73
12 August 2020 POLITECNICO DI TORINO Repository ISTITUZIONALE Combined finite-discrete element modelling of key instabilities which characterise deep-seated landslides from massive rock slope failure / Piovano, Giovanna. - (2012). Original Combined finite-discrete element modelling of key instabilities which characterise deep-seated landslides from massive rock slope failure Publisher: Published DOI:10.6092/polito/porto/2502740 Terms of use: openAccess Publisher copyright (Article begins on next page) This article is made available under terms and conditions as specified in the corresponding bibliographic description in the repository Availability: This version is available at: 11583/2502740 since: Politecnico di Torino

POLITECNICO DI TORINO Repository ISTITUZIONALE · Appendix A General formulation of conventional numerical methods A.1 Introduction The general formulation of continuum and discontinuum

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

12 August 2020

POLITECNICO DI TORINORepository ISTITUZIONALE

Combined finite-discrete element modelling of key instabilities which characterise deep-seated landslides from massiverock slope failure / Piovano, Giovanna. - (2012).

Original

Combined finite-discrete element modelling of key instabilities whichcharacterise deep-seated landslides from massive rock slope failure

Publisher:

PublishedDOI:10.6092/polito/porto/2502740

Terms of use:openAccess

Publisher copyright

(Article begins on next page)

This article is made available under terms and conditions as specified in the corresponding bibliographic description inthe repository

Availability:This version is available at: 11583/2502740 since:

Politecnico di Torino

Appendix A

General formulation ofconventional numerical methods

A.1 Introduction

The general formulation of continuum and discontinuum methods, as well as the simula-tion of fracture process using different numerical tecniques, described in this appendix,is based on the previous work made by Jing (2003).

A.2 Numerical methods in rock engineering

In rock engineering several modelling methods have been developed for the study anddesign of rock engineering structures. In the literature these approaches are divided intotwo main groups which differ for the type of representation of the problem used.

In the discrete representation, engineering problems are represented by an adequatemodel using a finite number of well-defined components. The behaviour of such com-ponents is either well known, or can be independently treated mathematically. Theglobal behaviour of the system can be determined through well-defined inter-relationsbetween the individual components (elements). In such problems, termed discrete, thediscrete representation and solution of such systems by numerical methods are usuallystraightforward.

In other problems, the definition of such independent components may require an in-finite sub-division of the problem domain, and the problem can only be treated using the

251

252 General formulation of conventional numerical methods

mathematical assumption of an infinitesimal element, implying in theory an infinite num-ber of components. This usually leads to differential equations to describe the systembehaviour at the field points. Such systems are termed continuous representation and haveinfinite degrees of freedom. To solve such a continuous problem by numerical methods,the problem domain is usually subdivided into a finite number of sub-domains (ele-ments) whose behaviour is approximated by simpler mathematical descriptions with fi-nite degrees of freedom. These sub-domains must satisfy both the governing differentialequations of the problem and the continuity condition at their interfaces with adjacentelements. This is the so-called discretization of a continuum. It is an approximation ofa continuous system with infinite degrees of freedom by a discrete system with finitedegrees of freedom. The continuum assumption implies that at all points in a problemdomain, the materials cannot be torn open or broken into pieces. Of course, at the micro-scopic scale, all materials are discrete systems. However, representing the microscopiccomponents individually is intractable mathematically and unnecessary in practice.

The concepts of continuum and discontinuum are therefore not absolute but relative,depending especially on the problem scales. Due to the differences in the underlyingmaterial assumptions, different numerical methods have been developed for continuousand discrete systems. The most commonly applied numerical methods for rock mechan-ics problems are:

continuum methods

• FINITE DIFFERENCE METHOD (FDM). The FDM method is the oldest numericalmethod, and it is the basis of the explicit approach of the DEMs. The FDM is adirect approximation of the governing Partial Differential Equations (PDEs) by re-placing partial derivatives with differences at regular or irregular grids imposedover problem domains. The solution of the system equation is obtained after im-posing the initial and boundary conditions;

• FINITE ELEMENT METHOD (FEM). The FEM is perhaps the most widely appliednumerical method in engineering because its flexibility in handling material het-erogeneity, non-linearity and boundary conditions. It is also the basis of the im-plicit approach of the DEM. The FEM requires the division of the problem domaininto sub-domains (elements) of standard shapes (triangle, quadrilateral, tetrahed-ral, etc.) with fixed number of nodes at the vertices and/or on the sides. Poly-nomial functions are used to approximate the behaviour of PDEs at the elementlevel and generate the local algebraic equations representing the behaviour of theelements. The local elemental equations are then assembled, according to the topo-logic relations between the nodes and elements, into a global system of algebraic

A.2 Numerical methods in rock engineering 253

equations whose solution then produces the required information in the solutiondomain, after imposing the properly defined initial and boundary conditions;

• BOUNDARY ELEMENT METHOD (BEM). The BEM requires discretization at theboundary of the solution domains only. The information required in the solutiondomain is separately calculated from the information on the boundary, which isobtained by solution of a boundary integral equation, instead of direct solutionof the PDEs, as in the FDM and FEM. It has greater accuracy over the FDM andFEM at the same level of discretization and is also the most efficient technique forfracture propagation analysis.

discontinuum methods

• DISCRETE ELEMENT METHOD (DEM). The DEM for modelling a discontinuum fo-cuses mostly on applications in the fields of fractured geological media. The DEMrepresents the fractured medium as assemblages of blocks connected by fractures.The equations of motion of these blocks are solved through continuous detectionand treatment of contacts between the blocks. The blocks can be rigid or be de-formable with FDM or FEM discretizations. Large displacements caused by rigidbody motion of individual blocks, including block rotation, fracture opening andcomplete detachments is straightforward in the DEM, but impossible in the FDM,FEM or BEM;

• DISCRETE FRACTURE NETWORK (DFN) method. The DFN method is an alternat-ive DEM for fluid flow in fractured rock masses. It simulates fluid flow throughconnected fracture networks, with the matrix permeability either ignored or ap-proximated by simple means. The stress and deformation of the fractures are gen-erally ignored as well. This method is conceptually attractive for simulating fluidflow in fractured rocks when the permeability of the rock matrix is low comparedto that of the fractures.

hybrid continuum/discontinuum models

• Hybrid FEM/BEM;

• Hybrid DEM/DEM;

• Hybrid FEM/DEM, and

• Other hybrid models.

254 General formulation of conventional numerical methods

A.3 Continuum methods

A.3.1 FINITE DIFFERENCE METHOD (FDM)

The FDM is the oldest numerical method to obtain approximate solutions to PDEs inengineering. The basic concept of FDM is to replace the partial derivatives of the object-ive function (e.g. displacement) by differences defined over certain spatial intervals inthe coordinate directions ∆x, ∆y, 4z, which yields a system of algebraic simultaneousequations of the objective functions at a grid (mesh) of nodes over the domain of interest(Figure A.1) (Wheel, 1996). Solution of the simultaneous algebraic system equations,incorporating boundary conditions defined at boundary nodes, will then produce therequired values of the objective function at all nodes, which satisfy both the governingPDFs and specified boundary conditions. The conventional FDM utilizes a regular gridof nodes, such as a rectangular grid as shown in Figure A.1a.

Using a standard FDM scheme, the so-called 5-point difference scheme (Figure A.1b),the resultant FDM equation at grid node (i, j) will be expressed as combinations of func-tion values at its four surrounding nodes. For a Navier equation of equilibrium for elasticsolids in 2D, the FDM equation of equilibrium at point (i, j) is given as

ui,jx = a1ui−1,j

x + a2ui,j−1x + a3ui,j+1

x + a4ui+1,jx + a5ui+1,j+1

x + a6Fi,jx

ui,jy = b1ui−1,j

y + b2ui,j−1y + b3ui,j+1

y + b4ui+1,jy + b5ui+1,j+1

y + b6Fi,jy (A.1)

where coefficients ak and bk (k = 1, 2, ..., 6) are functions of the grid intervals ∆x and∆y and the elastic properties of the solids, and Fi,j

x and Fi,jy are the body forces lumped

at point (i, j), respectively. Assembly of similar equations at all grid points will yield aglobal system of algebraic equations whose solution can be obtained by direct or iterativemethods. FDM schemes can also be applied in the time domain with properly chosentime steps, ∆t, so that function values at time t can be inferred from values at t− ∆t.

The fundamental nature of FDM is the direct discretization of the governing PDEs byreplacing the partial derivatives with differences defined at neighboring grid points. Thegrid system is only a convenient way of generating objective function values at samplingpoints with small enough intervals between them, so that errors thus introduced aresmall enough to be acceptable. No local trial (or interpolation) functions are employedto approximate the PDE in the neighborhoods of the sampling points, as is done in FEMand BEM.

The conventional FDM with regular grid systems does suffer from shortcomings,most of all in its inflexibility in dealing with fractures, complex boundary conditionsand material inhomogeneity. This makes the standard FDM generally unsuitable for

A.3 Continuum methods 255

progressive failure mechanism. This approach can beapplied to modelling the rock mass response toengineering actions in different field circumstances,leading to enhanced understanding and hence enablingthe engineer to design more coherently.

3. Numerical techniques for rock mechanics:

states-of-the-art

3.1. Finite Difference Methods

3.1.1. Basic concepts

The FDM is the oldest numerical method to obtainapproximate solutions to PDEs in engineering, espe-cially in fluid dynamics, heat transfer and solidmechanics. The basic concept of FDM is to replacethe partial derivatives of the objective function (e.g.displacement) by differences defined over certain spatialintervals in the coordinate directions, Dx; Dy; Dz; whichyields a system of algebraic simultaneous equations ofthe objective functions at a grid (mesh) of nodes over thedomain of interest (Fig. 8a) (Wheel, 1996 [9]). Solutionof the simultaneous algebraic system equations, incor-porating boundary conditions defined at boundarynodes, will then produce the required values of theobjective function at all nodes, which satisfy boththe governing PDFs and specified boundary conditions.The conventional FDM utilizes a regular grid of nodes,such as a rectangular grid as shown in Fig. 8a.Using a standard FDM scheme, the so-called 5-point

difference scheme (Fig. 8b), the resultant FDM equationat grid node ði; jÞ will be expressed as combinations offunction values at its four surrounding nodes. For aNavier equation of equilibrium for elastic solids in 2-D,the FDM equation of equilibrium at point ði; jÞ is givenas

ui;jx ¼ a1u

i1;jx þ a2u

i;j1x þ a3u

i;jþ1x þ a4u

iþ1;jx

þ a5uiþ1;jþ1x þ a6F

i;jx ;

ui;jy ¼ b1u

i1;jy þ b2u

i;j1y þ b3u

i;jþ1y þ b4u

iþ1;jy

þ b5uiþ1;jþ1y þ b6F

i;jy ; ð1Þ

where coefficients ak and bkðk ¼ 1; 2;y; 6Þ are functionsof the grid intervals Dx and Dy; and the elasticproperties of the solids, and Fi;j

x and Fi;jy are the body

forces lumped at point ði; jÞ; respectively. Assembly ofsimilar equations at all grid points will yield a globalsystem of algebraic equations whose solution can beobtained by direct or iterative methods. FDM schemescan also be applied in the time domain with properlychosen time steps, Dt; so that function values at time t

can be inferred from values at t Dt:The fundamental nature of FDM is the direct

discretization of the governing PDEs by replacing thepartial derivatives with differences defined at neighbour-ing grid points. The grid system is only a convenient wayof generating objective function values at samplingpoints with small enough intervals between them, sothat errors thus introduced are small enough to beacceptable. No local trial (or interpolation) functionsare employed to approximate the PDE in the neighbour-hoods of the sampling points, as is done in FEM andBEM. It is therefore the most direct and intuitivetechnique for the solution of the PDEs. The conven-tional FDM with regular grid systems does suffer fromshortcomings, most of all in its inflexibility in dealingwith fractures, complex boundary conditions andmaterial inhomogeneity. This makes the standardFDM generally unsuitable for modelling practical rockmechanics problems. However, significant progress hasbeen made in the FDM so that irregular meshes, such asquadrilateral grids (Perrone and Kao, 1975, [10]) andthe Voronoi grids (Brighi et al., 1998, [11]) can also beused. Although such irregular meshes can enhance theapplicability of the FDM for rock mechanics problems,however, the most significant improvement comes fromthe so-called Control Volume or Finite Volumeapproaches.

