastic modulus of the loading system decreases. Also, maximum contact softening increases as pillar size increases. A maximum of 1.4 meters in the unstable width to height ratio three pillar is measured. While cumulative values of contact softening show a distinguishing difference in magnitude between stable and unstable failures, the local 86
Colorado School of Mines
Figure 4.30: Broken contacts in pillar strength tests measurements do not express the same trend. Rather, local maximum contact softening is higher in unstable failures than stable failures but not by a large enough degree to use with confidence to distinguish stable and unstable failures. The contact softening in each tests shows that damage in the model occurs in planes that resemble shear planes. Both stable failures and unstable failures damage similarly in so far as planes of damage form in similar locations for similarly sized pillars. Although, concentrations of contact softening are noticeable in the damaged areas of the unstable failures. This can better be seen by comparing the 35 GPa tests, which are the most stable, to the 5 GPa tests, which is an unstable failure. Localization of failure along a plane would contribute to higher values for individual grid pixels and could explain the trend of higher local contact softening for unstable failures. The maximum local damping work follows the trend previously demonstrated by cumu- lative damping work in Figure 4.22. The maximum local damping work for the unstable failures is noticeably higher compared to the stable failures. As the loading system elastic modulus increases and the failures become more stable, the damping work decreases further. Also, for stable failures the damping work is more distributed throughout the model. The 87
Colorado School of Mines
Figure 4.35: Damping work during unstable failure of width to height ratio three pillar with decreased value range (kJ) values for stable failures and then increases with the degree of instability. The ideal indicator can perform both functions, to identify instability and to quantify the degree of instability. According to the results in this chapter, cumulative values more completely describe the failure and are more reliable than the maximum instantaneous values because of less vari- ability in trends. The cumulative damping work and cumulative kinetic energy are superior indicators because they identify unstable failure when compared to stable failures and give a qualitative measure of the intensity of failure. Otherindicatorsareaffectedbythesizeofthemodelsuchascumulativemeanunbalanced force and contact softening. These indicators might be useful in conjunction with damping work and kinetic energy to confirm instabilities. The performance of contact softening as a viable stability indicator in the pillar tests was a surprising result given the highly variable nature of contact softening in the EPC tests. Continued caution should be exercised when using this indicator for anything but a damage indicator. Grid based measurements showed that damping work and contact softening can not just provide a picture of damage in the model but also support identification of instability by cumulative indicators. The magnitude of local damping work, as tracked using the grid based technique, can indicate instability when compared to stable failure. The localization of damage, as shown with contact softening and damping work, further supports the deter- mination of failure stability. During unstable failure, damage appears to localize along a 92
Colorado School of Mines
CHAPTER 5 UNSTABLE FAILURE IN AN IN SITU PILLAR MODEL In underground coal mining, as areas are mined out, failure occurs on the edges of the pillars, or ribs. As the excavated area increases and the pillar size decreases, the failed area proceeds into the pillar. Stable or unstable failure of the rib material can occur while the pillar as a whole retains load bearing capacity. In order for unstable failures to occur on the rib of the pillar two conditions must be met. First, the material must fail, and second, the loading system stiffness must be less than the material’s post-peak stiffness. In this chapter, an in situ pillar (ISP) model is used to investigate failure of pillar ribs. The model is first described in detail. Then the model is verified by comparing analytical solutions to an elastic FLAC2D model and then to the ISP model. The coal material, modeled using PFC2D, is mined in using a realistic mining sequence and failure of the pillar near the rib is observed. Stability indicators are used to distinguish between stable and unstable failure, namely, damping work, kinetic energy and mean unbalanced force. Spatial measurements of damping work and contact softening are then used to support the stability indicator results. 5.1 Model Description A two-dimensional mechanically coupled DEM/FDM model similar to that of the pre- vious chapters is utilized. The geometry and load application scheme is modified in order to simulate an in situ pillar panel under development and then further mining. The failing material is modeled using the displacement softening model (DSM) and the surrounding mine and pillar core is modeled in FLAC2D. In situ stresses are installed to simulate a deep coal mining scenario and the DSM material is slowly removed in order to simulate the min- ing process. As the DSM material is mined, installed stresses redistribute and cause failure 94
Colorado School of Mines
to occur. The failure is characterized by tracking stress measurements and damping work, kinetic energy, and mean unbalanced force. 5.1.1 ISP Geometry, Boundary Conditions, and Material Properties Figure 5.1 shows the geometry and boundary conditions for the in situ pillar model. The grey and blue areas indicate FLAC2D zones of different grid types and the yellow region is the PFC2D assembly. The blue area, labeled FLAC2D Inner Grid, is a fine grid comprised of square zones one-fifteenth of a meter on each side. This fine grid is intended to achieve a high resolution of stress measurement near the PFC2D assembly and to comply with the recommended coupling boundary ratio of four to five PFC2D elements to one FLAC2D zone. The grey area labeled FLAC2D Outer Grid is graded outward to increase computational efficiency and adhere to memory constraints. Directly above and below the inner grid, the grid is graded only vertically, retaining constant zone width. To the right of the inner grid, the grid is graded only horizontally, retaining constant zone height. In the remaining areas the grid is graded in both directions. Listing F.17 shows the FLAC2D grid generation file. The dimensions of the model are symmetric about the vertical center of the PFC2D part. The top, left and right boundaries are fixed with a roller boundary while the bottom boundaryispinned. Theredarrowindicatesthedirectionofexcavationandtheblackdashed lines show the locations of two FLAC2D interfaces. The placement of the dashed lines is exaggerated to indicate that the interfaces are located one zone width within the FLAC2D grid. The PFC2D part is composed of twenty square pbricks stacked two high and ten wide. ThematerialpropertiesandelementsizeofthePFC2DmaterialarethesameastheDSM used in the previous chapters and shown in Table 3.2. The only difference is the number of pbricks in the horizontal direction. The FLAC2D material is divided in two sections. They are the surrounding rock and coal regions, both are elastic. The FLAC2D coal spans the same height of the PFC2D coal and extends from the right edge of the PFC2D part to the right boundary of the model. The zones in the FLAC2D coal region are assigned a shear modulus, bulk modulus and Poisson’s ratio corresponding to the DSM elastic modulus and 95
Colorado School of Mines
Poisson’s ratio shown in Table 3.4. The surrounding rock material is assigned bulk and shear moduli corresponding to an elastic modulus of 35 GPa and a Poisson’s ratio of 0.25. The constitutive model used for the interfaces is a Mohr-Coulomb elastic perfectly plastic model. The assigned properties are given in Table 5.1. A zero degree friction angle and low cohesion are assigned to simulate a slick interface with low strength. Table 5.1: FLAC2D interface properties 5.2 ISP Model Execution Theuseofthediscreteelementmethodformodelingrockrequiresinitialstepstogenerate the material and install in situ stresses. For coupled applications, where a continuum model performsthefunctionofasurroundingrock, additionalstepsareneeded toinsertthePFC2D material into its assigned region and bring the coupled system into equilibrium. During any part of the initialization procedure, if unbalanced forces in the system are high, contact bonds can be broken. Any damage inflicted upon the system at this stage is unrealistic and should be minimized. Careful initialization of the model ensures that the expected DEM material properties are retained. Unlike the slender pillar model of the previous chapter, the in situ pillar model utilizes in situ stresses for load application. During initialization, the free mining face creates an opportunity for high unbalanced forces to destabilize the DEM system. In order to prevent unnecessary damage to the system, the stresses are installed in the PFC2D part separately, then the mechanical coupling is initialized and the coupled model is equilibrated with in- stalled stresses. Then the left boundary of the PFC2D part is slowly released to create a free mining face with minimal initial damage. Following model initialization, the PFC2D mate- rial is deleted in thin slices to simulate the mining process. The following sections provide a 97
Colorado School of Mines
Figure 5.2: PFC2D stress installation screen shot, 16 MPa vertical stress target execute the equilibration process in FLAC2D and Listing F.20 and Listing F.21 are used for PFC2D. Additional custom files necessary for the in situ pillar (ISP) model runs in this chapter are shown in Listing F.22 through Listing F.26 The final step in initializing the coupled model is to remove the pressure boundary on the PFC2D part and bring the model to equilibrium. Listing F.27 shows the sequence of commands to initiate this process in FLAC2D and Listing F.28 shows the commands for PFC2D. The functions needed to excavate material and bring the model to equilibrium au- tomaticallyareshowninListingF.25andListingF.26forFLAC2DandPFC2Drespectively. The pressure must be removed gradually so that minimal damage is imposed due to sud- den deconfinement. A pressure reducing function performs this task. Calling the function bdry loop in both FLAC2D and PFC2D starts a pressure reduction process in which the pressure is reduced incrementally and brought to equilibrium after each reduction step until the pressure is reduced to zero. Then, finite strength is restored to the PFC2D element contacts. 99
Colorado School of Mines
5.2.3 Excavation Excavation then begins by calling the function slice loop in both FLAC2D and PFC2D. Excavation proceeds by deleting elements to the left of an advancing mining face position. Once a selection of elements is deleted, the model is cycled until equilibrium is achieved. Then the face position moves forward one mining increment. The mining increment used in this study is equal to the average element diameter. The single layer of FLAC2D zones adjacent to the PFC2D model and under the interface are deleted as the mining face passes by. The model is saved at 0.5 m mining face advance increments until the mining distance limit is reached. In this study, eight of the ten meters of PFC2D material is mined. 5.3 Model Verification To verify the performance of the coupled in situ pillar model, first, analytical solutions for closure of a tabular excavation and associated abutment vertical stress are compared to a FLAC2D model. Then closure and abutment stress is compared between elastic versions of the coupled in situ pillar model and FLAC2D. Closure of the excavation and abutment vertical stresses are compared for a tabular excavation span of 6 m at 630 m depth with a lateral earth pressure coefficient of K = 0.3 and a surrounding rock elastic modulus of 35 E GPa. 5.3.1 FLAC2D Measurements The stresses and displacements are captured in each model using FLAC2D zones and grid points respectively. Figure 5.3 shows the FLAC2D grid for the ISP model at the final excavation stage. This figure includes an inset of the FLAC2D grid adjacent to the PFC2D part of the model. The inset shows the locations of zones used for stress measurement and grid points used for displacement measurement in the roof. In order to plot vertical stress as a function of position, stress is averaged among the two zones adjacent a grid point at a given position x. While only mine roof measurement locations are shown, a mirrored scheme is used for the mine floor. The vertical stress presented for comparison is the average of the 100
Colorado School of Mines
roof and the floor stress for each x position. Closure is the sum of roof and floor grid point displacement for a given position, x, where displacements toward the excavation are positive. Figure 5.3: ISP model FLAC2D grid measurement locations 5.3.2 Closure and Vertical Stress in FLAC2D Equation 5.1 is the closure of a tabular excavation at a distance x from the center of e the width of the excavation where the span is 2l , the in situ vertical stress is σ , and the e v surrounding rock has shear modulus G, and Poisson’s ratio ν. Figure 5.4 shows a rectangular excavationinwhichthedimensionsarelabeledandtheclosureisdemonstrated. Thissolution assumes plane strain conditions, that the closure at the edge of the excavation is zero, and the extent of the rock in the vertical and horizontal directions is infinite. Abutment vertical stress (σ ) into the abutment a distance x is given by Equation 5.2 Ozbay [62]. z a 2σ (1 ν) s = v − l2 x2 (5.1) G e − e ￿ 101