(a) (b)

Cell 8

Cell 7Cell 6

Cell 5

Cell 4Cell 3 Cell 2

Cell 1

lk

ji

P

∆x

∆y

i,j i+1,j

i,j+1

i-1,j

i,j-1

Fig. 8. (a) Regular quadrilateral grid for the FDM and (b) irregular quadrilateral grid for the FVM (after Wheel, 1995 [9]).

L. Jing / International Journal of Rock Mechanics & Mining Sciences 40 (2003) 283–353 293

Figure A.1: (a) Regular quadrilateral grid for the FDM and (b) irregular quadrilateralgrid for the FVM (Wheel, 1996).

modeling practical rock mechanics problems. However, significant progress has beenmade in the FDM so that irregular meshes, such as quadrilateral grids (Perrone andKao, 1975) and the Voronoi grids (Brighi et al., 1998) can also be used. Although suchirregular meshes can enhance the applicability of the FDM for rock mechanics problems,however, the most significant improvement comes from the so-called Control Volume orFinite Volume approaches.

A.3.1.1 Finite volume approach of FDM and its application to stress analysis

The Finite Volume Method (FVM) is also a direct approximation of the PDEs, but inan integral sense. An elastostatic problem with body Ω, is divided into a finite num-ber, N, of internal contiguous cells of arbitrary polyhedral (or polygonal in 2D cases)shape, called Control Volumes (CV), Ωk, with boundary Γk, of unit outward normal vec-tor nk

i , k = 1, 2, ..., N. The boundary Γk of Ωk is comprised of a number, Mk, polygonalside (faces or line segments), Γp

k , p = 1, 2, ..., Mk. Assuming isotropic, linear elasticityand using Gauss’ divergence theorem, the Navier–Cauchy equation of equilibrium interms of stress can be rewritten in terms of displacement as

256 General formulation of conventional numerical methods

N

∑k=1

[Mk

∑p=1

ˆΓp

k

tki dΓ +

ˆΩk

fidΩ

]=

N

∑k=1

[Mk

∑p=1

ˆΓp

k

σkijn

pj dΓ + Fk

x

]= 0 (A.2)

where Fki = ρgiVk is the body force vector of the CV of volume Vk lumped at its cen-

ter, ρ is the material density and gi is the body force intensity vector, such as gravityacceleration.

The task is to formulate the integrals into algebraic functions of the displacementsat nodes defining the boundary sides Γp

k of Ωk, which vary with different grid schemes.For an unstructured quadrilateral grid system (Figure A.1b), a typical cell P (CV), withits center at node P, has four sides (ij, jk, kl, li) and four nodes (i, j, k, l), surrounded byeight neighboring cells with center nodes I, J, ..., O. The integral terms in Equation A.2for the cell P are written in terms of displacement variables at the centers of cells (Wheel,1996), written as

Apupx + ∑

rArur

x + Bpupy + ∑

rBrur

y + FKx

Cpupy + ∑

rCrur

y + Dpupx + ∑ Drur

x + FKy (A.3)

where coefficients Ap, Ar, Bp, Br, Cp, Cr, Dp, Dr are functions of the cell geometry and theelastic properties of the solids, with r = 1, 2, ..., 8 running through the eight surroundingcells.

The FDM/FVM approach is therefore as flexible as FEM in handling material in-homogeneity and mesh generation. As a branch of the FDM, the FVM can overcome theinflexibility of the grid generation and boundary conditions in the traditional FDM withunstructured grids of arbitrary shape.

The FDM/FVM approaches are therefore specially suited to simulate non-linear be-havior of solid materials. The reason is its special advantage of no-matrix-equation-solving formulation and data structure, so that integration of non-linear constitutiveequations is a straightforward computer implementation step, rather than iterative prediction-mapping integration loops required in FEM.

At present, the most well-known computer codes for stress analysis for non-linearrock engineering problems using the FVM/FDM approach is perhaps the FLAC codegroup (ITASCA Consulting Group, 1993b), with a vertex scheme of triangle or quadri-lateral grids.

A.3 Continuum methods 257

A.3.2 FINITE ELEMENT METHOD (FEM)

The FEM requires the division of the problem domain into a collection of sub-domains(elements) of smaller sizes and standard shapes (triangle, quadrilateral, tetrahedral, etc.)with fixed number of nodes at the vertices and/or on the sides, the discretization. Trialfunctions, usually polynomial, are used to approximate the behavior of PDEs at the ele-ment level and generate the local algebraic equations representing the behavior of theelements. The local elemental equations are then assembled, according to the topologicrelations between the nodes and elements, into a global system of algebraic equationswhose solution then produces the required information in the solution domain, afterimposing the properly defined initial and boundary conditions.

The FEM has been the most popular numerical method in engineering sciences, in-cluding rock mechanics and rock engineering. Its popularity is largely due to its flexib-ility in handling material inhomogeneity and anisotropy, complex boundary conditionsand dynamic problems, together with moderate efficiency in dealing with complex con-stitutive models and fractures. All these merits were very appealing to researchers andpracticing engineers alike during early development in the 1960s and 1970s when themain numerical method in engineering analysis was the FDM with regular grids. Sincethen, the FEM method has been extended in many directions.

Basically, three steps are required to complete an FEM analysis:

1. domain discretization;

2. local approximation;

3. assemblage and solution of the global matrix equation.

The domain discretization involves dividing the domain into a finite number of internalcontiguous elements of regular shapes defined by a fixed number of nodes (e.g., triangleelements with three nodes in 2D and brick elements with eight nodes in 3D). A basicassumption in the FEM is that the unknown function, ue

i over each element, can be ap-

proximated through a trial function of its nodal values of the system unknowns, uji , in a

polynomial form. The trial function must satisfy the governing PDF and is given by

uei =

M∑j=1

Nijuji (A.4)

where the Nij are often called the shape functions (or interpolation functions) defined inintrinsic coordinates in order to use Gaussian quadrature integration, and M is the order

258 General formulation of conventional numerical methods

of the elements. Using the shape functions, the original PDF of the problem is replacedby an algebraic system of equations written as

N

∑i=1

[Ke

ij

] ue

j

=

N

∑i=1

( f ei ) or Ku = F (A.5)

where matrix[Ke

ij

]is the coefficient matrix, vector

ue

j

is the nodal value vector of

the unknown variables, and vector

f ei

is comprised of contributions from body forceterms and initial/boundary conditions.

For elasticity problems, the matrix[Ke

ij

]is called the element stiffness matrix given by[

Keij

]=

ˆΩi

([Bi] [Ni])T ([Di]

[Bj])

dΩ (A.6)

where matrix [Di] is the elasticity matrix and matrix [Bi] is the geometry matrix de-termined by the relation between the displacement and strain. The global stiffness matrixK is banded and symmetric because the matrices [Di] are symmetric. Material inhomo-geneity in FEM is most straightforwardly incorporated by assigning different materialproperties to different elements (or regions). To enforce the displacement compatibilitycondition, the order of shape functions along a common edge shared by two elementsmust be the same, so that no displacement discontinuity occurs along and across theedge.

“Infinite elements” have also been developed in FEM to consider the effects of aninfinite far-field domain on the near-field behavior, most notably the “infinite domainelements” of Beer and Meek (1981) and the “mapped infinite elements” of Zienkiewiczet al. (1983), with focus on geomechanical applications. The mapped infinite elementsare simply implemented using special shape functions that project boundary nodes atinfinite distances in one or two directions, where the displacements are either zero orhave prescribed values. Additional nodes are needed at the imaginary infinite locations.The infinite domain element technique does not require additional infinite nodes, butrequires a “decay function” to describe the manner in which the displacements vary frommesh boundary to infinity. The shape functions used in the infinite element formulationsare singular at the “infinite” nodes.

A.3.3 BOUNDARY ELEMENT METHOD (BEM)

The BEM requires discretization at the boundary of the solution domains only, thus redu-cing the problem dimensions by one and greatly simplifying the input requirements. The

A.3 Continuum methods 259

information required in the solution domain is separately calculated from the informa-tion on the boundary, which is obtained by solution of a boundary integral equation,instead of direct solution of the PDEs, as in the FDM and FEM.

Unlike the FEM and FDM methods, the BEM approach initially seeks a weak solutionat the global level through an integral statement, based on Betti’s reciprocal theorem andSomigliana’s identity. For a linear elasticity problem with domain Ω boundary Γ of unitoutward normal vector ni and constant body force fi for example, the integral statementis written as

cijuj +

ˆΓ

t∗ijujdΓ =

ˆΓ

u∗ijtjdΓ +

ˆΓ

∂u∗ij∂n

f jdΓ (A.7)

where uj and tj are the displacement and traction vectors on the boundary Γ the terms u∗ijand t∗ij are called displacement and traction kernels. The term cij is called the free termdetermined by the local geometry of the boundary surfaces, cij = 1 when the field pointis inside the domain Ω. The solution of the integral Equation A.7 requires the followingsteps:

1. Discretization of the boundary Γ with a finite number of boundary elements. For 2D prob-lems, the elements are 1D line segments which may have one node at the center ofthe element (constant element), two nodes at the two ends of the line segment (lin-ear elements) or three nodes with two end nodes and one central node (quadraticelements). Let N denote the total number of boundary elements. The boundaryintegral equation then is re-arranged into a sum of local integrals over all elements

cijuj +N

∑k=1

ˆΓk

t∗ijujdΓ =N

∑k=1

ˆΓk

u∗ijtjdΓ +N

∑k=1

ˆΓk

∂u∗ij∂n

f jdΓ (A.8)

2. Approximation of the solution of functions locally at boundary elements by (trial) shapefunctions, in a similar way to that used for FEM. The difference is that only 1D shapefunctions with intrinsic coordinate −1 6 ξ 6 1 is needed for 2D BEM problems,and 2D shape functions with two intrinsic coordinates−1 6 ξ 6 1 and−1 6 η 6 1are needed for 3D problems. The displacement and traction functions within eachelement are then expressed as the sum of their nodal values of the element nodes:

ui =m

∑k=1

Nkuki , ui =

m

∑k=1

Nktki (A.9)

where m is the element order (m = 1, 2 or 3 for 2D problems, for example), anduk

i and tki are the nodal displacement and traction values at node k, respectively.