Colorado School of Mines
σ x v a σ = (5.2) z x2 l2 a − e The tabular excavation is modeled in FL￿AC2D using a vertical line of symmetry about the center of the excavation. Roller boundary conditions are applied to the grid edges and the rib is fixed. A vertical stress of 16 MPa is used and lateral earth pressure coefficient of K = 0.3 is used to generate the corresponding horizontal stress. E (a) Excavation dimensions (b) Excavation closure Figure 5.4: Static deformation of a tabular excavation in elastic medium Figure 5.5 shows the comparison of closure for the analytical and FLAC2D solutions. The difference between the FLAC2D and analytical closure at the center of a 6 meter wide excavation is approximately 0.01% of the excavation height (2m). Figure 5.6 and Figure 5.7 show the vertical stress comparison for the FLAC2D and analytical solutions near the rib and for the full width of the coal respectively. Figure 5.6 shows that the FLAC2D model experiences a sharp increase in stress at the rib similar to the analytical solution, but not as high in magnitude. Figure 5.7 includes two horizontal lines indicating the virgin stress magnitude at 16 MPa. As the distance from the rib increases, the vertical stress decreases to meet the virgin stress value. 102
Colorado School of Mines
Figure 5.7: Vertical stress comparison between FLAC2D and analytical solution, full model span Figure 5.8 shows closure for the FLAC2D and coupled ISP models for a 6 m by 2 m tabular excavation. In these models, the rib displacement is not fixed, so the value of closure at the rib is non zero. The closure values for each model are in close agreement indicating that the DSM material satisfactorily simulates elastic displacement due to a given stress field. Figure 5.9 and Figure 5.10 show vertical stress in the FLAC2D and coupled ISP models near the rib and for the full span of the coal respectively. Near the rib, vertical stress values are similar for each model for respective depths. Each result exhibits an stress increase at the coal rib. The coupled ISP model shows an unexpected decrease in stress at the rib, where the FLAC2D model does not show this decrease. This inward movement of the stress concentration in the coupled model is due to larger horizontal displacement of the rib in the coupled ISP model resulting in slight rib destressing and therefore, inward movement of the stress concentration. Figure 5.10 shows the full span of the coal material. The plot terminates just after 16 meters because stress is the average of two adjacent zones. Because of zone gradation, the 104
Colorado School of Mines
zone adjacent to the the grid point near x equal 16 meters is nearly 4 meter in width. The dotted lines in Figure 5.10 show the installed virgin stress state. As distance from the rib increases, vertical stress approaches these values for each depth. The irregular stress pattern, as measured in the FLAC2D zones, is due to roughness of the PFC2D assembly. This stress pattern is periodic as a result of using identical pbricks to create the PFC2D assembly. Figure 5.8: Closure comparison between FLAC2D and coupled ISP model solutions 5.3.3 Effect of the Coupling Boundary on Vertical Stress Figure 5.10 shows a stress increase in the coupled ISP model near the right coupling boundary. The stress increase is due to the mismatch of material behavior at the coupling boundary and can influence results and interpretation of results if the area of interest is near this boundary. In order to examine this effect in greater detail, elastic simulations for various entry widths are performed. The model is initialized, then the PFC2D material to the left of the appropriate excavation boundary is deleted, then the model is cycled until equilibrium was achieved. Roof stress is then analyzed, as opposed to average roof and floor stress, so detail in stress changes could be seen. 105
Colorado School of Mines
Figure 5.11 shows vertical stress in the roof for various entry widths. The right coupling boundary is located at 10 meters. Comparing the rib stress for different entry widths shows an incremental increase at all locations. At a entry width of 16 meters, the mismatch in material behavior results in a lower than expected rib stress and higher than expected abut- ment stress . Therefore, if the area of interest is within two to three meters of the coupling boundary, analysis of model output should consider the effect of the coupling boundary. Figure 5.11: Vertical roof stress for various entry widths 5.4 Results Results are presented here for the ISP model with inelastic DSM material. As the exca- vation proceeds material fails at the rib due to the redistribution of stress. Stress profiles reveal the extent of failure by denoting the location of maximum stress for a given entry width. Damping work, kinetic energy, and mean unbalanced force are utilized to detect occurrences of unstable failure. Then, grid based, spatial measurements of contact softening and damping work are used to support identifier and stress results. 107
Colorado School of Mines
5.4.1 Zone Stress Measurements Roof stress profiles identical to Figure 5.11 are shown in Figure 5.12. As the excavation widens, high stresses on the edge of the material cause failure and redistribution of vertical stress inward, towards the unfailed portions of the pillar. As the entry widens, more stress must be carried by the pillar so at the point of maximum stress the magnitude increases for successive excavation steps. Figure 5.12: Deep depth roof stress profiles for various entry widths Thedegreeoffailureatspecificinstancesduringtheexcavationprocesscanbedetermined from the stress profile in Figure 5.12. The degree of failure is indicated by comparing the difference in vertical stress to the location of maximum stress. A low value of residual stress on the pillar rib compared to the maximum stress indicates extensive failure. Whereas if the pillar rib exhibited a greater amount residual stress, it would indicate that the material is still capable of bearing load and therefore has been damaged to a lesser degree. The degree of failure can be quantified by calculating the gradient of vertical stress from the pillar rib to the point of maximum stress. The vertical stress gradient is calculated by dividing the change in stress from the rib to the point of maximum stress by the distance between the rib and the x position of maximum stress. Figure 5.13 shows the rib stress 108
Colorado School of Mines
these plots, the vertical lines are the excavation steps, when a layer of elements is deleted. Since these plots are showing results from the same model, these mining steps are identical in each plot. The spaces in between the excavation steps is are the number of steps required to bring the simulation to equilibrium. There are 158 excavation steps in the simulation as the mining face advances from x equals 4 to 6 meters. The lines showing the time step of mining each have the same width. The appearance of thicker lines is an indication of several mining steps occurring closely to one another. ItcanbeseeninFigure5.14thattwoparticularminingstepsresultinsignificantincreases in the cumulative damping work performed. The first is at approximately 6.8 106 steps and × the other is near 8.4 106 steps. In Figure 5.15, significant increases in the instantaneous × kinetic energy are present during these steps and also in Figure 5.16, large amounts of unbalanceforcearepresent. Ascomparedtootherminingsteps, thenumberofstepsrequired for equilibration and the increase in identifier value signifies that unstable failures of the rib occurred at these locations. Despite the fact that the mining increments are very small in distance, it is rational to assume that there be some instability resulting from removal of elements. By comparing a typical stable mining step and the first unstable mining step, a significant difference in identifier behavior emerges. The stable step chosen is when the mining face is at x equals 4.709 meters, at timestep 6.678 106. The unstable mining step is when the mining face is × at x equals 4.772 meters, at timestep 6.784 106. Figure 5.17 shows the damping work in × the stable and unstable mining steps. The damping work accumulates during both mining steps, although the increase in damping work in the unstable case is an order of magnitude higher than the stable case. As seen in the SPCS tests, a large relative increase in damping work can indicate unstable failure. By plotting the incremental increase of damping work versus the mining extent in meters, the magnitude of damping work performed during each mining step can be clearly compared. Figure 5.18 shows the amount of damping work performed between excavation steps. Typ- 110
Colorado School of Mines
(a) Stable (b) Unstable Figure 5.17: Damping work during a stable and unstable mining steps ically, the amount of damping work performed during a mining step is below 5 kJ, but the damping work at the two unstable mining steps at 6.8 and 8.4 million calculation steps cor- respond to the high values near 4.8 meters and 5.65 meters. Figure 5.18 also reveals that there is a second tier of instability with intensities between 5 and 20 kJ. It should be noted that this value of energy is of a numerical nature and should not be considered as the amount of energy associated with a real unstable failure. Additional work must be undertaken to assess the accuracy of energy calculation using the ISP model. The kinetic energy and mean unbalanced force also exhibit a difference in magnitude between stable and unstable cases. However, aside from a change in magnitude, another revealing difference between the stable and unstable cases is shown in kinetic energy and mean unbalanced force. Figure 5.19 shows instantaneous kinetic energy in the unstable and stable mining steps and Figure 5.20 shows the instantaneous mean unbalanced force. For both cases, kinetic energy and mean unbalanced force increase immediately after the elements are deleted. In the stable step, the model equilibrates steadily as shown by steadily decreasing identifier value. In the unstable step, there is a secondary increase in indicator value unrelated to the initial deconfinement due to mining. This failure results from the mechanism for instability in which excess energy stored in the loading system is unable to 114
Colorado School of Mines
indicator trend. Region B indicates stable mining, Regions A indicates quasi stable mining, and Region C indicates and unstable failure. Figure 5.18 shows that the damping work magnitudes performed during the stable, quasi-stable, and unstable steps are in agreement with failure stability interpretation from Figure 5.14, Figure 5.15, and Figure 5.16. 5.4.4 Grid Based Measurements The previous section showed that stability indicators could detect the mechanism for unstable failure and revealed a distinct difference between an example of a stable mining step and an unstable failure resulting from a single mining step. By tracking the damping work and contact softening using the grid based measurement technique, the damage and instability in the model during each of these steps can be observed spatially. Figure 5.21 shows the damping work before and after the stable and unstable mining steps. Each of the images shows the rib region of the model, from x equals 4.8 to 7 meters. For comparability, each of the images is shaded according to the same scale. The scale used, from 0 to 160 J, is shown next to the image of the state before the stable mining step. Each of the images shows a parabolic shaped area of failure. There are subtle increases in the magnitude of damping work as mining progresses from the beginning of the stable mining step to the beginning of the unstable mining step. The image showing the state after the unstable mining step indicates that a large amount of instability occurs along the outer edge of the parabolic damage region. Figure 5.22 shows the damping work after the unstable mining step with the scale reset to resolve the larger pixel values. This image reveals a possible new failure surface further into the pillar than the most inner failure surface seen before the unstable mining step. Also, the magnitude of the damping work performed along the new failure surface far exceeds the values seen in the unstable mining step in Figure 5.21. Figure 5.23 shows the contact softening for the stable and unstable mining steps. The shading scale is common to each image and is given next to the image depicting the model state before the stable mining step. The maximum value is set to 0.5 m of contact softening to highlight the pattern of damage. As mining progresses the softening of the rib exceeds 117
Colorado School of Mines
the maximum value on the scale and therefore the outer most pixels are white. The dashed line adjacent each image denotes the edge of the measurement grid at x equals 4.8 meters. Figure 5.23: Contact softening in the rib during the ISP test Figure 5.23 shows that damage has accumulated near the rib in the form of planes of failure which extend from the rib corners inwards, toward the vertical centerline of the pillar. Before the stable mining step, one such failure plane is depicted along with a region of damaged material near central part of the rib. After the excavation of the stable mining step, damage accumulates in these two areas. In subsequent mining steps, the material at the rib softens but there is no significant accumulation of damage along the failure plane, as seen in the state of the model before the unstable mining step. After the unstable mining step, the contact softening shows the formation of a new failure plane. 119
Colorado School of Mines
CHAPTER 6 CONCLUSIONS AND FUTURE WORK To summarize the work in this thesis briefly, first, an appropriate DEM model for study- ing compressive failure stability was chosen. Then a series of model behaviors were defined to use as indicators of failure stability. These were evaluated during a series of pillar strength tests and the most appropriate indicators were identified. Then the failure of an excavation loaded under in situ, mining conditions was investigated using the indicators on a global and also localized basis. In general, the numerical models behaved acceptably for the purpose of studying unstable compressive failure in western U.S. coal and the methods used to dis- tinguish between stable and unstable failure were successful. The following is a concise list of conclusions on a chapter by chapter basis. Then a list of suggested future work is given followed by some additional research questions inspired by this work. Chapter Conclusions Ch 2. Background Information on Unstable Failure in Underground Coal Mining A need exists to ensure failure stability in deep underground western U.S. coal mines • due to the high probability and potential risks associated with unstable failure in western U.S. coal mines. A theoretical background for the mechanism of unstable compressive failure in brittle • rocks exists, but additional work is needed to include stable and unstable failure modes in mechanistic numerical studies. TheDEMcodePFCoffersfeaturessuchasemergentrocklikebehaviorsandanimplicit • time steping solution scheme that allows for multi-stage simulation of unstable failure. 121
Colorado School of Mines
To the best of the author’s knowledge no previous work successfully simulates unstable • compressive failure using a discrete element model. Ch 3. Evaluation of Two DEM Models for Simulating Unstable Failure in Compression The BPM and DSM described in Chapter 3 are two discrete element models that are • capable of simulating a western United States coal with post-peak softening behavior for the purpose of studying compressive failure stability. Currently, the BPM requires a ’black-box’ type of computer algorithm to determine • microparameters, because the combination of parameters necessary to define charac- teristic post-peak behavior is not known. The DSM requires an iterative calibration that can be conducted manually. The key • microparameter influencing the post-peak softening behavior is the contact plastic softening limit, U . pmax Triaxial tests results revealed lower than desired friction angle for the BPM, an ex- • pected result. The DSM triaxial tests showed a higher than desired friction angle. The DSM model showed more consistent behavior during the failure stability (EPC) • tests in that post peak behavior remained consistent for stable failures than did the BPM model. Furthermore, the transition from stable to unstable failure mode with various loading system stiffnesses was more defined with the DSM, while the BPM exhibited a fairly large quasi-stable region. The BPM exhibits a clear dependency of post-peak softening on loading rate. For • lower loading rates, the BPM post-peak stiffness increases in magnitude. The DSM exhibits an effect on post-peak behavior for different loadings rate, however • the general softening characteristic of the material is retained. 122
Colorado School of Mines
The DSM is a more appropriate DEM to use in studying failure stability than the • BPM based on consistency of behavior in stable and unstable failure mode and the independence of DSM post-peak softening by loading rate. Chapter 4. Indicators of Unstable Compressive Failure in DEM Coal Strength Tests Cumulative indicators better represent the failure of the model because they embody • information from the entire failure rather than from one calculation step as was the case with maximum instantaneous values. Although, trends of instantaneous values also indicate the behavior in the model. In both the EPC tests and the SPCS tests the damping work and kinetic energy • differentiated between stable and unstable failure and provided a qualitative indication of the magnitude of failure. Some indicators are affected by the size of the model as shown in SPCS tests. The • mean unbalanced force, for example appears to decrease as model size increases. So, this indicator should be used in conjunction with damping work and kinetic energy. The contact softening indicator does not clearly distinguish between stable and unsta- • ble failure when analyzed globally. However, this indicator could be used to provide information on location and extent of damage in the model. Grid based measurements for damping work and contact softening showed higher local • valuesforunstablefailuresandsimilarvaluesforstablefailuresanddepictedthefailure patterns in the models. Chapter 5. Indicators of Unstable Compressive Failure in DEM Coal Strength Tests The stress gradient after arbitrarily chosen mining steps suggested increased possibility • of unstable failures as the excavation is expanded to exceed four times the excavation height. 123
Colorado School of Mines
When instabilities occur, an increase in indicator value will occur that is independent • of the initial removal of elements. The mean unbalanced force, damping work, and kinetic energy indicate two significant unstable failures with increased magnitude of values after increasing the excavation width beyond four times the seam height.. Plots of damping work, kinetic energy, and mean unbalanced force versus step show • that fewer steps are required to equilibrate the model during stable mining. A larger number of steps are needed to equilibrate the model when unstable failure occurs. As revealed by the grid based measurements, single mining steps can result in the • initiation of significant unstable failures. Future Work TheDEMmodelsandstabilityindicatorsinthisthesisareapplicabletoinvestigatingspecific mechanisms of unstable failure and conditions that influence them. By changing existing model parameters the identifiers can be used to potentially study the effect mine conditions haveontheintensityandfrequencyofunstablefailures. Inthiscontext, additionalnumerical analysis should be conducted on the following topics: The effect of the coal/mine contact condition • TheFLAC2DpartofthemodelinChapter5containsaninterfacewithMohr-Coulomb strength properties with perfectly plasticity. This interface is intended to simulate the contact condition between the coal and a competent adjacent rock. While it is difficulttodeterminetheactualmaterialpropertiesofdiscontinuitiesinminestheeffect of different idealized discontinuity behaviors on unstable compressive failure can be evaluated. For example, using the Mohr-Coulomb with perfect plasticity interface, the strength of the discontinuity could be changed to simulate various levels of horizontal confinement on the coal due to contact conditions. By improving the constitutive law 124
Colorado School of Mines
of the effect of unstable slip along the interface could considered. Unstable failure researcher Gu demonstrated that a discontinuity with a softening post-peak behavior cansimulatestableandunstableslip, analogoustocompressivefailurestabilitycriteria. By modifying the discontinuity plasticity law in FLAC2D by means of user defined FISH function to include a softening post-peak behavior the effect of unstable slip on compressive unstable failure could be studied [28]. Depth of the mine • It is widely agreed upon that unstable failure is more probable as depth of the mining activity increases. By initializing a series of ISP models at different depths. The effect of depth on frequency and intensity of unstable failure can be evaluated. Various types of coal • Thepost-peakbehaviorofcoaliskeptconstantthroughoutthisstudy. UsingtheDSM, coal materials with different levels of brittleness can be calibrated and tested under similar conditions. Both EPC and ISP tests on these coals could serve to confirm Cook’s stiffness stability criteria concept on a theoretical basis. Mining rate • In this thesis, the criteria for model equilibrium is set to simulate the onset of static equilibrium. The mining rate in actual coal mines is known to effect mine stability [90]. Pillar design schemes • Various combinations of pillar sizes are used to offer support in gateroad entries in longwall coal mining. The ISP model provides an opportunity to study the effects of pillar design and pillar loads on stability in these entries. Additional Research Questions 125
Colorado School of Mines
Failure localization • An interesting result that arose from this study is that of the failure localization due to unstable failure. In chapters four and five the grid based measurements of damping workandcontactsofteningshowedamoredispersedtypeoffailureresemblingcrushing for stable failures and localized failure along planes, resembling shear bands, for un- stable failures. This behavior is more prevalent in BPM models as compared to DSM models [49]. This result suggests that a different failure mechanism is in effect when failure is unstable. DEM models hold promise in studying the failure mechanism due to their micromechanical nature. Effects of model properties on failure pattern such as particle assembly, particle size, and contact and bonding models should be system- atically tested to determine the nature of failure localization in DEM. If not already sufficiently performed, laboratory testing could reveal if there is a physical analogue. Alternatives to the DSM for studying unstable failure • The BPM has been used widely to simulate the failure of rock because it exhibits physical properties, such as increased strength with confinement and the Poisson ef- fect, of rock with out the explicit assignment of such properties [68][5]. The difficulty in calibrating a velocity dependent post-peak behavior in the BPM, in part, lead to the selection of the DSM for the work in this thesis. The DSM exhibits undesirable properties, such as an unrealistically high friction angle and high Poisson’s ratio. How- ever, the ease at which post-peak softening is calibrated is key to simulating unstable behavior in in situ loading conditions. Improvements should be made on these exist- ing models to cope with their respective drawbacks by closely examining the effect of contact laws on post-peak behavior. A simple comparison between BPM and DSM suggests that some form of softening behavior must be in action on the contact level. Alternatives to the these contact models should be thoroughly reviewed and possible usage of other numerical methods should be considered. 126
Colorado School of Mines
cp buf(2) = mv H 2.0 md ravg ; rectangle height − ∗ cp bufn = 2 cp write ; cp bufn = 3 cp read cpp nseg = cp buf(1) cpp yoff = cp buf(2) cpp rad = cp buf(3) cpp nseg0 = cpp nseg ; ; cpp excavate slp init EPC slp getwhat = 0 ; coords slp getlist cbi init ; command SET dt dscale ; Assumes that FLAC is running in static mode ( − default). ; By making PFC2D also run in static mode, we insure − ; that the displacements during one step in each code ; will be the same. Static mode means timestep of − unity , ; so velocities have units of [meters/step ]. SET fishcall 0 cpp getvel SET fishcall 3 cpp putfor SET fishcall #FC BALL DEL cpp delball SET fishcall #FCNEWQUIT cpp delballremove end command ; cpp coupled view oo=out(’ Coupling scheme successfully initialized . ’) ∗∗∗ end ; −−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−− def cpp cyc calm ; ; Controls synchronous cycling between PFC2D and FLAC. PFC2D is −−−−− the ; controlling process such that when FLAC is in slave mode, calls − to ; [cpp cyc] from PFC2D will force both codes to take one step . ; The coupling scheme assumes that cycling occurs only by calling ; [cpp cyc] from PFC2D. DO NOT ISSUE CYCLE COMMANDS DIRECTLY AND ; DO NOT TYPE ESCAPE WHILE CYCLING IN EITHER CODE! ; 162
Colorado School of Mines
APPENDIX B - INVESTIGATING THE EFFECT OF INTERNAL STRAIN ENERGY ON POST-PEAK BEHAVIOR OF THE BPM AND DSM In response to the results of the LRC tests, an additional study was conducted in order to test a possible explanation for the velocity dependent post-peak behavior. The interaction between internal strain energy and strength increase due to confinement could possibly lead to the observation of different levels of characteristic softening under different loading rates. Stored strain energy accrued during elastic deformation might lead to instability in the post- peak region of the DEM material if the post-peak characteristic is such that the strain energy cannot be absorbed during failure. At the moment of failure, this loading condition would lead to sudden failure with no post-peak softening. Although, if the loading rate is high, the amount of work added to the system each step may exceed the amount of energy released. If so, a higher material strength in the post-peak region could be observed due to increased confinement. In other words, a post-peak softening characteristic would be observed that is not a characteristic property of the material, but dependent upon the loading condition. Three specimens are subjected to a modified UCS test to investigate the effect of elastic strain energy on DEM stability in the post-peak region. Here, a UCS test is conducted as in Chapter 3.4, but here, the specimen is loaded just beyond failure and loading is halted when the vertical stress on the specimen is ninety five percent of the strength. At this point the model is cycled in order to determine stability of the specimen. If no change in stress occurs then the specimen is stable, if the stress continues to drop then the specimen is unstable. The three specimens tested are the two DEM specimens from Chapter 3 and a recalibrated DSM to with a steeper post-peak softening curve. First, the results of the tests on the DEMs from Chapter 3 are discussed then the results for the two DSM tests are discussed. FigureB.1showsthestressversusstepcurvesforthemodifiedUCStestonthetwoDEMs from Chapter 3. The black dot on each line designates the point where loading is halted. 169