260 General formulation of conventional numerical methods

Substitution of Equations A.9 into Equation A.8 and for

Tij =´

Γkt∗ijNjdΓ, Uij =

´Γk

u∗ijNjdΓ,

Bij =´

Γkf j

∂u∗ij∂n dΓ

(A.10)

Equation A.7 can be written in matrix form as[Tij (l, k)

] uj (k)

=[Uij (l, k)

] tj (k)

+ Bi (k)

(2N × 2N) (2N × 1) (2N × 2N) (2N × 1) (2N × 1)(A.11)

where i, j = 1, 2 for 2D and i, j = 1, 2,3 for 3D problems, respectively, l, k =1, 2, ..., N and

Tij (l, k) = cijδlk +

ˆΓk

t∗ijNjdΓ (A.12)

3. Evaluation of the integrals Tij, Uij and Bi with point collocation method by setting thesource point P at all boundary nodes successively.

4. Incorporation of boundary conditions and solution. Incorporation of the boundary con-ditions into the matrix (Equation A.11) will lead to final matrix equation

[A] x = b (A.13)

where the global matrix [A] is a mixture of Tij, and Uij, the unknown vector (x) is acomposite of both unknown displacements and unknown boundary tractions, andthe known vector b is the sum of the body force vector Bi and the products ofTij with known displacements and Uij with known tractions, respectively. The res-ultant Equation A.13 is usually fully populated and asymmetric, leading to fewerchoices for efficient equation solvers, compared with the sparse and symmetricmatrices encountered in the FEM. The solution of Equation A.13 will yield the val-ues of unknown displacements and tractions at boundary nodes. Therefore allboundary values of displacements and tractions are obtained.

5. Evaluation of displacements and stresses inside the domain. For practical problems, it isoften the stresses and displacements at some points inside the domain of interestthat have special significance. Unlike the FEM in which the desired data are auto-matically produced at all interior and boundary nodes, whether some of them areneeded or not, in BEM the displacement and stress values at any interior point, Pmust be evaluated separately by

ui (P) = −M

∑k=1

Tijukj +

M

∑k=1

Uij tkj +

M

∑k=1

Bk (A.14)

A.3 Continuum methods 261

σij (P) = −M

∑l=1

Skijulk +

M

∑l=1

Dkij tlk (A.15)

where kernels Tij, Uij, Skij, Dkij and Bi must be reevaluated according to the newposition of the source point inside the domain (closed-form formula for them areavailable in many text books on the BEM), usually without singularities unless thepoint is very close to boundary, and uk

j and tkj are known or calculated displacement

and traction vectors at all boundary nodes.

The main advantage of the BEM is the reduction of the computational model dimen-sion by one, with much simpler mesh generation and therefore input data preparation,compared with full domain discretization methods such as the FEM and FDM. Usingthe same level of discretization, the BEM is often more accurate than the FEM and FDM,due to its direct integral formulation. In addition, solutions inside the domain are con-tinuous, unlike the point wise discontinuous solutions obtained by the FEM and FDMgroups. The solution domains of BEM can be divided into several sub-domains withdifferent material properties, and this will often reduce the calculation time as well. Themethod is also suitable for considering infinite domains (full or half space/plane), dueto its use of the fundamental solutions.

However, in general, the BEM is not as efficient as the FEM in dealing with mater-ial heterogeneity, because it cannot have as many sub-domains as elements in the FEM.The BEM is also not as efficient as the FEM in simulating non-linear material behavior,such as plasticity and damage evolution processes, whereas is more suitable for solvingproblems of fracturing in homogeneous and linearly elastic bodies. The BEM formu-lation described above is called the direct formulation in which the displacements andtractions in the equations have clear physical meanings, are the basic unknowns of theboundary integral equations which are explicitly described on the problem boundary,and can be directly obtained by the solution of the integral equations. In the indirectformulation, on the other hand, the basic unknowns have no physical meanings and arejust fictitious source densities related to the physical variables such as displacements andtractions. The typical indirect BEMs are the Displacement Discontinuity Method (DDM)by Crouch (1976) for 2D problems and Weaver (1977) for 3D problems. The basic conceptof the indirect approach is to place the finite domain of interest into an imaginary infin-itely large domain (full or half plane or spaces) to derive the boundary integral equationsrelating the physical variables, such as displacements and tractions, to fictitious sourcedensities, such as fictitious load (stress) or displacement discontinuity.

262 General formulation of conventional numerical methods

A.4 Discontinuum methods

A.4.1 DISCRETE ELEMENT METHOD (DEM)

Rock mechanics is one of the disciplines from which the DEM originated (Cundall, 1971).The method has a broad variety of applications in rock mechanics, soil mechanics, struc-tural analysis, granular materials, material processing, fluid mechanics, multi-body sys-tems, etc. It is one of most rapidly developing areas of computational mechanics.

The DEM for modeling a discontinuum is relatively new compared with FDM, FEMand BEM and focuses mostly on applications in the fields of fractured or particulategeological media. The essence of the DEM is to represent the fractured medium as as-semblages of blocks formed by connected fractures in the problem domain, and solvethe equations of motion of these blocks through continuous detection and treatment ofcontacts between the blocks. The blocks can be rigid or be deformable with FDM or FEMdiscretizations. Large displacements caused by rigid body motion of individual blocks,including block rotation, fracture opening and complete detachments is straightforwardin the DEM, but impossible in the FDM, FEM or BEM.

The theoretical foundation of the method is the formulation and solution of equationsof motion of rigid and/or deformable bodies using implicit (based on FEM discretiza-tion) and explicit (using FVM discretization) formulations.

The key concept of DEM is that the domain of interest is treated as an assemblageof rigid or deformable blocks (particles, bodies) and the contacts among them need tobe identified and continuously updated during the entire deformation/motion process,and represented by proper constitutive models.

To formulate a DEM method to simulate the mechanical processes in rock mechanicsapplications, the following problems must be solved:

1. space sub-division and identification of block system topology;

2. representation of block deformation (rigid or deformable, using FVM or FEM);

3. developing an algorithm for contact detection (penalty function, Lagrange multi-plier, or augmented Lagrange multiplier);

4. obtaining constitutive equations for the rock blocks and fractures;

5. integration of the equations of motion of the blocks/particles (dynamic relaxation;time-marching FVM).

For rigid block analysis, an explicit time-marching scheme is used to solve the dynamicequations of motion of the rigid block system, based on a dynamic or static relaxation

A.4 Discontinuum methods 263

scheme, or an FDM approach in the time domain. For deformable block systems, thesolution strategies are different for the treatment of block deformability. One is explicitsolution with finite volume discretization of the block interiors, without the need forsolving large-scale matrix equations. The other is an implicit solution with finite elementdiscretization of the block interiors, which leads to a matrix equation representing thedeformability of the block systems, similar to that of the FEM.

The most representative explicit DEM methods is the Distinct Element Method cre-ated by (Cundall, 1980,1988) with the computer codes UDEC and 3DEC for 2D and 3Dproblems of rock mechanics (ITASCA Consulting Group, 1993c).

The implicit DEM was represented mainly by the Discontinuous Deformation Ana-lysis (DDA) approach, originated by Shi (1988). The method uses standard FEM meshesover blocks and the contacts are treated using the penalty method. In terms of devel-opment and application, the DDA approach occupies the front position. DDA has twoadvantages over the explicit DEM: permission for relatively larger time steps and closed-form integrations for the stiffness matrices of elements. An existing FEM code can alsobe readily transformed into a DDA code while keeping all the advantageous features ofthe FEM.

Another similar development, called the combined finite-DEM (Munjiza et al., 1995;Munjiza and Andrews, 2000), considers not only the block deformation but also fractur-ing and fragmentation of the rocks.

A.4.2 Explicit DEM - Distinct Element Method

The Distinct Element Method was originated in the early 70s by a landmark paper on theprogressive movements of rock masses as 2D rigid block assemblages (Cundall, 1971).The technique of the explicit DEM is presented comprehensively in Cundall and Hart(1992).

A.4.2.1 Block discretization

Blocks are represented as convex polyhedra in 3D with each face a planar convex poly-gon having a finite number of rectilinear edges. Their 2D counterparts are general poly-gons with a finite number of straight edges (Figure A.2). The 2D polygons can be eitherconvex or concave, but the 3D polyhedral must be convex. These blocks are formed byfractures which are represented in the problem domain either individually (for larger-scale fractures) or by a fracture sets generator (for smaller-scale fracture sets) using ran-dom distributions (based on site or modeling requirement data) of dip angles, dip dir-ections, spacing and apertures of the sets. The vertices (corners), edges and faces of

264 General formulation of conventional numerical methods

and apertures of the sets. The vertices (corners), edgesand faces of individual blocks and their connectionrelations are identified during the block generationprocess.The deformable blocks are further divided into a finite

number of constant strain triangles in 2-D or tetrahedrain 3-D. These triangles or tetrahedra form a mesh of theFVM (zones). Rectangular element meshes can also beused for 2-D problems when the problem geometry isfavourable.

3.4.1.2. Representation of deformation. An explicit,large strain Lagrangian formulation for the constantstrain elements is used to represent the elementdeformations. The displacement field of each ele-ment varies linearly and the faces or edges of theelements remain as planar surface or straight linesegments. Higher order elements may also be used, butcurved boundary surfaces (or edges) may be obtained,which may in turn complicate the contact-detectionalgorithm.Based on Gauss’ theorem to convert volume (area)

integrals into surface (line) integrals, the increments ofelement strain can be written

DeijEDt

2

XN

k¼1

½ðvmi Þnj7ðvm

j ÞniDSk; ð23Þ

where DSk is the area (or length) of the kth boundaryface (or edge) with unit normal nk

i ; and vmi is the mean

value of velocity over DSk: The summation extends overthe N faces (or edges) of an element (zone). The sign‘‘+’’ is used if i ¼ j; otherwise, the sign ‘‘’’ is used. Dt

is the time step. The stress increments are obtained byinvoking the constitutive equations for the blockmaterials.

3.4.1.3. Representation of contacts. Kinematically,block contacts are determined by the smallest distancebetween two blocks, pre-set in the codes or models.When this distance is within a prescribed threshold, apotential contact between these two blocks is numeri-cally established. The contact-detection algorithm in theDistinct Element Method programs determines thecontact type (different touching patterns betweenvertices, edges and faces), the maximum gap (if twoblocks do not touch but are separated by a gap close tothe pre-set tolerance), and the unit normal vectordefining the tangential plane on which sliding can takeplace. Table 1 lists all types of contacts.Mechanically, the interaction between two contacting

blocks is characterized by a stiffness (spring) in thenormal direction and a stiffness and friction angle(spring-slip surface series) in the tangential directionswith respect to the fracture surface (contact plane, seeFig. 16a). Interaction forces developed at contact pointsare determined as linear or non-linear functions ofthe deformations of springs and slip surfaces (i.e., therelative movements of blocks at contact points) andresolved into normal and tangential components,depending the constitutive models of the contacts (pointcontacts or edge/face contacts).The concept of contact ‘overlap’, though physically

inadmissible in block kinematics—because blocksshould not interpenetrate each other—may be acceptedas a mathematical means to represent the deformabilityof the contacts. However, it does present a numericalshortcoming that is difficult to overcome when thenormal forces or stresses at contact points are large. Inthis case, even with high normal stiffness, the ‘overlap’may be too excessive to be acceptable and thecalculation has to be stopped to implement someremedial measure (for example, to increase the normal

triangle element

tetrahedral elementa tetrahedral element

(a) (b) (c)

Fig. 15. Discretization of blocks by: (a) constant strain triangles; (b) constant strain tetrahedral; and (c) a typical tetrahedral element.

Table 1

Types of contacts for polygons and polyhedral

Block shapes Contact types

Arbitrary polygons (convex or concave) (2-D block) vertex-to-vertex, vertex-to-edge, edge-to-edge

Convex polyhedral (3-D block) vertex-to-vertex, vertex-to-edge, vertex-to-face, edge-to-edge, edge-to-face, face-to-face

L. Jing / International Journal of Rock Mechanics & Mining Sciences 40 (2003) 283–353 309

Figure A.2: Discretization of blocks by: (a) constant strain triangles; (b) constant straintetrahedral; and (c) a typical tetrahedral element (Jing, 2003).

individual blocks and their connection relations are identified during the block genera-tion process. The deformable blocks are further divided into a finite number of constantstrain triangles in 2D or tetrahedra in 3D. These triangles or tetrahedra form a mesh ofthe FVM (zones). Rectangular element meshes can also be used for 2D problems whenthe problem geometry is favorable.

A.4.2.2 Representation of deformation

An explicit, large strain Lagrangian formulation for the constant strain elements is usedto represent the element deformations. The displacement field of each element varieslinearly and the faces or edges of the elements remain as planar surface or straight linesegments. Higher order elements may also be used, but curved boundary surfaces (oredges) may be obtained, which may in turn complicate the contact-detection algorithm.Based on Gauss’ theorem to convert volume (area) integrals into surface (line) integrals,the increments of element strain can be written

∆εij ≈∆t2

N

∑k=1

[(vm

i ) nj ±(

vmj

)ni

]∆Sk (A.16)

where ∆Sk is the area (or length) of the kth boundary face (or edge) with unit normal nki ,

and vmi is the mean value of velocity over ∆Sk. The summation extends over the N faces

(or edges) of an element (zone). The sign “+” is used if i + j; otherwise, the sign “-” isused. ∆t is the time step. The stress increments are obtained by invoking the constitutiveequations for the block materials.

A.4 Discontinuum methods 265

A.4.2.3 Representation of contacts

Kinematically, block contacts are determined by the smallest distance between two blocks,pre-set in the codes or models. When this distance is within a prescribed threshold,a potential contact between these two blocks is numerically established. The contact-detection algorithm in the Distinct Element Method programs determines the contacttype (different touching patterns between vertices, edges and faces), the maximum gap(if two blocks do not touch but are separated by a gap close to the pre-set tolerance), andthe unit normal vector defining the tangential plane on which sliding can take place.

Table A.1 lists all types of contacts. Mechanically, the interaction between two con-tacting blocks is characterized by a stiffness (spring) in the normal direction and a stiff-ness and friction angle (spring-slip surface series) in the tangential directions with re-spect to the fracture surface (contact plane, see Figure A.3a).

Block shapes Contact types

Arbitrary polygons (convex orconcave) (2-D block)

vertex-to-vertex, vertex-to-edge,edge-to-edge

Convex polyhedral (3D block)vertex-to-vertex, vertex-to-edge,vertex-to-face, edge-to-edge, edge-to-face,face-to-face

Table A.1: Types of contacts for polygons and polyhedral.

Interaction forces developed at contact points are determined as linear or non-linearfunctions of the deformations of springs and slip surfaces (i.e., the relative movementsof blocks at contact points) and resolved into normal and tangential components, de-pending the constitutive models of the contacts (point contacts or edge/face contacts).The concept of contact “overlap”, though physically inadmissible in block kinematics(because blocks should not interpenetrate each other) may be accepted as a mathemat-ical means to represent the deformability of the contacts. However, it does present anumerical shortcoming that is difficult to overcome when the normal forces or stressesat contact points are large. In this case, even with high normal stiffness, the “overlap”may be too excessive to be acceptable and the calculation has to be stopped to implementsome remedial measure (for example, to increase the normal stiffness) and start again. Italso presents a problem for fluid flow calculation in which the apertures of fractures maybecome negative if “overlap” occurs at contact points. The mathematical representationof the contact “overlap” is thus not fully compatible with physical reality.

266 General formulation of conventional numerical methods

stiffness) and start again. It also presents a problem forfluid flow calculation in which the apertures of fracturesmay become negative if ‘overlap’ occurs at contactpoints. The mathematical representation of the contact‘overlap’ is thus not fully compatible with physicalreality.

3.4.1.4. Numerical integration of the equations of

motion. An explicit central difference scheme is appliedin the Distinct Element Method to integrate theequations of motion of the block system, as opposedto the implicit approach utilized in other continuum-based numerical methods. The unknown variables(contact forces or stresses) on the block boundary orin the internal elements are determined locally at eachtime step from the known variables on the boundaries,in the elements and their immediate neighbours. There isno need to set up and solve a matrix form of theequations of motion. The non-linearity in the materialbehaviour (of the fractures or intact blocks) can behandled in a straightforward manner.The equations of motion for a rigid block, in terms of

translational and rotational velocities, are written

vðtþDt=2Þi ¼ v

ðtDt=2Þi þ

Pfi

mþ bi

Dt;

oðtþDt=2Þi ¼ oðtDt=2Þ

i þP

Mi

IDt; ð24Þ

where m is the block mass, I is the moment of inertia, bi

are the volume force components of the block and Mi

are the components of the resultant moment. Thedisplacement at the next time step is then given by

uðtþDtÞi ¼ u

ðtÞi þ v

ðtþDt=2Þi Dt;

yðtþDtÞi ¼ yðtÞi þ oðtþDt=2Þ

i Dt; ð25Þ

where yi is the angular displacement of the block.

For deformable blocks, the equations of motion arewritten for grid points—the vertices of internal differ-ence elements. The central difference scheme is similar tothe first equation in Eq. (24) with only one modificationto the resultant out-of-balance force, fi

fi ¼ f ci þXN

k¼1

sijðnkj DSkÞ; ð26Þ

where f ci is the resultant contact force if the grid point ison the boundary of the block. The symbol N denotes thenumber of difference elements connected by this gridpoint.At each time step, the kinematic quantities (velocities,

displacements and accelerations) are first calculated andthe contact forces or stresses, as well as the internalstresses of the elements, are then obtained via constitu-tive relations for contacts.In the general calculation procedure, two basic tasks

are performed in turn. The kinematic quantities areupdated first, followed by invoking the constitutiverelations to provide the corresponding forces andstresses, see Fig. 17.

3.4.1.5. Applications and remarks. Due mainly to itsconceptual attractions in the explicit representation offractures, the DEM, especially the Distinct ElementMethod, has been enjoying wide application in rockengineering. A large quantity of associated publicationshas been published, especially in conference proceed-ings: it is not practical to list these even at a moderatelevel for this review. Therefore, a few representativereferences, mainly in international journals, are givenhere to show the wide range of the applicability of themethods:

* Tunnelling, underground excavations and mining:Barton (1991), Jing and Stephansson (1991), Nor-dlund et al. (1995), Chryssanthakis et al. (1997),Hanssen et al. (1993), Kochen and Andrade(1997), McNearny and Abel (1993), Souley et al.(1997a, b), Sofianos and Kapenis (1998) and Loriget al. (1995) [252–262];

* Rock dynamics: Zhao et al. (1999) and Cai and Zhao(2000) [263,264];

* Nuclear waste repository design and performanceassessment: Chan et al. (1995), Hansson et al. (1995),Jing et al. (1995, 1997) [265–268];

* Reservoir simulations: Gutierrez and Makurat (1997)[269];

* Fluid injection: Harper and Last (1989, 1990a, b)[270–272];

* Rock slopes, caving and gravity flow of particlesystems: Zhu et al. (1999) [273];

* Laboratory test simulations and constitutive modeldevelopment for hard rocks: Jing et al. (1993, 1994),Lanaro et al. (1997) [274–276];

K

K

n

t

(a)

l l

(b)

∆Ut∆Un

Old position

New position

F

F

t

n

llll 1 2 3 4

(c) (d)

C 1

C 2

C 3 C4

Φ

Fig. 16. Mechanical representation of contacts in the 2-D DEM.

L. Jing / International Journal of Rock Mechanics & Mining Sciences 40 (2003) 283–353310

Figure A.3: Mechanical representation of contacts in the 2D DEM (Jing, 2003).

A.4.2.4 Numerical integration of equations of motion

An explicit central difference scheme is applied in the Distinct Element Method to integ-rate the equations of motion of the block system, as opposed to the implicit approachutilized in other continuum based numerical methods. The unknown variables (contactforces or stresses) on the block boundary or in the internal elements are determined loc-ally at each time step from the known variables on the boundaries, in the elements andtheir immediate neighbors. There is no need to set up and solve a matrix form of theequations of motion. The non-linearity in the material behavior (of the fractures or intactblocks) can be handled in a straightforward manner.

The equations of motion for a rigid block, in terms of translational and rotationalvelocities, are written

v(t+∆t/2)i = v(t−∆t/2)

i +

[∑ fim

+ bi

]∆t

ω(t+∆t/2)i = ω

(t−∆t/2)i +

∑ MiI

∆t (A.17)

where m is the block mass, I is the moment of inertia, bi are the volume force componentsof the block and Mi are the components of the resultant moment. The displacement at

A.4 Discontinuum methods 267

the next time step is then given by

u(t+∆t)i = u(t)

i + v(t+∆t/2)i ∆t

θ(t+∆t)i = θ

(t)i + ω

(t+∆t/2)i ∆t (A.18)

where θi is the angular displacement of the block.For deformable blocks, the equations of motion are written for grid points - the ver-

tices of internal difference elements. The central difference scheme is similar to the firstequation in Equation A.17 with only one modification to the resultant out-of-balanceforce, fi

fi = f Ci +

N

∑k=1

σij

(nk

j ∆Sk)

(A.19)

where f ic is the resultant contact force if the grid point is on the boundary of the block.