Colorado School of Mines
The DSM model shows a slight decrease in stress indicating a brief period of instability. The BPM model shows shows a greater decrease in stress. The small amount of instability in the DSM model reflects the material’s ability to regain stability after some energy release, whereas the instability of the BPM specimen shows that stability is not regained. Figure B.1: Stability test stress-strain curves for Chapter 3 DEM models The stability concept tested in Chapter 3 using the EPC tests states that if the energy stored within the loading system at the point of failure cannot be absorbed by the specimen then the failure will be unstable. During unstable failure, the characteristic post-peak be- havior of the material is hidden, and the strain measurement taken at the platen-specimen boundary reflects the rebound of the platens. In the case of the BPM tested here, when loading is halted, work is no longer done on the system by the loading mechanism. If the failure is unstable, another source of energy must be acting on the system. Elastic strain energy in the BPM specimen could cause unstable failure if the specimen is not able to dissipate all of the stored energy during the failure process. DEM material is not perfectly linear in the elastic region or in the post-post peak. Al- though a simplification of the behavior is useful in illustrating a possible mechanism for 170
Colorado School of Mines
failure stability of DEM specimens under rigid loading conditions. Figure B.2 shows a schematic of a stress strain curve where the elastic region and post-peak region are made linear for simplification. The hatched area U is stored elastic strain energy up to the point E of failure. The unhatched region, U , can be thought of as the capacity of energy storage in C the specimen during failure. In order for failure to be stable, U < U . If U > U , then E C E C | | | | the available energy is greater than the material’s capacity to store the energy and failure will be unstable. Figure B.2: Strain energy regions for linear elastic material with linear softening Figure B.3 shows two stress-strain curves, one for the DSM from Chapter 3 and one for the recalibrated DSM, labeled DSMr. The DSMr curve is calibrated to have similar elastic modulus and strength, but with a steeper post-peak curve. The steeper post-peak curve reflects a post-peak behavior in which less energy can be absorbed during failure than by the DSM specimen. The curves in Figure B.3 are both non-linear in both elastic and post-peak regions. So, making an exact determination of available strain energy versus energy storage capacity would require a determination of elastic unloading behavior at any given point in the post-peak curve. Although, an estimation on the likelihood of failure stability can be 171
Colorado School of Mines
made based on the energy criteria above. DEM materials with shallow post-peak softening compared to the elastic modulus will most likely be stable under rigid loading, materials with steep post-peak softening will likely be unstable and material with post-peak softening approximately equal in magnitude to the elastic modulus are questionable and should be closely examined. By approximating the post-peak behavior versus the pre-peak behavior in Figure B.3, it is likely that DSMr will be unstable under rigid loading conditions. Figure B.3: DSM characteristic stress-strain curves Figure B.4 shows two stress-strain curves, one for the DSM and one for the DSMr. The black dots show the point at which loading is halted and the model is cycled to test stability. Figure B.4 shows that the DSM has a partial instability and the DSMr material fails com- pletely after loading is halted and the model is cycled. Total failure of the material indicates that the DSMr is unstable under rigid loading. The instability of the DSMr specimen sup- ports the claim that internal strain energy magnitude in reference to the post-peak softening characteristic plays a significant role in modeling rock behavior with DEM. 172
Colorado School of Mines
endif ; Total SE pb = SE pbn + SE pbb + SE pbs else SE pb = 0.0 endif ; Total Strain Energy at contact cp −−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−− cp se = SE cp + SE pb end def pfc sof if c model(cp) = ’udm softening ’ then if c prop(cp, ’ sof broken ’) = 0 cp sof = c prop(cp, ’ sof softened ’) ; ratio of amt. yielded to yield limit , 0 = no yield sof nbcnt = sof nbcnt + 1 dumA = gridcell cont nbcount(cell indexY , cell indexX) dumA = dumA + 1 gridcell cont nbcount(cell indexY , cell indexX) = dumA endif else cp sof = 0.0 endif sof tot = sof tot + cp sof end def pfc wd bp wd = 0.0 ;This is the amount of work done in one step on one ball ;and it should be zeroed before pfc wd calculates the work ; in order to get a cummulative value for each grid cell . ; ; Incremental values can be computed by zeroing the grid array ; after each step , but this is not wise because the histories for ;each grid must be recorded every step for the values to be useful ; quantitatively . Otherwise , they can be a qualitative indicator . ; 181
Colorado School of Mines
history 501 syyf1 1 history 502 syyf3 1 ; ; history # syyrn 1 ; Lines deleted for brevity , fill in ; ; all history syyfn 1 commands ; history 596 syyf191 1 history 597 syyf193 1 history 598 syyf194 1 history 599 syyf195 1 history 600 syyf196 1 history 601 syyf197 1 history 602 syyf198 1 history 603 syyf199 1 history 604 syyf200 1 history 605 syyf201 1 ; history 701 rdL 1 history 702 rdL 3 ; ; history # rdL n ; Lines deleted for brevity , fill in ; ; all history rdL n commands ; history 796 rdL 191 history 797 rdL 193 history 798 rdL 194 history 799 rdL 195 history 800 rdL 196 history 801 rdL 197 history 802 rdL 198 history 803 rdL 199 history 804 rdL 200 history 805 rdL 201 ; history 901 fdL 1 history 902 fdL 3 ; ; history # fdL n ; Lines deleted for brevity , fill in ; ; all history fdL n commands ; history 996 fdL 191 history 997 fdL 193 history 998 fdL 194 history 999 fdL 195 history 1000 fdL 196 history 1001 fdL 197 history 1002 fdL 198 261
Colorado School of Mines
ABSTRACT Knowledge of geomechanical properties is beneficial if not essential for drilling and completion operations in the oil and gas industry. The Unconfined Compressive Strength (UCS) is the max- imum compressive force applied to cylindrical rock samples without breaking under unconfined conditions. Unconfined Compressive Strength (UCS) is one of the key criteria to ensure safe, effi- cient, and successful drilling operations, and estimation of UCS is vital to avoid wellbore stability problems that are inversely correlated with the pace of drilling operations. Furthermore, UCS is an essential input to ensure the success of completion operations such as acidizing and fracturing. Different methods are available to estimate UCS. The common practice to estimate UCS is to conduct experiments with a laboratory testing setup. These laboratory experiments are considered the most accurate way to measure UCS, but they are destructive, time-consuming, and expensive. Alternatively, empirical equations are derived to estimate UCS from well-logging tool readings. These empirical equations are generally derived from physical properties such as interval transit time, porosity, and Young’s modulus. However, most of these equations are not generic, and their applicability for other formation types is limited. The limitations of existing methods to estimate UCS promoted the development of data-driven solutionstoestimateUCS.Thedata-drivenmethodsincludebutarenotlimitedtobasicregression, machine learning, and deep learning algorithms. Data-driven methods to identify patterns in the datatoestimategeomechanicalparametersareconsideredtobeimplementedfordrillingoperations. This study proposes methods to assist safe and successful drilling operations while eliminating the need for coring, saving a vast amount of time and money by estimating UCS from drilling parameters instantaneously. The goal is to develop a machine-learning algorithm to analyze and process high-frequency data to estimate UCS instantaneously while drilling, allowing safer and more efficient drilling operations. The drilling data used to train, validate, and test the machine learning model is re-purposed from data collected during drilling in a previous study. The algorithm consists of a data processing method called Principal Component Analysis (PCA) to indicate the importance of each parameter by quantifying their variance contribution. Random Forest machine learning algorithm is utilized iii
Colorado School of Mines
ACKNOWLEDGMENTS I want to acknowledge the help of people and organizations that made this research possible by contributing to my research, studies, and daily life. First and foremost, I would like to thank my sponsor, MTA (General Directorate of Mineral Research and Exploration of Turkey), for financially sponsoring me through my studies to achieve excellence in the exploration of Geothermal resources in Turkey. I want to remind the memories of Dr. Tutuncu and thank her for opening the doors of CSM to me. She will always be remembered for her legacy here at CSM. I will remember her as a hard-working, deeply knowledgeable, kind, and lovely advisor. I want to thank Dr.Eustes for supporting me through all my studies. His trust in me motivated me to complete my research. I want to thank him for being there to answer my questions and for his wisdom. I want to express my most profound appreciation to my committee members for their help and for making this research possible. I am honored to gain an opportunity to learn from professors and staff here at CSM. I would like to thank Dr. Eustes, Dr. Ozkan, Dr. Prasad, Prof. Crompton, and Dr. Fleckenstein for their classes. I would like to thank all the Petroleum Engineering Department members, especially Denise and Rachel, for their support throughout my studies. Sincere thanks to my friends, Hazar, Deep, Santiago, Mansour, Mansour Ahmed, Val, Gizem, and Roy, for supporting me through all hardness. Thanks to you guys, I look at life from a broader perspective. Special thanks to Nehra for her company, love, and support that motivated me to complete my studies and made my life better in every way. Thank you for being there for me. Finally,Iwouldliketothankmyparents,AlimeandAbdullah,forbeingmyteachersthroughout my life. Everything I achieved was thanks to you two. xiii
Colorado School of Mines
CHAPTER 1 INTRODUCTION Historical data has always been an essential part of the oil and gas industry. The industry has become data-intensive with the recent advancements in data collection due to more durable and reliable sensors. However, the amount of data utilized to improve the efficiency of future operations is still a fraction of the data collected. The oil and gas industry is becoming more aware of the potential uses of these data. The utilization of data is being recognized as the most efficient method for reducing cost by increasing operational efficiency and creating safer, more sustainable developments (Løken et al. 2020). Specifically, the drilling industry has started investing more into the automation of drilling operations due to efficiency, safety, and cost concerns. In the last decade, the increasing computa- tional powers and the digitization of rig parts have allowed the industry to utilize machine learning algorithms (ML) for most drilling operations. By implementing data-driven solutions through ML algorithms,theindustryisworkingonbuildingautomateddrillingsystemsthatcanconductdrilling operations without human input or recommend an efficient solution for the safety of operations. One objective of the drilling industry is to increase the efficiency of drilling operations by reducing capital and operational expenses with the implementation of data-driven solutions. Knowing sub- surface conditions and geomechanical properties is essential to achieving this objective. Especially by gaining more knowledge about geomechanical properties, wellbore stability can be improved while drilling by avoiding hole collapse, stuck pipe, tight hole, kicks, and loss circulation. Drilling parameters have been recognized and used as an indicator of formation parameters, and estimating geomechanical properties from drilling parameters has been stated as an important topic, with studies conducted since the late 1950s (Combs 1968; Cunningham and Eenink 1959). Early studies completed by Bourgoyne and Young (1974) showed that pore pressure could be de- termined from drilling parameters with 1 lb/gal standard deviation on Gulf Coast, and Majidi et al. (2017) observed similar results in estimating formation pore pressure from MSE and drilling parameters from a study with a similar intent. Some of these models for pore pressure estima- tion by Jorden and Shirley (1966) as well as Rehm and McClendon (1971) are still being used in 1