The symbol N denotes the number of difference elements connected by this grid point.At each time step, the kinematic quantities (velocities, displacements and accelerations)are first calculated and the contact forces or stresses, as well as the internal stresses ofthe elements, are then obtained via constitutive relations for contacts. In the general cal-culation procedure, two basic tasks are performed in turn. The kinematic quantities areupdated first, followed by invoking the constitutive relations to provide the correspond-ing forces and stresses.

A.4.2.5 Applications

Due to the explicit representation of fractures, the DEM, especially the Distinct ElementMethod, hase been widely applicated in rock engineering. The applications of DEM con-centrate on hard rock problems and have increasing focus on coupled hydro-mechanicalbehavior - because of the dominating effects of the rock fractures on these aspects, andso where the explicit representation of fractures is necessary (Sharma et al., 2001). Forthe softer and weaker rocks, equivalent continuum models are more applicable becausethere is less difference between the deformability of the fractures and the rock matrix.

Despite the advantages of DEM, lack of knowledge of the geometry of the rock frac-tures limits its more general applications. In general, the geometry of fracture systemsin rock masses cannot be known and can only be roughly estimated. The adequacy ofthe DEM results in capturing the rock reality are therefore highly dependent on the in-terpretation of the in situ fracture system geometry - which cannot be even moderatelyvalidated in practice. Of course, the same problem applies also to the continuum models,

268 General formulation of conventional numerical methods

such as the FEM or FDM, but the requirement for explicit fracture geometry representa-tion in the DEM highlights the limitation and makes it more acute. Monte Carlo fracturesimulation may help to reduce the level of uncertainty, with increased computation cost.

A.4.3 Implicit DEM - Discontinuum Deformation Analysis method

DDA originated from a back analysis algorithm for determining a best fit to a deformedconfiguration of a block system from measured displacements and deformations (Shiand Goodman, 1985). It was later further developed to perform complete deformationanalysis of a block system (Shi, 1988).

By the second law of thermodynamics, a mechanical system under loading (externaland/or internal) must move or deform in a direction which produces the minimum totalenergy of the whole system. For a block system, the total energy consists of the potentialenergy due to different mechanisms like external loads, block deformation, system con-straints, kinetic and strain energy of the blocks and the dissipated irreversible energy.The minimization of the system energy will produce an equation of motion for the blocksystem, the same as that used in the FEM. For a system of N blocks, each having minodes (i = 1, 2, ..., N), the total number of nodes is m1 + m2 + ... + mN = M, and eachnode has two orthogonal displacement variables, u and v. Assuming, without losinggenerality, that nodes are numbered sequentially blockwise, the minimization will yield(2M× 2M) simultaneous equations, written symbolically as

k11 k12 k13 . . . k1Nk21 k22 k23 . . . k2Nk31 k32 k33 . . . k3N

......

......

...kN1 kN2 kN3 . . . kNN

d1d2d3...

dN

=

f1f2f3...

fN

or [K] D = F (A.20)

where diagonal sub-matrices kij is a (2mi × 2mi) matrix representing the sum of con-tributing sub-matrices for the ith block of mi nodes. Vector di is a (2mi × 1) vector ofdisplacement variables of the ith block and vector fi is a (2mi × 1) vector of resultantgeneral forces acting on the ith block. The off-diagonal sub-matrices kij (i 6= j) repres-ent the sum of contributing sub-matrices of contacts between blocks i and j and otherinter-block actions like bolting. The matrix [K] can also be called the global “stiffnessmatrix”.Compared with the explicit approach of the DEM, the DDA method has four basic ad-vantages over the explicit DEM:

A.4 Discontinuum methods 269

1. the equilibrium condition is automatically satisfied for quasi-static problems withoutusing excessive iteration cycles;

2. the length of the time step can be larger, and without inducing numerical instabil-ity;

3. closed-form integrations for the element and block stiffness matrices can be per-formed without the need for Gaussian quadrature techniques;

4. it is easy to convert an existing FEM code into a DDA code and include manymature FEM techniques without inheriting the limitations of the ordinary FEM,such as small deformation, continuous material geometry, and reduced efficiencyfor dynamic analysis. However, matrix equations are produced and need to besolved, using the same FEM technique.

The DDA method has emerged as an attractive model for geomechanical problems be-cause its advantages cannot be replaced by continuum-based methods or explicit DEMformulations.

A.4.4 DISCRETE FRACTURE NETWORK (DFN) method

The DFN method is a special discrete model that considers fluid flow and transportprocesses in fractured rock masses through a system of connected fractures. The DFNmodel is established on the understanding and representation of the two key factors:fracture system geometry and transmissivity of individual fractures.

The stochastic simulation of fracture systems is the geometric basis of the DFN ap-proach and plays a crucial role in the performance and reliability of the DFN model, inthe same way as the DEM. The key process is to create PDFs of fracture parameters relat-ing to the densities, orientations and sizes, based on field mapping results using boreholelogging data and scan line or window mapping techniques, and generate the realizationsof the fractures systems according to these PDFs and assumptions about fracture shape(circular discs, ellipses or polygons), (Dershowitz, 1984; Billaux et al., 1989).

Numerical techniques have been developed for the solution of flow fields for in-dividual fracture elements using closed-form solutions, the finite element model, theboundary element model, the pipe model and the channel lattice model.

Closed-form solutions exist, at present, only for planar, smooth fractures with parallelsurfaces of regular shape (i.e. circular or rectangular discs) for steady-state flow (Long,1983) or for both steady state and transient flow (Amadei and Illangasekare, 1992). Forfractures with general shapes, numerical solutions must be used. The FEM discretizationtechnique is perhaps the most well-known techniques used in the DFN flow models and

270 General formulation of conventional numerical methods

mapping results using borehole logging data and scan-line or window mapping techniques, and generate therealizations of the fractures systems according to thesePDFs and assumptions about fracture shape (circulardiscs, ellipses or polygons), (Dershowitz, 1984; Billauxet al., 1989) [387–388].A critical issue in this technique is the treatment of

bias in estimation of the fracture densities and tracelengths from conventional straight scanline or rectan-gular window mappings. A notable recent developmentusing circular windows (Mauldon, 1998; Mauldon et al.,2001) [389,390] provides an important step forward inthis regard.

3.5.3. Solution of the flow fields within fractures

Numerical techniques have been developed for thesolution of flow fields for individual fracture elementsusing closed-form solutions, the finite element model,the boundary element model, the pipe model and thechannel lattice model.Closed-form solutions exist, at present, only for

planar, smooth fractures with parallel surfaces ofregular shape (i.e. circular or rectangular discs) forsteady-state flow (Long, 1983) [391] or for both steady-state and transient flow (Amdei and Illangaseekare,1992) [392]. For fractures with general shapes, numericalsolutions must be used. The FEM discretizationtechnique is perhaps the most well-known techniquesused in the DFN flow models and has been used in theDFN codes FRACMAN/MAFIC and NAPSAC. Thebasic concept is to impose an FEM mesh overthe individual discs representing fractures in space(Fig. 20a) and solve the flow equations. The apertureor transmissivity field within the fracture can be eitherconstant or randomly distributed. Similarly, the BEMdiscretization can also be applied with the boundaryelements defined only on the disc boundaries (Fig. 20b),

with the fracture intersections treated as internalboundaries in the BEM solution. The compatibilitycondition is imposed at the intersections of discs. SeeElsworth (1986a, b) [372,373] and Robinson (1986) [393]for detailed formulations.The pipe model represents a fracture as a pipe of

equivalent hydraulic conductivity starting at the disccentre and ending at the intersections with otherfractures (Fig. 20c), based on the fracture transmissivity,size and shape distributions (Cacas, 1990) [394]. Thechannel lattice model represents the whole fracture by anetwork of regular pipe networks (Fig. 3.13d). The pipemodel leads to a simpler representation of the fracturesystem geometry, but may have difficulties to properlyrepresent systems of a number of large fractures.The channel lattice model is more suitable for

simulating the complex flow behaviour inside thefractures, such as the ‘‘channel flow’’ phenomena (Tsangand Tsang, 1987) [395], and is computationally lessdemanding than the FEM and BEM models since thesolutions of the flow fields through the pipe elements areanalytical.The fractal concept has also been applied to the DFN

approach in order to consider the scale dependence ofthe fracture system geometry and for upscaling thepermeability properties, using usually the full boxdimensions or the Cantor dust model (Barton andLarsen, 1985; Chil!es, 1988; Barton, 1992) [396–398].Power law relations have been also found to exist fortrace lengths of fractures and have been applied forrepresenting fracture system connectivity (Renshaw,1999) [399].

3.5.4. Issues of importance and difficulty

The influence of the rock matrix on flow in rockfractures is usually not considered in the DFN models.However, the related effects also need to be estimated

(a) (b)

(c)

Equivalent pipes Equivalent pipe networks

BE elementsIntersection

FE elementsIntersection

(d)

Fig. 20. Representation of rock fractures for the flow equation solution: (a) FEM; (b) BEM; (c) equivalent pipes; and (d) channel lattice model.

L. Jing / International Journal of Rock Mechanics & Mining Sciences 40 (2003) 283–353316

Figure A.4: Representation of rock fractures for the flow equation solution: (a) FEM; (b)BEM; (c) equivalent pipes; and (d) channel lattice model (Jing, 2003).

has been used in the DFN codes FRACMAN/MAFIC (Dershowitz et al., 1993) and NAP-SAC (Herbert, 1994). The basic concept is to impose a FEM mesh over the individualdiscs representing fractures in space (Figure A.4a) and solve the flow equations. Theaperture or transmissivity field within the fracture can be either constant or randomlydistributed. Similarly, the BEM discretization can also be applied with the boundaryelements defined only on the disc boundaries (Figure A.4b), with the fracture intersec-tions treated as internal boundaries in the BEM solution. The compatibility condition isimposed at the intersections of discs.

The pipe model represents a fracture as a pipe of equivalent hydraulic conductivitystarting at the disc center and ending at the intersections with other fractures (FigureA.4c), based on the fracture transmissivity, size and shape distributions.

The channel lattice model represents the whole fracture by a network of regular pipenetworks (Figure A.4d). The pipe model leads to a simpler representation of the fracturesystem geometry, but may have difficulties to properly represent systems of a numberof large fractures. The channel lattice model is more suitable for simulating the complexflow behavior inside the fracturesand is computationally less demanding than the FEMand BEM models since the solutions of the flow fields through the pipe elements areanalytical.