Colorado School of Mines
the industry; thanks to their practicality. The objective of these studies is to indicate the impor- tance of estimating formation parameters fo and efficiency of the operations. Likewise, Unconfined Compressive Strength (UCS) has been known as an essential parameter as it is a key input to avoiding possible wellbore failures by implementing a robust mud weight window and deciding an aggressiveness of bit (Nabaei et al. 2010). The Unconfined Compressive Strength is the maximum compressive strength that a cylindrical rock sample can withstand under unconfined conditions. The UCS is also known as a uniaxial compressive strength because the compressive stress is applied along only one axis while measuring therockstrength. Theimpactofrockhardness,alsoknownasrockstrength,ondrillingperformance has always been an important issue and has been investigated since the early 1960s (Spaar et al. 1995). Inaddition,unconfinedcompressivestrengthisoneoftheessentialparameterswhendeciding on bit aggressiveness (Spaar et al. 1995). Theearlystudiesindicatedastrongcorrelationbetweenrockhardnessanddrillingperformance, and it is also observed that other drilling parameters such as weight on bit, revolution per minute, and bit type are required to predict the drilling efficiency among rock hardness measurements (Gstalder and Raynal 1966). The study indicates that estimation or measurement of UCS is essential to avoid wellbore stability problems while drilling. In addition, a study conducted by Spaar et al. (1995) shows a strong correlation between the formation drillability with UCS and friction angle as these parameters are essential for bit selection and the selection of appropriate aggressiveness for a bit can improve overall drilling performance substantially. The empirical equations derived from well logging tool readings and rock strength tests run in laboratory conditions are the most common methods to estimate UCS. However, data-driven solutions to estimate these parameters are becoming more common as these methods are getting more robust thanks to studies conducted to observe their veracity and versatility with available geomechanical and drilling data. Also, an exponential increase in the number of drilling operations conducted in unconventional reservoirs brought the need for a more sophisticated and cheaper method to estimate geomechanical parameters as these reservoirs commonly have non-linear be- havior and coring in a horizontal section of the wells drilled through the unconventional reservoirs is harder to conduct. These reasons indicate a need for a faster and cheaper method to estimate geomechanical parameters. 2
Colorado School of Mines
1.1 Motivation This thesis proposes to build a data-driven solution to estimate UCS instantaneously from drilling parameters by utilizing Random Forest regression model. The reviewed studies conducted by a vast amount of scientists indicate that the data-driven methods can improve the efficiency of operations in various ways by introducing sophisticated solutions such as predicting ROP (Hegde et al. 2015), estimating drilling optimization parameters (Nasir and Rickabaugh 2018), indicating the development of dominant water channels (Chen et al. 2019), predicting casing failures (Song and Zhou 2019), and predicting possible drilling incidents (AlSaihati et al. 2021). This study is solely motivated to provide key input parameter to avoid potential wellbore stability problems and drilling accidents by indicating rock strength changes within the formation. Furthermore, the final model from this study can be adjusted and developed to integrate into a fully automated drilling system as instantaneous UCS pattern indicator, which will be one of the essential steps for the next level of drilling automation. 1.2 Objectives The main objectives of this thesis are: • Develop a machine-learning model that can be trained to estimate unconfined compressive strength from drilling parameters instantaneously. • Provide changes in Unconfined Compressive Strength based on drilling parameters instanta- neously to avoid possible wellbore failures and drilling accidents. • Utilize principal component analysis to analyze feature importance using available drilling data. • StudytheimplementationofRandomForestregressionalgorithmtobuildarobustregression model to estimate certain geomechanical parameters (i.e., UCS, and Mechanical Specific Energy). 1.3 Thesis Organization This thesis consists of six chapters. The summary of each chapter is presented as follows: 3
Colorado School of Mines
CHAPTER 2 OVERVIEW 2.1 The Unconfined Compressive Strength Theunconfinedcompressivestrengthcanbedefinedasthemaximumamountofforceacylindri- calrocksamplecanwithstandunderunconfinedconditions. Themostcommonmethodstoestimate UCS are laboratory experiments and empirical equations derived from well logging tool readings. The laboratory experiments are conducted by using a testing setup that measures the maximum stress a sample can withstand; this method is referred to as a direct method to measure UCS, and the empirical equations derived from well-logging tool readings are referred to as an indirect method to estimate UCS (Ceryan et al. 2013). The American Society for Testing and Materials (ASTM) and the International Society for Rock Mechanics (ISRM) standardized the laboratory testing procedures that should be followed to estimate UCS. The laboratory testing procedures can vary with respect to stress distribution created around the sample. These laboratory testing methods use compressive, tensile, and triaxial stress distributions for different scenarios. The most common test method to measure the rock strength is called the uniaxial compressive test method, which is determined by applying compressive stress to the sample vertically until the sample fails (Brook 1993). The stress distributions for different test methods are shown in Figure 2.1. Themostcommonlimitationwhileconductingtheselaboratoryexperimentsisthequalityofthe coresample, asthebrokenorchippedsamplewillresultinachangeinstressdistributionwithinthe sample. The standardization of core sample preparation and test procedures by ASTM includes a detailed description to ensure accurate measurements. For example, according to American Society for Testing and Materials (2014), the minimum diameter of the test sample should be approximately 47-mm, and the length to diameter ratio of the test sample should be between 2.0:1 and 2.5:1 to satisfy the criterion in the majority of cases. As mentioned before, in addition to laboratory experiments, several empirical equations were derived to estimate the rock strength, as coring operation required to obtain rock samples for these experiments is time-consuming and expensive. The empirical equations can vary to the parameter 5
Colorado School of Mines
Most laboratory testing methods used to estimate the UCS are accurate if the test sample fits within the predetermined standards, but these methods are destructive and time-consuming. On the other hand, the applicability of empirical equations to estimate UCS is limited. The empirical equations could be a better option if the time is limited. However, the accuracy of empirical equa- tions is as good as the accuracy of the well-logging tool readings, which can introduce inaccuracy to the estimations. The limitations of laboratory testing methods and empirical correlations moti- vated the industry to develop an additional method to estimate UCS. With this motivation, the oil and gas industry has started to conduct studies to implement data-driven solutions as a suitable option, among other methods, to estimate parameters that require significant time and budget to measure (i.e., UCS). Especially, gaining even a limited amount of knowledge about UCS became necessaryasUCSchangescouldhelpindicatepotentialwellborestabilityproblemsinadvance. The wellbore stability and drilling problems cost the industry a vast amount of money every year since wellbore stability problems can lead to stuck pipes, stuck tools due to differential sticking, and excessive mud losses due to tensile fractures. Some of these wellbore stability problems, such as tensilefracturesandabreakout, canoccurduetoimpropermudweightwindow(i.e., excessivemud weight, low mud weight). Excessive mud weight can lead to tensile fractures, which will cause loss circulation and increase the chances of differential sticking, whereas stress around the wellbore will cause breakout if the mud weight can’t withstand the compressive strength of the rock (Al-Wardy and Urdaneta 2010). The study conducted by Al-Wardy and Urdaneta (2010) indicated that the time required to deliver a well located in North Oman could be reduced from 36.8 days in 2009 to an average of 30.1 days in 2010 by understanding the geomechanics of the field. A better understanding of geomechanics is achieved by building a geomechanical model of the particular area and complet- ing wellbore stability analysis by using vertical stress values from density logs, elastic properties and rock strength through DSI logs, minimum and maximum horizontal stress values from Min- frac/XLOT data, and stress orientation from BHI image logs. In addition to fracture tests and datacollectedfromlogs, theavailableformationanddrillingparameterssuchasporepressure, mud weights, drilling reports, and wellbore trajectory are used to complete the geomechanical model and wellbore stability analysis. This study showed that understanding geomechanics of even a par- ticular area of a field can help reduce wellbore stability problems considerably and improve overall 7
Colorado School of Mines
drilling performance. Also, in the study, it is stated that there is a sensitive correlation between the wellbore stability problems with UCS and minimum horizontal stress. A similar study was conducted by Klimentos (2005) to optimize drilling performance by provid- ing optimum drilling parameters and to estimate pore pressure values and wellbore stability plan through a geomechanical model. The study focuses on optimizing drilling performance in deep- waterwells,especiallywhiledrillingshalyformation. Toachievetheobjectiveofprovidingoptimum drilling parameters, the initial Mechanical Earth Model (MEM) is developed by using well logging tool readings, mud-logs, and drilling information. Then, a proper match between logs is completed to indicate the lithology and porosity of sections. Later, the overburden stress is estimated by integrating density log readings with the MEM. The exponential extrapolation model is used to estimate the missing values for the sections where density log readings were missing to study the overall rock strength through in-situ stresses from overburden and pore pressure. After combining previous efforts with estimated pore pressure of shaly formation through compaction theory and pressure/sonic log readings, the final MEM is completed. It is indicated that the determination of MEM and using optimum mud weight windows minimized washouts and loss circulation, and it allowed them to understand better the necessity of using casing strings. The results indicated that drilling performance could be improved by understanding the geomechanics of formation as the number of days to drill and construct a well was reduced by 15 days, and $4+ million was saved on total drilling cost. ThestudiesontheimportanceofgainingknowledgeaboutUCSondesignphasesofdrillingand completion operations are conducted by Brehm et al. (2006) and Al-Awad (2012). Brehm et al. (2006) completed a case study on Shenzi Field regarding the anisotropic behavior of the formations and the impacts on wellbore stability. Then, Al-Awad (2012) conducted a study that focused on the simple correlation between UCS and apparent cohesion, and throughout the study, impacts of wellbore stability issues as a result of lack of accuracy on rock strength estimation are included. Furthermore, comparison studies are conducted to question the accuracy of empirical equations to prove the versatility and veracity of these solutions. The research conducted by Meulenkamp and Alvarez (1999) compared the performance of empirical equations and used Machine Learning (ML) algorithmstoestimateUCSofdifferentrocksamplestoindicateifMLcanbeavaluabletoolforthe industry. Then, Changetal.(2006)conductedastudyontheaccuracyofempiricalequationsusing 8
Colorado School of Mines
a vast range of data and studied the applicability of 31 different empirical equations. Yurdakul et al. (2011) also conducted a similar study comparing the accuracy of the simple regression model and Artificial Neural Network (ANN) model in estimating UCS of sedimentary rock samples from 17 different regions of Turkey. Furthermore, Barzegar et al. (2016) expanded the coverage of the study and compared the per- formanceofdifferentMachineLearning(ML)andDeepLearning(DL)models. Later, Negaraetal. (2017) introduced elemental spectroscopy to the prediction model and searched for the potential impact of grain size on UCS by using the ML and DL model. The reviewed studies showed that data-driven solutions are becoming more robust. Brehmetal.(2006)completedacasestudyonthewellborelocatedattheShenziFieldinGreen Canyon blocks 653 and 654 regarding the wellbore stability issues. The study focused on building a complex geomechanical model for wellbores where the main problems are anisotropic failure and lost circulation. The study indicates the importance of using complex and comprehensive geomechanical models while drilling the wellbore with weak rocks and overpressured zones. This combination could limit the available mud weight window, and cause wellbore stability problems if the geomechanical model does not explain these phenomena in detail. The study also states that basic geomechanical modeling, built using the earth’s mechanical properties and in-situ earth’s effectivestressesoftheregion,bringsanovergeneralizedapproachtowellborereactionwhiledrilling a formation where anisotropic failure occurs as a consequence of weakly bedded rocks. It is also statedthatthesebasicgeomechanicalmodelscanbeimprovedtounderstandwellborefailuresbetter if it is applied correctly by building the model with accurate data. The complexity of these models is directly correlated with the accuracy quantified of the mechanical properties (i.e., pore pressure, in-situ stress magnitudes, stress orientation, rock strength). Further discussions showed that these wellbore stability problems could be avoided if the mud weight used while drilling at Shenzi was updated based on the anisotropic behavior of shale reservoirs. The model built includes changes in in-situ stress magnitudes, stress orientation, and UCS estimations. The results indicated that the significant wellbore stability and lost circulation problems in previous exploration operations are substantially reduced and turned into manageable drilling problems. Al-Awad (2012) conducted research focusing on the correlation between UCS and the apparent cohesion of rocks. The study also points out how important it is to know rock strength, aka UCS, 9
Colorado School of Mines
before designing drilling and completion operations to avoid possible wellbore stability problems such as sloughing shale, stuck drill pipe, tensile fractures, and breakouts. Also, in the study, it is mentioned that possible wellbore stability problems in producing wells such as sand production, perforation instability, subsidence, mechanical damage and how these problems can be foreseen in the design stage if rock strength is known. The correlation model between apparent cohesion and UCS is developed using available data from 300 different rock samples. The results showed that a simple correlation between rock apparent cohesion and UCS can be developed, and the correlation can estimate the rock mechanical properties with a 10% average error. The research study conducted by Meulenkamp and Alvarez (1999) compared the accuracy of estimation of UCS values by utilizing regression techniques, aka empirical correlations, and Neural Network. In the study, the data set contains records of 194 rock samples ranging from weak sandstones to very strong granodiorites, and the Equotip hardness tester is used as an index test forrockstrengthproperties. Thecomparisonwascompletedbetweenthreedifferentmethods: curve fitting, multivariate regression, and NN algorithm. The coefficient of determination (R2) of NN algorithm trained by this data set is determined as 0.967, while R2 measured 0.957 for Multivariate Regression relation and 0.910 for curve fitting relation. The results indicated that 0.967 of actual UCSvaluesstaywithintheregressionlinefitbyNN.EventhoughR2valuesforNNandMultivariate Regression relation are similar, the results would possibly change if a more extensive data set is used. Also, in the study, it is observed that the statistical relations underestimated high UCS values and overestimated low UCS values as the statistical relations were based on the mean of all predictions. Even though ML has limitations, the high accuracy of predictions indicates that it is possible to develop algorithms and implement them for field applications. Furthermore, the study indicated that ML algorithms reduce the cost and time to derive empirical correlations or conduct destructive experiments. Chang et al. (2006) reviewed and summarized the empirical equations that are derived to estimatetheunconfinedcompressivestrengthandinternalfrictionangleofsedimentaryrock(shale, limestone, dolomite, and sandstone) by using physical properties (such as porosity, velocity, and modulus). The author describes the importance of deriving efficient empirical correlations by pointing out the difficulty of retrieving core samples of overburden formations, where wellbore instability problems generally occur. In the study, overall, 31 empirical equations are reviewed, 10
Colorado School of Mines
and it is observed that most of the equations are unique for certain data gathered from a specific location, while some of the equations perform well. The empirical equations summarized for the prediction of UCS in sandstone vary by input values such as interval transit time, aka P-wave velocity (Fjær et al. 1992; McNally 1987; Moos et al. 1999), porosity (Vernik et al. 1993), and modulus (Bradford et al. 1998). Also, the empirical equations to predict UCS values in shales from porosity (Horshud 2001; Lashkaripour and Dusseault 1993), velocity (Horshud 2001; Lal 1999), and modulus (Horshud 2001) are reviewed and listed in the study. An example of well-performing equations is a relation between strength and porosity for sandstone and shale. The study also emphasized that the velocity readings were from dry rock samples, which causes a lower estimated value of UCS because of the inaccuracy introduced by the difference between dynamic and static moduli. However, it is also noted that the empirical equations derived by using the laboratory results are sufficient to estimate the lower boundary for UCS. The study conducted by Yurdakul et al. (2011) compared the predictive models for UCS of carbonate rocks from Schmidt Hardness. The Schmidt hammer that was initially designed to measure the strength of concrete can also be used to predict rock strength. The test considers the distance traveled by the energy transferred by the spring, and it measures the Schmidt hardness value based on the percentage of initial extension of the spring. The study compares the prediction results from the first-degree polynomial simple regression model and the artificial neural network (ANN) based model. The data set for this study was collected from 37 different natural stones collectedfrom19differentnaturalprocessingplantsfromdifferentcitiesinTurkey. Thecomparison between models was made considering Variance account for (VAF), coefficient of determination (R2), and root mean square error (RMSE). A lower value of RMSE indicates a more accurate prediction, while a lower value of VAF indicates a less accurate prediction in the model. Also, R2 valueshouldbecloserto1ifthemodelperfectlyfitstheavailabledata. TheobtainedVAF,RMSE, R2 indicators for the simple regression model are 12.45, 46.51, and 0.39, while 95.84, 7.92, and 0.96 for the ANN-based model. The results showed that the ANN-based model performs significantly better than the simple regression model, and an updated model can be developed to predict UCS values in sedimentary rocks. TheperformanceofvariousMLmethodsthatcanpredictUCSiscomparedintheresearchcom- pleted by Barzegar et al. (2016). The focus of the study is described as to evaluate the performance 11
Colorado School of Mines
of Adaptive neuro-fuzzy inference system (ANFIS), Mamdani fuzzy logic (MFL), Multi-layer per- ception (MLP), Sugeno fuzzy logic (SFL), and support vector machine (SVM) for the prediction of UCS of rocks in the Azarshahr area in northwest Iran. The fuzzy logic is described as an approach to computational methods that consider the degree of truth rather than absolute truth, and this allows the fuzzylogic toprovide arrays of possible true values. The multi-layer perceptionis acom- mon ANN approach for the prediction models that include layers to process data and learn from it. The adaptive neuro-fuzzy interference system is summarized as a feed-forward neural network function to check for the best fuzzy decision rule. The support vector machine is a soft computing learning algorithm mainly used for classification, pattern recognition, regression analysis, and pre- diction. The data set for the study include P-wave velocity, porosity, Schmidt rebound hardness, and UCS measured in the laboratory from 85 core samples. For the models, the data set is divided into two subsets: training (80% of data) and testing (20% of data). The performance of the models was assessed based on root mean square error (RMSE), mean absolute error (MAE), and coefficient of determination (R2). The results indicated that SVM model outperformed the other models with the lowest RMSE (2.14 MPa), MAE (1.351 MPa), and the highest R2 (0.9516). Negara et al. (2017) introduced elemental spectroscopy to consider grain size effects on UCS. The support-vector regression (SVR) is utilized to predict UCS. In this study, laboratory testing is the primary method to collect UCS data. The X-ray fluorescence (XRF) analysis is used for elemental spectroscopy. For the models, the data set was collected from the measurements of 35 core samples. The data gathered from seven of these core samples counted as outliers, and only 28 of them were used for the data set. The SVR is a supervised learning method that utilizes the necessary algorithms to analyze and recognize patterns. The quantitative measures to evaluate the performance of the model were the coefficient of determination and the mean absolute percentage error. The results indicated that the model built with SVR could predict UCS with a small error eventhoughasmallnumberofsampleswereusedtotrainthemodel. Also,aninfluenceofelemental spectroscopyonUCSpredictionisobserved. Thisinfluenceisdescribedastheeffectofgraindensity on rock strength. 12
Colorado School of Mines
2.2 Principal Component Analysis Principal Component Analysis (PCA) is a multivariate statistical method that can effectively reduce data dimensionality while preserving the variation within the data set. By preserving the variation,thePCAallowsMachineLearning(ML)algorithmstobetrainedwiththesameorsimilar patterns in the data set, which is essential for building a robust ML model. PCA is defined by Gupta et al. (2016) as a dimensionality reduction technique that uses an orthogonal transformation to convert a set of observations of possibly correlated or dependent variables into a set of linearly uncorrelated variables, which are called Principal Components. Kong et al. (2017) reported that PCA was first published by Pearson (1901) and developed by Hotelling (1933), and the modern applications and concepts are formed by Jolliffe (2002). PCA wasusedtoconductstudiesrelatedtohistorymatching,seismicinterpretation,patternrecognition, reservoir uncertainty evaluation, data compression, image processing, and high-resolution spectral analysis (Iferobia et al. 2020). Kong et al. (2017) explained feature extraction as a process of extraction measurements that are invariant and insensitive to the variations within each subspace of the data batch. The feature extractionisanessentialstepofthetaskofpatternrecognitionandthecompressionofdatabecause bothtasksrequirethesmallestpossibledistortionondatawhilereducingthenumberofcomponents. Also, feature extraction is a data processing technique that outlines a high-dimensional space to a low-dimensional space with minimal information loss. Principal component analysis (PCA) is one of the widely known feature extraction methods, while independent component analysis (ICA) and minor component analysis (MCA) are variants of the PCA. ICA is usually applied for blind signal separation, and MCA is commonly used for total least square problems (Kong et al. 2017). The scope of this study will be limited to the PCA, and the analysis will be conducted by using Python (Van Rossum and Drake 2009). However, comprehensive information regarding PCA is provided in Chapter 3 to clarify the concepts. PCA seems a complicated and time-consuming method once described in mathematical terms, but with the increase in computational power in the last decade, now it is possible to apply PCA to a million data points in less than a minute. With that, the applications of PCA started to become moreandmorecommonoverthelastdecadeintheindustry. TherecentapplicationsofPCAinclude 13
Colorado School of Mines