A.5 Hybrid models 271

A.5 Hybrid models

Hybrid models are frequently used in rock engineering, basically for flow and stress-deformation problems of fractured rocks. The main types of hybrid models are the hy-brid BEM/FEM and DEM/BEM models. The hybrid DEM/FEM models are also de-veloped. The BEM is most commonly used for simulating far-field rocks as an equi-valent elastic continuum, and the FEM and DEM for the non-linear or fractured near-field where explicit representation of fractures or non-linear mechanical behavior, suchas plasticity, is needed. This harmonizes the geometry of the required problem resolu-tion with the numerical techniques available, thus providing an effective representationof the effects of the far-field to the near-field rocks.

A.5.1 HYBRID FEM/BEM MODELS

The hybrid FEM/BEM was first proposed in Zienkiewicz et al. (1977) as a general stressanalysis technique. The standard technique is to treat the BEM region as a “super” ele-ment with an artificially “symmetrized” stiffness matrix, using the least-square tech-niques, so that it can be easily inserted into the symmetric FEM stiffness matrix for thefinal solution, which is easier to handle than the non-symmetric BEM stiffness matrix.However, such artificial “symmetrization” introduces additional errors into the final sys-tem equations. The coupling can also be performed in the opposite direction, i.e. treatthe FEM region as a “super” BEM element, and insert the corresponding FEM stiffnessmatrix into the final BEM stiffness matrix; this leads to a asymmetric stiffness matrix forthe final equation, which needs additional computational efforts for solution.

The hybrid BEM/FEM models are as efficient computationally as the FEM, with theadditional advantage of being able to deal with the non-linear behavior of materials inthe FEM region, using the FEM’s advantages. However, this advantage may be affectedby the symmetrized BEM equation. A possible step forward in this direction is to use theGalerkin double integration techniques in the BEM region so that the final BEM stiffnessmatrix is automatically symmetric, and therefore can be directly inserted in the finalhybrid BEM/FEM matrix without errors caused by artificial “symmetrization”.

A.5.2 HYBRID DEM/BEM MODELS

The hybrid DEM/BEM model was implemented only for the explicit Distinct ElementMethod, in the code group of UDEC and 3DEC. The technique was created by Lorig andBrady (1982), and was implemented into UDEC by Lemos (1987). The basic concept is totreat the BEM region (which surrounds the DEM region) as a “super” block having con-

272 General formulation of conventional numerical methods

inside the blocks, but it does produce displacements ofblocks, often of large scale. In the continuum approach,the rigid body motion mode of deformation is generallynot included because it does not produce strains in theelements. Therefore, a continuous system reflects mainlythe ‘‘material deformation’’ of the system and thediscrete system reflects mainly the ‘‘member (unit, orcomponent) movement’’ of the system.The choice of continuum or discrete methods depends

on many problem-specific factors, but mainly on theproblem scale and fracture system geometry. Fig. 4 illus-trates the alternative choices for different fracturecircumstances in rock mechanics problems. Continuumapproaches should be used for rock masses with no frac-tures or with many fractures, the behaviour of the latterbeing established through equivalent properties establishedby a homogenization process (Fig. 4a and d). Thecontinuum approach can be used if only a few fracturesare present and no fracture opening and no complete blockdetachment is possible (Fig. 4b). The discrete approach ismost suitable for moderately fractured rock masses wherethe number of fractures too large for continuum-with-fracture-elements approach, or where large-scale displace-ments of individual blocks are possible (Fig. 4c).Modelling fractured rocks demands high performance

numerical methods and computer codes, especiallyregarding fracture representations, material heterogeneityand non-linearity, coupling with fluid flow and heattransfer and scale effects. It is often unnecessarilyrestrictive to use only one method, even less one code,to provide adequate representations for the most sig-nificant features and processes: hybrid models or multipleprocess codes are often used in combination in practice.There are no absolute advantages of one method over

another, as is explained further in the later part of this

review. However, some of the disadvantages inherent inone type can be avoided by combined continuum-discrete models, termed hybrid models. In 1984, Lorigand Brady [3] presented an early computational schemein which the far-field rock is modelled as a transverselyisotropic continuum using the BEM and the near-fieldrock as a set of discrete element blocks defined by rockfractures. This type of hybrid BEM-DEM is shown inFig. 5. The complex rock mass behaviour caused byfractures and matrix non-linearity in the near-field of theexcavation can be efficiently handled by the DEM orFEM, surrounded by a BEM representation of the far-field region with linear material behaviour withoutfractures. The basis for such simple representation ofthe far-field is the fact that the gradients of variationof the physical variables, such as stress, displacement orflow, decrease rapidly with distance from the excavation.Therefore, if the interface between the near-field

(a) (b)

(c) (d)

Persistentdiscontinuities

Sets of discontinuities

continuum continuum

Fig. 4. Suitability of different numerical methods for an excavation in a rock mass: (a) continuum method; (b) either continuum with fracture

elements or discrete method; (c) discrete method; and (d) continuum method with equivalent properties.

Continuum for the far-field

Discontinuum for the near field

Boundary elements

excavation

on the interfaceBoundary elementson the outer boundary

Fig. 5. Hybrid model for a rock mass containing an excavation—using

the DEM for the near-field region close to the excavation and the BEM

for the far-field region.

L. Jing / International Journal of Rock Mechanics & Mining Sciences 40 (2003) 283–353290

Figure A.5: Hybrid model for a rock mass containing an excavation-using the DEM forthe near-field region close to the excavation and the BEM for the far-field region (Jing,2003).

tacts with smaller blocks along the interfaces with the DEM region (Figure A.5), whichcan be treated in standard DEM contact representations. The key conditions are:

1. the kinematic continuity along the interfaces of the two regions during the time-marching process;

2. the elastic properties of the two regions near the interface are similar.

Condition 2 indicates that blocks in the DEM regions must be deformable, i.e. not berigid blocks. In the case of mixed rigid and deformable block systems, special equationsof motion need to be developed to handle such cases.

A.5.3 Hybrid FEM/DEM models

Besides the above mainstream hybrid formulations, there are other coupling techniqueswhich take advantage of different numerical methods. Pan and Reed (1991) reported ahybrid DEM/FEM model, in which the DEM region consists of rigid blocks and the FEMregion can have non-linear material behavior. The algorithm places the FEM calculationsinto the DEM time-marching process. Since the blocks in DEM region are rigid andthe FEM region is an elastic continuum, the kinematical continuity condition along theinterface of the DEM and FEM regions may not be satisfied.

A.6 Numerical modeling of fracture process 273

The hybrid FEM/DEM model has been recently implemented in ELFEN, a commer-cial numerical code developed by Rockfield (Swansea, UK), since 1986. The developmenthas been undertaken collaboration with the Institute of Numerical Methods in Engineer-ing (INME) at University of Wales Swansea, which is internationally well-known forits pioneering work in this area. ELFEN is a dynamic solver based on the Finite Ele-ment Method (FEM) and specialized in problems of transition from a continuum to adiscontinuum material. It includes specific non linear fracture mechanics algorithms tosimulate the fracturing and produce discrete fractures (Klerck et al., 2004). If the fracturecriterion within the intact rock (represented by FEM) is met, then a crack (representedby DEM) is initiated. Adaptive re-meshing allows the fracture process through the FEMmesh to be tracked and visualized; thus contact properties can be assigned to pre-existingcracks and newly generated cracks.

A.6 Numerical modeling of fracture process

Modeling fractured rocks demands high performance numerical methods and computercodes, especially regarding fracture representations, material heterogeneity and non-linearity, coupling with fluid flow and heat transfer and scale effects.

It is often restrictive to use only one method to provide adequate representations forthe most significant features and processes: hybrid models or multiple process codesare often used in combination. There are no absolute advantages of one method overanother. However, some of the disadvantages inherent in one type can be avoided bycombined continuum-discrete models, termed hybrid models.

A.6.1 Fracture analysis with FDM/FVM

Explicit representation of fractures is not easy in FDM/FVM because the finite differenceschemes in FDM and interpolations in FVM require continuity of the functions betweenthe neighboring grid points. During the early development of FVM approaches, it is pos-sible to represent weakness zones of certain thickness as collections of cells of differentmaterials, which are not permitted to have openings or to be detached from their neigh-boring cells. The FDM/FVM deformable models have been used to study the mechan-isms of macroscopic fracturing processes, such as shear-band formation in the laboratorytesting of rock and soil samples and slope stability. This is achieved as a process of ma-terial failure or damage propagation at the grid points or cell centers, without creatingfracture surfaces in the models.

274 General formulation of conventional numerical methods

A.6.2 Fracture analyses with FEM

Representation of rock fractures in the FEM has been motivated by rock mechanics needssince the late 1960s, with the most notably contributions from Goodman et al. (1968),Zienkiewicz et al. (1970), Ghaboussi et al. (1973), Katona (1983), Desai et al. (1984).

Assuming that the contact stresses and relative displacements along and across therock fractures of a theoretical zero thickness (Figure A.6a), Goodman et al. (1968) pro-posed a “joint element” which can be readily incorporated into an FEM process.The well-known “Goodman joint element” in rock mechanics literature, has been widelyimplemented in FEM codes and applied to many practical rock engineering problems.Also, it has been extended to consider peak and post-peak behavior in the shear direc-tion. However, its formulation is based on continuum assumptions, so that large-scaleopening, sliding, and complete detachment of elements are not permitted. Because ofthe zero thickness of the joint element, numerical ill-conditioning may arise due to largeaspect ratios (the ratio of length to thickness) of joint elements.Zienkiewicz et al. (1970) proposed a six-node fracture element with two additional nodesin the middle section of the element, and a small thickness (Figure A.6c). The elementscan, therefore, be curved. The formulation may be seen as a “degenerate” ordinary solidelement of narrow thickness, and is subject to numerical ill-conditioning when the aspectratio is too large.

Using the relative displacements between the two opposite surfaces of fractures asthe independent system unknowns, Ghaboussi et al. (1973) proposed an FEM joint ele-ment based on the theory of plasticity (Figure A.6b). The use of the relative displacementcomponents across and along the fractures of finite thickness reduces the number of un-knowns of the fracture elements by half, defined at two nodes instead of four nodes as inGoodman’s joint elements. A finite thickness t is also used. The normal and shear straincomponents of the element are defined as the corresponding ratios of relative normaland shear displacements over the fracture thickness. An elasto-plastic relation betweenthe normal and shear stresses and the normal and shear strains of the fracture element isformulated and can be implemented in the usual manner for continuum FEM analysis.This formulation is more robust in terms of numerical ill-conditioning as compared withthose proposed in Goodman et al. (1968) and Zienkiewicz et al. (1970), due to the use ofthe relative displacements.

The “thin-layer” elements developed by Desai et al. (1984) are also based on a con-tinuum assumption; these are a solid element with a specially developed constitutivemodel for contact and frictional sliding.