the spectral decomposition of seismic data, noise reduction of gamma-ray spectrometry maps, predicting possible casing failures, identifying the possible correlations between elemental data, estimation of dominant water channel development in oil wells, and estimation of geomechanical properties in unconventional reservoirs. Guo et al. (2009) utilized PCA to conduct a spectral decomposition of seismic data technique recently introduced to use as an interpretation tool that can help identify hydrocarbons. The merit of this interpretation technique is to develop an adequate form of data representation and reduction because, typically, the interpreter might generate 80 or more spectral amplitude and phase components. In the study conducted by Guo et al. (2009), 86 spectral components ranging from 5 Hz through 90 Hz were generated using an interpreter, and PCA was utilized to reduce the number of spectral components. It is observed that only three principal components in the total of 86 components were able to capture 83% of the variation in the seismic data. The results indicated that flow channel delineation could be mapped using RGB (Red, Green, and Blue) colors stack for the three largest principal components. de lima and Marfurt (2018) conducted a study with a similar motivation as Guo et al. (2009). In this study, PCA was used to reduce the noise of the gamma-ray spectrometry maps and reduce thenumberofcomponentsinthedatasetfromfourtothree. Initially, thegamma-rayspectrometer data consists of TC, K, eTH, and eU. The map displayed after implementing PCA and K-means clustering on PC1 and PC2 indicated a better correlation with the traditional geological map compared to the map created by only clustering of TC, K, eTH, and eU without PCA. Song and Zhou (2019) conducted a study to predict possible casing failures using PCA and gradientboostingdecisiontreealgorithms(GBDT).Thegradientboostingdecisiontreealgorithmis amachinelearningalgorithmthatutilizesdecisiontreesandcombinestheoutputofweakandstrong decision trees, aka boosting, to create a robust learning algorithm. This study applies the proposed method to the data set obtained from an oil field in mid-east China. The data set was created based on the parameters affecting the casing failure. Some of these parameters were the outside diameter of the casing, the thickness of perforation, and casing wall thickness. The PCA is used to reducedatasetdimensionality,whileGBDTisutilizedtodevelopthemachinelearningclassification model. The results indicated that using PCA with GBDT increased prediction accuracy on casing failure compared to classic methods (decision tree, Na¨ıve Bayes, Logistic Regression, multilayer 14
Colorado School of Mines
perceptron classifier). Also, it is stated that the algorithm created by using PCA and GBDT can successfully predict a timeline for preventive maintenance on offset wells (Song and Zhou 2019). EventhoughthePCAiscommonlyusedtoreducedimensionality,akathenumberofvariablesin thedataset,PCAhasotherpracticalapplications. ThePCAcanbeutilizedtoidentifyacorrelation between the components within the data. The study conducted byElghonimy and Sonnenberg (2021) focuses on observing a correlation between major and trace elements within the elemental data obtained from the Niobrara Formation in the Denver Basin. In the study, elemental data of samples is measured using a handheld XRF analyzer on full core from the Niobrara Formation. The variability of elemental concentrations is analyzed using PCA, and it is compared with the core facies to display the history of deposition and the conditions through the deposition process. The results showed that the application of PCA on the data set created by the integration of XRF measurements and core facies indications made a clear display of these elements in five major categories. Also,itisstatedthattheseidentifiedmajorcategoriescanactasanintermediaryforthe different deposited elements to indicate the history of deposition within the Niobrara Formation. Chen et al. (2019) studied a possible application of PCA as a recognition method for the dominant water channel development in oil-producing wells. The study was conducted using SZ Oilfield’s data located in Liaodong of Bohai bay. An evaluation index system was created to build a comprehensive evaluation method to consider every parameter. The parameters grouped in two main categories, dynamic response parameters and the parameters causing the channel to advance. The parameters considered for the evaluation index system are as follows: • Dynamic response parameters. – Dimensionless pressure index, – Pressure index, – Average water cut, – Water absorption profile coefficient, – Apparent water injectivity index increase, – Water injection intensity increase. • Parameter causing the channel to advance. 15
Colorado School of Mines
– Total water injection volume/unit thickness, – Apparent water injection intensity, – Viscosity of crude oil, – Effective thickness of reservoir, – Permeability contrast, – Average permeability. In this study, the focus of utilizing PCA is to test the objectivity of the method since the increasing number of parameters induces subjectivity to the recognition algorithm. An evaluation index system is created to analyze the causes of the dominant channel, and based on this system, an artificial learning method that recognizes the dominant channel is developed using PCA. The decision system was created based on the comprehensive evaluation index of the well group. If the calculated comprehensive evaluation index of the well group was higher than the average value, the well group was assumed to be developing a dominant channel. The results showed that the application of PCA to compute comprehensive evaluation index values reduced the subjectivity introduced by a large number of parameters. Also, it is stated that the method can provide technical support for further enhancing oil recovery by recognizing a pattern of dominant channel development in producing wells. Furthermore, another study conducted by Iferobia et al. (2020) shows that the significant num- ber of drilling operations conducted in unconventional reservoirs has shown that the prediction of UCS in the shale reservoirs is essential due to its complex and non-linear behavior. However, the modelswithasingleloginputparameter(sonic)wereinsufficientfortheseformations(Iferobiaetal. 2020). In the study, 21,708 data points of acoustic parameters are used to create a model with a principal component – multivariate regression, and the results indicated that the model could predict UCS values with 99% accuracy. Previous studies conducted on possible wellbore stability problems induced by a lack of knowl- edge of geomechanical properties indicated that the geomechanical modeling for drilling and com- pletion operations is open for improvement. Furthermore, the previous works conducted on the correlation between drillability and UCS show that the estimation of UCS is essential to increase 16
Colorado School of Mines
the overall drilling performance. ML algorithms are powerful tools for predicting UCS on forma- tions with complex and non-linear behavior. Recent studies show that data-driven solutions are reliable resources to support the decision-making process for drilling and completion operations. ML can be a powerful tool to estimate geomechanical parameters such as UCS in formations with complex and non-linear behavior. The literature review shows that implementing these methods can vastly increase overall drilling performance and efficiency of completion operations. 2.3 Tree-Based Algorithms Tree-based algorithms are supervised learning methods that are considered to be the most efficient machine learning method. Tree-based algorithms can be used for both regression and clas- sification problems. Also, they are suitable for non-linear relationships. The tree-based algorithms are simple yet powerful learning methods. The most popular tree-based learning algorithms are decision trees, gradient boosting, and random forest. Tree-based methods create and partition the feature space into a set of subspaces and fit a simple model (or constant) in each space. By using these feature spaces, a decision for each entry is given based on conditions set on each node of every tree. Thedecisiontreeisasupervisedlearningalgorithmwithadefinedtargetavailable. Thedecision tree is commonly used for classification problems. There are two types of decision trees: categorical variable decision trees and continuous variable decision trees. Categorical decision trees are built to solve classification problems, while continuous variable decision trees are commonly used to solve regression problems. Similarly, for both decision trees, the algorithm splits the data set into two or more homogenous subsets regarding the highest number of splitters or differentiators in input values. The decision tree algorithms are popular as they are easy to understand and useful for data exploration, but their tendency to overfit the data set is the most common difficulty of these methods. Regressionandclassificationtreesaresimplydecisiontreeswithmorenodes(orleaves),splitting the data set into smaller subsets. As it is mentioned before, the main differences between them are the type of input values and objective set while training the algorithm. While training, the splittingprocesscreatesafullygrowntreeforbothcases, andthesplitprocesscancauseoverfitting as the information given to each tree will be similar. The model parameters can be adjusted to 17
Colorado School of Mines
avoid overfitting, and validation techniques like pruning can be applied. The constraints of the tree size are simply the model parameters such as minimum samples for each node, minimum samples for each terminal node, maximum features to consider for split, and the maximum depth of the tree. The pruning essentially prevents the model from being greedy in the decision process. While pruning, the tree is grown to a large depth, and nodes (or leaves) that give negative values as output are removed. Another advantage of tree-based algorithms is that they are suitable for ensemble learning methods. Ensemble learning is developing a group of predictive models to improve model stability and accuracy, and they are simply a boost for decision tree models. A well-developed model should maintain the balance between bias and variance error. This balance between two errors is known as the trade-off management of bias-variance errors. This trade-off between variance and bias can be optimized in decision tree models by applying ensemble learning methods. The most common ensemble learning methods are bagging, random forest, and boosting. The bagging and random forestarethemostcommonmethodsappliedforclassificationproblemswhereagroup(committee) of trees cast a vote for the prediction. Boosting also involves a committee of trees that cast a vote for the prediction, but unlike random forest, a group of weak learners can evolve, and committee members cast a weighted vote. Bagging also utilizes multiple classifiers and combines them to develop a classifier to reduce the variance of predictions. These ensemble methods simply develop a group of base learners and combine them to establish a strong composite predictor. Even though there are many ensemble learning methods, Random Forest (RF) is decided to be the most suitable algorithm for this study, and it will be used to build a regression model to estimate UCS from drilling parameters. 2.3.1 Random Forest Random forest is an ensemble learning method that includes multiple classifiers and uses a combination of classifiers. Random forest is technically a modification of the bagging method that utilizes many decorrelated trees and retains the average of predictions as an output. Random Forest (RF) is similar to Bosting in terms of performance, but tuning the hyperparameters to avoid overfitting makes RF the best option for this study. The algorithm of RF for regression or classification is as follows, 18
Colorado School of Mines
• Assume z=1 to Z: • Then a bootstrap sample A* of size N is drawn from the training data. • RF tree T is grown around the bootstrapped data by following the steps below for each z terminal node of the tree until the predetermined node size is achieved. – n number of variables are selected from the p variables. – The best variable point among the n is picked. – The node is split into two daughter nodes. • The output of ensemble tress comes as {T }Z z 1 The prediction at a new point x; Z 1 (cid:88) Regression : fZ(x) = T (x) (2.4) rf Z Z z=1 Classification : If the class prediction of thezth random forest tree assumed as C (x) b (2.5) ThenCZ(x) = majority vote{C (x)}Z(x) (2.6) rf Z 1 The process of bagging trees creates identically distributed variance within the predictions, which means that the bias of a single tree (bootstrap) is the same as the bagged trees. Therefore, the only hope of improving prediction accuracy is to reduce the variance. However, RF adaptively develops the trees to remove bias as the distribution between trees is not identical. IfthevarianceisexplainedbythetermsoftheRFalgorithmdescribedabove. Then,theaverage variance of Z independent identically distributed random variables, each with σ2 variance, is 1σ2. z If it is assumed that the variables are identically distributed but not necessarily independent from each other, with a positive correlation ρ , the average variance is 1−ρ ρσ2+ σ2 (2.7) Z The second term of the average variance disappears with the increasing number of random variables (Z). Hence the averaging greatly helps to reduce the variance. The focus of RF is to 19
Colorado School of Mines