The fracture element formulation in FEM has also been developed with interface ele-ment models in contact mechanics, using the FEM approach, instead of the continuumsolid element approximation as mentioned above. Katona (1983) developed an FEM in-

A.6 Numerical modeling of fracture process 275

‘‘mapped infinite elements’’ of Zienkiewicz et al. (1983)[42], with focus on geo-mechanical applications. Theoriginal concept was proposed by Bettess (1977) [43] forfluid mechanics problems. An infinite element formula-tion with body force terms was given recently by Cheng(1996) [44] with the emphasis also on geotechnicalproblems. The mapped infinite elements are simplyimplemented using special shape functions that projectboundary nodes at infinite distances in one or twodirections, where the displacements are either zero orhave prescribed values. Additional nodes are needed atthe imaginary infinite locations. The infinite domainelement technique does not require additional infinitenodes, but requires a ‘‘decay function’’ to describe themanner in which the displacements vary from meshboundary to infinity. The shape functions used in theinfinite element formulations are singular at the‘‘infinite’’ nodes.Because rock mechanics is one of the most stimulating

fields for development of numerical methods—withmany special challenges, such as fractures, propertyheterogeneity and anisotropy, material and geometricalnon-linearity, and scale and time effects—much FEMdevelopment work and application has been specificallyoriented towards rock mechanics problems, as illu-strated in the publications of Owen and Hinton (1980),Naylor et al. (1981), Pande et al. (1990), Wittke (1990),and Beer and Watson (1992) [45–49]. The FEM has beenthe most widely applied numerical methods for rockmechanics problems in civil engineering because it wasthe first numerical method with enough flexibility fortreatment of material heterogeneity, non-linear deform-ability (mainly plasticity), complex boundary condi-tions, in situ stresses and gravity. A typical recentdevelopment is given in Tang et al. (1998) [50] forsimulating fracturing processes in inhomogeneous rockswith FEM. Also, the method appeared in the late 1960sand early 1970s, when the traditional FDM with regulargrids could not satisfy these essential requirements forrock mechanics problems. It out-performed the conven-tional FDM because of these advantages.

3.2.2. Fracture analysis with the FEM

Representation of rock fractures in the FEM has beenmotivated by rock mechanics needs since the late 1960s,with the most notably contributions from Goodmanet al. (1968), Goodman (1976), Zienkiewicz et al. (1970),Ghaboussi et al. (1973), Katona (1983), Desai et al.(1984) [51–56].Assuming that the contact stresses and relative

displacements along and across the rock fractures of atheoretical zero thickness (Fig. 9a) follow a linearrelation with constant normal and shear stiffness, Kn

and Ks; Goodman et al. (1968) [51] proposed a ‘jointelement’ which can be readily incorporated into an FEMprocess, with its local equilibrium equation given by

kGuG ¼ fG; ð7Þ

where the matrix kG is a symmetric matrix with itsentries defined by the normal and shear stiffness, theelement’s length and its orientation to the globalcoordinate system, respectively. The vector uG ¼ðui

x; uiy; u

jx; u

jy; u

kx; u

ky ; u

lx; u

lyÞT is the nodal displacement

vector of the four nodes (i; j; k and l) defining the jointelement (Fig. 9b) and vector fG:The above formulation, the well-known ‘Goodman

joint element’ in rock mechanics literature, has beenwidely implemented in FEM codes and applied to manypractical rock engineering problems. Also, it has beenextended to consider peak and post-peak behaviour inthe shear direction. However, its formulation is based oncontinuum assumptions—so that large-scale opening,sliding, and complete detachment of elements are notpermitted. The displacements of a joint element are ofthe same order of magnitude as its neighbouringcontinuum elements, allowing the displacement compat-ibility condition to be kept along and across the jointelements. Because of the zero thickness of the jointelement, numerical ill-conditioning may arise due tolarge aspect ratios (the ratio of length to thickness) ofjoint elements.Zienkiewicz et al. (1970) [53] proposed a six-node

fracture element with two additional nodes in the

(a) (c)

(b) (d)

thickness =ts

nji

l

thickness = 0s

nl k

ji2/l 2/l

nm

l

kj

i

t

ξ

ηζp

o

n

m

l k

ji

Fig. 9. Fracture elements in FEM by (a) Goodman et al. (1968) [51], (b) Ghaboussi et al. (1973) [54], (c) Zienkiewicz et al. (1970) [53] and

(d) Buczkowski and Kleiber (1997) [60].

L. Jing / International Journal of Rock Mechanics & Mining Sciences 40 (2003) 283–353296

Figure A.6: Fracture elements in FEM by (a) Goodman et al. (1968), (b) Ghaboussi et al.(1973), (c) Zienkiewicz et al. (1970) and (d) Buczkowski and Kleiber (1997).

terface element model defined by mating pairs of nodes, without using the normal andshear stiffness parameters, and with three states - sticking, slipping and opening - basedon the Coulomb friction law.

Despite these efforts, the treatment of fractures and fracture growth remains the mostimportant limiting factor in the application of the FEM for rock mechanics problems, es-pecially when large number of fractures needs to be represented explicitly. The FEM suf-fers from the fact that the global stiffness matrix tends to be ill-conditioned when manyfracture elements are incorporated. Block rotations, complete detachment and large-scale fracture opening cannot be treated because the general continuum assumption inFEM formulations requires that fracture elements cannot be torn apart.

However, special algorithms have been developed in an attempt to overcome this dis-advantage, e.g. using discontinuous shape functions (Wan, 1990) for implicit simulationof fracture initiation and growth through bifurcation theory.

A special class of FEM, often called “enriched FEM”, has been especially developedfor fracture analysis with minimal or no re-meshing, as reported in Belytschko and Black(1999). The basic concept is direct representation of the objective function (such as dis-placements) with arbitrary discontinuities and discontinuous derivatives in FEM, butwithout need for the FEM meshes to conform to the fractures and no need for re-meshingfor fracture growth.

The treatment of fractures is at the element level. The surfaces of the fractures aredefined by assigned distance functions so that their representation requires only nodalfunction values, represented by an additional degree of freedom in the trial functions, a

276 General formulation of conventional numerical methods

discontinuity. Dunbar (1985) [154] showed the equiva-lence between the direct and indirect BEM approaches.

3.3.2. Fracture analysis with BEM

To apply standard direct BEM for fracture analysis,the fractures must be assumed to have two oppositesurfaces, except at the apex of the fracture tip wherespecial singular tip elements must be used. Denote Gc asthe path of the fractures in the domain O with its twoopposite surfaces represented by Gþ

c and Gc ; respec-

tively, Somigliana’s identity (when the field point is onthe boundary) can be written as

uj cijDuj þZG

tijuj dGþZGc

tijDuj dG

¼ZG

uij tj dGþ

ZGc

uijDtj dGþ

ZG

quij

qnfj dG; ð17Þ

where Dui and Dti are the displacement and tractionjumps across the two opposite surfaces of the fractures.Because of the very small thickness of fractures, the

two nodes at the opposite surfaces of a fracture will infact occupy the same coordinates. This will naturallylead to singular global stiffness matrices if the sameboundary conditions (or unknowns) are specified at thetwo opposite fracture surfaces. Also, any set of equaland opposite tractions on the fracture surfaces will leadto the same equation since Dti ¼ 0: In addition, thedisplacement difference Duj becomes an additionalunknown on Gc besides uj :To overcome these difficulties, two techniques were

proposed. One was to divide the problem domain intomultiple sub-domains with fractures along their inter-faces (Fig. 13a), by Blandford et al. (1981) [155]. Thisway, the stiffness matrices contributed by oppositesurfaces of the same fracture will belong to differentsub-domain stiffness matrices; thus, the singularity ofthe global matrix is avoided. This technique, however,requires the knowledge of fracturing paths (used fordeciding sub-regions) and growth rate (for decidingelement sizes), which is determined by the solution of theproblem itself, before the problem solution, and may not

be applicable for many practical problems withoutsymmetry in geometry and boundary conditions.The second technique is the Dual Boundary Element

Method (DBEM). The essence of this technique is toapply displacement boundary equations at one surfaceof a fracture element and traction boundary equationsat its opposite surface, although the two opposingsurfaces occupy practically the same space in the model.The general mixed mode fracture analysis can beperformed naturally in a single domain (Fig. 13b). Theterm DBEM was first presented in Portela (1992) andPortela et al. (1992, 1993) [156–158], and was extendedto 3-D crack growth problems by Mi and Aliabadi(1992, 1994) [159,160]. However, the original concept ofusing two independent boundary integral equations forfracture analysis, one displacement equation and an-other its normal derivative, was developed first byWatson (1979) [128]. Special crack tip elements, such asdeveloped in Yamada et al. (1979) and Aliabadi andRooke (1991) [161,162], are used at the fracture tips toaccount for the stress and displacement singularity.The DDM has been widely applied to simulate

fracturing processes in fracture mechanics in generaland in rock fracture propagation problems in particulardue to the advantage that the fractures can berepresented by single fracture elements without needfor separate representation of their two oppositesurfaces, as should be done in the direct BEM solutions.It was developed by Crouch and Starfield (1983) [131]with open fractures, and was extended to fractures withcontact and friction by Wen and Wang (1991) [163] andShen (1991) [164] for mechanical and rock engineeringanalyses, respectively. The fictitious unknowns are thedisplacement discontinuity Dui acting on the boundaryG of a finite body of domain O; inserted in an infinitelylarge half (or full) space. The displacement and tractionis given by

ui ¼ cijDuj þZG

uijDuj dGþ

ZGc

%uijDuj dG;

ti ¼ cijDuj þZG

tijDuj dGþZGc

%tijDuj dG ð18Þ

(a) (b) (c)

Ω

Γ

c

1 Ω

ΓcΓ

ΩΓ+

_

Γc

Γc Γ

Fig. 13. Illustrative meshes for fracture analysis with BEM: (a) sub-domain, direct BEM; (b) single domain, dual BEM; and (c) single domain DDM.

L. Jing / International Journal of Rock Mechanics & Mining Sciences 40 (2003) 283–353 303

Figure A.7: Illustrative meshes for fracture analysis with BEM: (a) sub-domain, directBEM; (b) single domain, dual BEM; and (c) single domain DDM (Jing, 2003).

jump function along the fracture and a crack tip function at the tips. In Belytschko et al.(2001), the enriched method was applied to a tunnel stability analysis with fracturessimulated as displacement discontinuities.

The “enriched” FEM with jump functions and crack tip functions has improved theFEM’s capacity in fracture analysis. Coupled with the FEM’s advantage in dealing withmaterial heterogeneity and non-linearity, this makes the “enriched” FEM suitable fornon-linear fracture analysis. One such example is the so-called “generalized finite ele-ment method” (GFEM), which was developed based on the partition of unity principle(Duarte et al., 2000). The mesh in GFEM is independent of the geometry of the domainof interest and therefore can be regular regardless of the object geometry. Fractures canbe simulated by their surrounding nodes “enriched” by jump functions and crack tipfunctions.

A.6.3 Fracture analyses with BEM

To apply standard direct BEM for fracture analysis, the fractures must be assumed tohave two opposite surfaces, except at the apex of the fracture tip where special singulartip elements must be used.