reduce the variance of bagged trees by reducing the correlation between them without increasing variance by a high margin. RF is one of the most common ML methods that help solving to both regression and classifica- tion problems. Also, the studies reviewed indicated that the common problems of the oil and gas industry, specifically the drilling industry, can took the opportunity to create models with RF that can predict essential parameters or indicate the drilling incidents. Hegde et al. (2015) conducted a study to oversee the possible applications of statistical learning techniques to predict the Rate of Penetration (ROP) values. The statistical learning methods were trees, bagged trees, and Random Forest (RF), and the models are evaluated based on their performance by comparing Root Mean Square Error (RMSE) values. Also, this study introduces a predictive model called Wider Windows Statistical Learning Model (WWSLM), which considers manyinputparameterssuchasWOB,RPM,anddepthtocompensatefortheeffectofhighlithology variation on drilling parameters by utilizing trees and random forest. Another advantage of the model built by Hegde et al. (2015) is that the blind test subset is split from the complete data set and ensured that the model never learns from it, which reduces the probability of overfitting. Only surface drilling parameters are included in the training, validation, and test sets as input data. The tree method is described as a technique used for classification and regression purposes. Baggingissummarizedasjointlyusingbootstrappinganddecisiontreestocompensateforthehigh lithology variation. Bootstrapping is a method to reduce variation by repeatedly sampling from the same data set that yields multiple training sets. Random forest is described as a method that utilizes bootstrapping to increase the number of training subsets and combines it with decision trees. The main difference between RF and bagging is that RF uses a random subset of predictors to build trees, which eventually reduces the variance of Statistical Learning Model (SLM). The results indicated that RF provides the most accurate ROP predictions with RMSE of 7.4, which is three times lower than RMSE values of other models. It should be noted that the ROP values of the training data set ranging from 20 to 120 ft/hour. The optimization of drilling parameters by using Random Forest (RF) regression algorithm is studied by Nasir and Rickabaugh (2018). The study aimed to optimize drilling parameters to increase bit life by reducing the wear and tear while maximizing ROP and minimizing Mechanical Specific Energy (MSE). In the study, drilling parameters from the wells within 20 miles radius are 20
Colorado School of Mines
used as a training data set. Also, it should be noted that the drilling parameters were collected while using the same motor and bit for the entire vertical interval of these wells. The key drilling parameters investigated were surface RPM, mudflow rate, WOB, and formation type. A total of six different formations are encountered while drilling these wells. The models were trained to estimate ROP using Linear regression, Support Vector Regression (SVR), Random Forest, and Boosted Tree. The data is split into two categories, the training set (80% of the data set) and the validation set (20%). Performance indicators for the model were root mean square error (RMSE), mean absolute error (MAE) and mean average error percentage (MAEP). The results showed that RF could estimate the ROP values with 12% error after tuning the hyperparameters, while the other models had a higher percentage of errors. Also, the author stated that the model could be possiblyimprovedbyintroducingkeyvariablesimpactingdrillingperformance, suchascompressive strength and gamma-ray response. AlSaihati et al. (2021) studied the possible application of Random Forest (RF) to predict the anomalies in Torque and Drag (T&D) values to indicate a possible drilling incident such as a stuck pipe in advance. While building RF model, a pipe stuck case that took place while drilling a 5-7/8-inch horizontal section of 15,060 ft depth well is considered, and the model was built to indicate possible problems in this interval. It is assumed that the stuck pipe occurred while drilling the reservoir contact zone (5000 ft) because of high T&D, insufficient transfer of weight-on-bit (WOB), and poor hole cleaning. The pipe stuck occurred at a measured depth of 14,935-ft while tripping out after circulating the hole clean. The drilling parameters used as a variable to train RF model include hook load, flow rate, rotation speed, rate of penetration, standpipe pressure, and torque readings at the surface, and the data covers the timeline starting from the beginning of the horizontal section to one day prior to the stuck pipe incident. The number of data points used to train the RF algorithm is 7,186, 80% of the total data, and 1,797 data points used to test the algorithm, which refers to 20% of the total data. The MSE value of 0.06 and R value of 0.97 indicated that the RF algorithm could predict the anomalies accurately. Also, it is stated that the model detected anomalies for nine hours consecutively prior to the pipe stuck incident. Overall, the reviewed studies indicated that the performance of RF machine learning algorithm could be a robust tool, and its implementation as a solution for common problems is promising. Also, the studies indicated that RF is one of the most common algorithms to build efficient data- 21
Colorado School of Mines
CHAPTER 3 BACKGROUND INFORMATION ABOUT DATASET AND PCA Data collection and processing are essential for every project studying possible data-driven solutions for the common problems. In this research, the drilling data used to train a regression model to estimate UCS is collected for another research project with similar intents (Joshi 2021). In this chapter, the background information of the data set is provided. Also, the unique nature of this data set is discussed in detail. In addition, the fundamental principles of PCA are discussed in this chapter. Then, the implementation of PCA to indicate feature importance and explained variance by principal components are given in detail with the results. 3.1 Background Information about the Dataset As mentioned before, the data set needs to meet certain criteria to be used to train a machine learningalgorithm. Inotherwords, theaccuracy, consistency, uniquenessofdata, andcompleteness and validity of the data set should meet certain criteria. The possible problems in a data set can be duplicated, outlier, missingdatapoints, andinaccuratedatareadings. Initially, thedatasetusedin thisresearchwascollectedbyDr. Joshi,anditwascollectedwhiledrillingthroughsamplesprepared in laboratory conditions. While the analog samples were prepared to mimic field conditions while drilling, cryogenic samples were prepared to mimic various extraterrestrial conditions in Lunar. Joshi (2021) collected the drilling data using a laboratory experiment setup. The properties of these analog and cryogenic samples are given in Table 3.1 and Table 3.2, respectively. The initial planwastousethecompleterawdatasetcollectedfromeachofthesesamples, butunfortunately, a drilling data collected only from analog samples could be utilized to train the RF regression model for this study. A discussion regarding the uniqueness and challenges of the data set is provided later in this chapter. The final form of these cellular concrete samples, aka analog samples, before and after drilling while curing and demolding is presented in Figure 3.1. The setup is built on a frame with a 3-phase AC motor to transfer power to the masonry bit. The movement of the masonry bit is supplied by the stepper motor and guide rails. The experimental setup was controlled by a variable frequency drive and stepper motor. The picture of 23
Colorado School of Mines
3.1.2 Weight on Bit (WOB) Weight on bit calculations was based on the subtraction of axial force measurement on air for each bit from axial force measurements while drilling. The reference axial force of each bit on-air is measured and hardcoded to the model to automate the WOB calculation process (Joshi 2021). The WOB at any point is calculated as: WOB = Total axial force −Total axial force (3.2) i i air 3.1.3 Rate of Penetration (ROP) The rate of penetration is calculated by the difference in measured depth at each time interval. The ROP can be calculated by the following equation. Drilling Depth −Drilling Depth i i−1 ROP = (3.3) i time −time i i−1 3.1.4 Normalized Field Penetration Index (N-FPI) The field penetration index was originally defined to evaluate the energy required to overcome therockstrengthforatunnelboringmachine(TarkoyandMarconi1991),(Hassanpouretal.2011). By the original definition, the FPI is calculated as: kN F FPI = (cutter) = n (3.4) mm P rev Where, F cutter load (Cutter Load (kN)/Number of Cutters) and P is the penetration of cutter n per revolution (ROP/RPM). The cutter load (F ) can be replaced by the normalized drilling force (WOB) to calculate n Normalized Field Penetration Index (N-FPI) for the drilling systems. The cutter load F for n drilling systems can be calculated as: N WOB F = ( ) = (3.5) n mm2 (π)d2 4 The final unit of N-FPI will be ( N )/(mm) mm2 rev 26
Colorado School of Mines
3.1.5 Mechanical Specific Energy (MSE) Teale (1965) defined the Mechanical Specific Energy (MSE) as the amount of energy required to excavate a unit volume of rock. MSE can be calculated as: MJ WOB(N) 103Torque(N.m)×RPM MSE = ( ) = + (3.6) m3 (π)d2(mm2) (π)d2(mm2)×ROP(mm/min) 4 4 3.2 Uniqueness with Dataset and Challenges Every data set collected includes noise and outliers that need to be removed before using for any machine learning or deep learning application. The process of filtering these noise and outliers from raw data before labeling is called processing. The challenges introduced by raw data can vary, and its uniqueness is the key to building a robust machine learning model. Likewise, the data set used in this study needed processing before being used as training data for machine learning models, but the challenges were not limited to only noise. Initially, torque values were planned to be measured using a bridge-based shaft-to-shaft torque sensor placed between shaft and bit, but torque readings were incredibly noisy due to electromagnetic interference, mechanical noise, and ambient noise (Joshi 2021). The electromagnetic interference caused by the three-phase AC motor is located extremely close to the torque sensor. The vibrations in the bit while drilling result in mechanical noise in torque readings, while ambient noise from The Earth Mechanics Institute (EMI) at the Colorado School of Mines campus was detected by torque sensors. Filtering the noise introduced by the environment and experimental setup was removed in a sampled data, but this process required a vast amount of time as signal smoothing, outlier removal, and significant signal filtering needed to be applied to the complete data set. After consideration, Joshi (2021) decided to build a regression model to predict torque values from other drilling parameters and Variable Frequency Drive (VFD) outputs, as the filtering process was computationally expensive andrequiredsignificantsignalconditioning. Toclarifyhowtorquedataiscalculated,anarchitecture ofthealgorithmcalled“TheLunarMaterialCharacterizationwhileDrillingAlgorithm”ispresented in Figure 3.3. 27
Colorado School of Mines
Figure3.3: ThearchitectureofalgorithmbuiltbyJoshi(2021)©2021byDeepR.Joshi.,reprinted with permission from Deep R. Joshi As it is described in Figure 3.3, a regression model used to calculate Torque values from VFD output and other drilling parameters right after raw data is classified as drilling and non-drilling data. Later, Torquevalueswereassumedtobesimilartotorquesensorreadings. Anotherchallenge introduced by this dataset is the lack of variation within UCS. Even though the complete data set contains approximately 55+ observations on eight different variables, only seven different UCS values are present. Initially, this was assumed to be a unique part of the data set. Later, it was understood that the lack of variability in target values while building RF regression model was causing significant overfitting and leading to high prediction accuracy while estimating UCS. A histogram graph of UCS within the complete data set is given in Figure 3.4. This issue and its impacts on this study are discussed in detail in Chapter 4. Initially, in the scope of this study, these predicted torque values were assumed as a unique part of the data set. However, the prediction of torque values possibly reduced the variation within the data set and impacted this study significantly with the contribution of other data set challenges. This combination led to changes in the path of this study. 3.3 Principal Component Analysis Brief information about Principal Component Analysis (PCA) is given in Chapter 2. PCA is defined as a dimensionality reduction method that utilizes orthogonal transformation to convert 28
Colorado School of Mines
Figure 3.4: Histogram of UCS - Complete Dataset a set of observations into a set of linearly uncorrelated variables (Gupta et al. 2016). As is de- scribed comprehensively in Chapter 2, PCA is commonly used to build data-driven solutions for the various problem caused by dimensionality. This section explains the fundamental principles and implementation of PCA to clarify the concept. In this study, the scope of PCA will be limited to feature importance indication and explained variance analysis in the data set. 3.3.1 The Concept of PCA Kong et al. (2017) stated that “the principal components (PC) are the directions in which the data have the largest variances and capture most of the information contents of data.”. The PCs are correlated with the eigenvectors inherent in the largest eigenvalues of the autocorrelation matrix of the data vectors. The expression of data vectors regarding the PCs is named PCA, while expressing the vectors regarding the MCs is named MCA. PCA or MCA is usually one-dimensional, but the actual applications have also been found to be multiple dimensional. The principal (or minor) components are referred to as the eigenvectors affiliated with r largest (or smallest) eigenvalues of the autocorrelation matrix of the data vectors, while r is the number of the principal (minor) components. The subspace covered by PC is called 29