Because of the very small thickness of fractures, the two nodes at the opposite sur-faces of a fracture will in fact occupy the same coordinates. This will naturally lead tosingular global stiffness matrices if the same boundary conditions (or unknowns) arespecified at the two opposite fracture surfaces.

A.6 Numerical modeling of fracture process 277

To overcome these difficulties, two techniques were proposed. One was to divide theproblem domain into multiple sub-domains with fractures along their interfaces (FigureA.7a), by Blandford et al. (1981). This way, the stiffness matrices contributed by oppos-ite surfaces of the same fracture will belong to different sub-domain stiffness matrices;thus, the singularity of the global matrix is avoided. This technique, however, requiresthe knowledge of fracturing paths (used for deciding sub-regions) and growth rate (fordeciding element sizes), which is determined by the solution of the problem itself, beforethe problem solution, and may not be applicable for many practical problems withoutsymmetry in geometry and boundary conditions.

The second technique is the Dual Boundary Element Method (DBEM) (Portela, 1992).The essence of this technique is to apply displacement boundary equations at one surfaceof a fracture element and traction boundary equations at its opposite surface, althoughthe two opposing surfaces occupy practically the same space in the model. The generalmixed mode fracture analysis can be performed naturally in a single domain (FigureA.7b). The original concept of using two independent boundary integral equations forfracture analysis, one displacement equation and another its normal derivative, was de-veloped first by Watson (1979).

The DDM has been widely applied to simulate fracturing processes in fracture mech-anics in general and in rock fracture propagation problems in particular due to the ad-vantage that the fractures can be represented by single fracture elements without needfor separate representation of their two opposite surfaces, as should be done in the directBEM solutions. It was developed by Crouch and Starfield (1983) with open fractures, andwas extended to fractures with contact and friction by Shen (1991) for rock engineeringanalyses.

278

References

Amadei, B. and Illangasekare, T. (1992). “Analytical solutions for steady and transientflow in non-homogeneous and anisotropic rock joints.” International Journal of RockMechanics and Mining Science and Geomechanics Abstracts, 29(6): 561–572.

Beer, G. and Meek, J. L. (1981). “Infinite domain elements.” International Journal forNumerical Methods in Engineering, 17(1): 43–52.

Belytschko, T. and Black, T. (1999). “Elastic crack growth in finite elements with minimalremeshing.” International Journal for Numerical Methods in Engineering, 45(5): 601–620.

Belytschko, T., Moës, N., Usui, S., and Parimi, C. (2001). “Arbitrary discontinuities infinite elements.” International Journal for Numerical Methods in Engineering, 50(4): 993–1013.

Billaux, D., Chiles, J. P., Hestir, K., and Long, J. C. S. (1989). “Three-dimensional stat-istical modeling of a fractured rock mass-an example from the Fanay-Augéres Mine.”International Journal of Rock Mechanics and Mining Science and Geomechanics Abstracts,26(3-4): 281–299.

Blandford, G., Ingraffea, A., and Liggett, J. (1981). “Two-dimensional stress intensityfactor computations using the boundary element method.” International Journal forNumerical Methods in Engineering, 17(3): 387–404.

Brighi, B., Chipot, M., and Gut, E. (1998). “Finite differences on triangular grids.” Nu-merical Methods for Partial Differential Equations, 14(5): 567–579.

Buczkowski, R. and Kleiber, M. (1997). “Elasto-plastic interface model for 3D-frictionalorthotropic contact problems.” International Journal for Numerical Methods in Engineer-ing, 40(4): 599–619.

279

280 General formulation of conventional numerical methods

Crouch, S. and Starfield, A. (1983). Boundary element methods in solid mechanics. GeorgeAllen and Unwin.

Crouch, S. L. (1976). “Solution of plane elasticity problems by the displacement discon-tinuity method. I. Infinite body solution.” International Journal for Numerical Methods inEngineering, 10(2): 301–343.

Cundall, P. A. (1971). “A computer model for simulating progressive large scale move-ments in blocky rock systems.” In “Proceedings of International Symposium on RockFracture,” volume 2. ISRM, Nancy.

Cundall, P. A. (1980). “UDEC - a generalized distinct element program for modellingjointed rock.” Technical Report Report PCAR-l-80, Peter Cundall Associates, EuropeanResearch Office, US Army Corps of Engineers.

Cundall, P. A. (1988). “Formulation of a three-dimensional distinct element model-PartI. A scheme to detect and represent contacts in a system composed of many polyhedralblocks.” Int. J. Rock Mech. Min. Sci. and Geomech., 25(3): 107–116.

Cundall, P. A. and Hart, R. D. (1992). “Numerical modelling of discontinua.” EngineeringComputations, 9(2): 101–113.

Dershowitz, W. S. (1984). Rock joint systems. Ph.D. thesis, Massachusetts Institute ofTechnology, Boston, USA.

Dershowitz, W. S., Lee, G., Geier, J., Hitchcock, S., and La Pointe, P. (1993). FracManuser documentation. Golder Associates Inc., Seattle WA.

Desai, C. S., Zaman, M. M., Lightner, J. G., and Siriwardane, H. J. (1984). “Thin-layerelement for interfaces and joints.” International Journal for Numerical and AnalyticalMethods in Geomechanics, 8(1): 19–43.

Duarte, C. A., Babuska, I., and Oden, J. T. (2000). “Generalized finite element methodsfor three-dimensional structural mechanics problems.” Computers and Structures, 77(2):215–232.

Ghaboussi, J., Wilson, E. L., and Isenberg, J. (1973). “Finite element for rock joints andinterfaces.” Journal of the Soil Mechanics and Foundations Division ASCE, 99(10): 849–862.

Goodman, R., Taylor, R., and Brekke, T. (1968). “A model for the mechanics of jointedrock.” J Soil Mech Div ASCE, 94(SM3): 637–659.

References 281

Herbert, A. (1994). NAPSAC (Release 3.0) summary document. AEA Technology, Harwell,UK.

ITASCA Consulting Group, L. (1993b). FLAC manuals.

ITASCA Consulting Group, L. (1993c). UDEC manuals.

Jing, L. (2003). “A review of techniques, advances and outstanding issues in numer-ical modelling for rock mechanics and rock engineering.” International Journal of RockMechanics and Mining Sciences, 40(3): 283–353.

Katona, M. (1983). “A simple contact-friction interface element with applications to bur-ied culverts.” International Journal for Numerical and Analytical Methods in Geomechanics,7(3): 371–384.

Klerck, P. A., Sellers, E. J., and Owen, D. R. J. (2004). “Discrete fracture in quasi-brittlematerials under compressive and tensile stress states.” Computer Methods in AppliedMechanics and Engineering, 193(27-29): 3035–3056.

Lemos, J. V. (1987). A hybrid distinct element computational model for the half-plane. Master’sthesis, Department of Civil Engineering, University of Minnesota.

Long, J. C. S. (1983). Investigation of equivalent porous media permeability in networks ofdiscontinuous fractures. Ph.D. thesis, Lawrence Berkeley Laboratory, University of Cali-fornia, Berkeley, CA.

Lorig, L. J. and Brady, B. H. G. (1982). “A hybrid discrete element-boundary elementmethod of stress analysis.” In R. E. Goodman and F. Heuze, editors, “Proceedings ofthe 23rd US Symposium Rock Mechanics, Berkeley,” pages 628–636.

Munjiza, A. and Andrews, K. R. F. (2000). “Penalty function method for combined finite-discrete element systems comprising large number of separate bodies.” InternationalJournal for Numerical Methods in Engineering, 49(11): 1377–1396.

Munjiza, A., Owen, D. R. J., and Bicanic, N. (1995). “A combined finite-discrete elementmethod in transient dynamics of fracturing solids.” Engineering Computations, 12(2):145–174.

Pan, X. and Reed, M. (1991). “A coupled distinct element-finite element method forlarge deformation analysis of rock masses.” International Journal of Rock Mechanics andMining Science & Geomechanics Abstracts, 28(1): 93–99.

282 General formulation of conventional numerical methods

Perrone, N. and Kao, R. (1975). “A general finite difference method for arbitrarymeshes.” Computers and Structures, 5: 45–58.

Portela, A. (1992). Dual boundary element incremental analysis of crack growth. Ph.D. thesis,Wessex Institute of Technology, Portsmouth University, Southampton, UK.

Sharma, V., Saxena, K., and Woods, R. (2001). Distinct element modelling in geomechanics.Balkema, Rotterdam.

Shen, B. (1991). Mechanics of fractures and intervening bridges in hard rocks. Ph.D. thesis,Royal Institute of Technology, Stockholm, Sweden.

Shi, G. (1988). Discontinuous deformation analysis-a new numerical model for the statics,dynamics of block systems. Ph.D. thesis, University of California, Berkeley, USA.

Shi, G. H. and Goodman, R. E. (1985). “Two dimensional discontinuous deformationanalysis.” International Journal for Numerical and Analytical Methods in Geomechanics,9(6): 541–556.

Wan, R. (1990). The numerical modeling of shear bands in geological materials. Ph.D. thesis,University of Alberta, Edmonton, AB.

Watson, J. O. (1979). “Advanced implementation of the boundary element method fortwo- and three-dimensional elastostatics.” In P. K. Banerjee and R. Butterfield, editors,“Developments in boundary element methods, vol. 1,” pages 31–63. Applied SciencePublishers, London.

Weaver, J. (1977). “Three-dimensional crack analysis.” International Journal of Solids andStructures, 13(4): 321–330.

Wheel, M. A. (1996). “A geometrically versatile finite volume formulation for planeelastostatic stress analysis.” The Journal of Strain Analysis for Engineering Design, 31(2):111–116.

Zienkiewicz, O., Emson, C., and Bettess, P. (1983). “A novel boundary infinite element.”International Journal for Numerical Methods in Engineering, 19(3): 393–404.

Zienkiewicz, O., Kelly, D., and Bettess, P. (1977). “The coupling of the finite elementmethod and boundary solution procedures.” International Journal for Numerical Meth-ods in Engineering, 11(2): 355–375.

A.6 Numerical modeling of fracture process 283

Zienkiewicz, O. C., Best, B., Dullage, C., and Stagg, K. (1970). “Analysis of non-linearproblems in rock mechanics with particular reference to jointed rock systems.” In“Proceedings of the 2nd International Congress on Rock Mechanics, ISRM, Belgrade,”.

284 General formulation of conventional numerical methods

Appendix B

Rock anisotropy

B.1 Introduction

In this appendix the attention is posed on rock anisotropy. A review of the deformabilityand strength anisotropy is done, in order to introduce the numerical modelling of someof the most common laboratory tests carried out on anisotropic rocks.

B.2 The nature of rock masses

The nature of rock masses is complex and resulting from a long history of tectonic forcesand other natural environmental effects.

In the context of continuum mechanics, the material and the forces applied to it haveto be considered. In rock mechanics, there is the intact rock divided by discontinuities toform the rock structure which is subjected to the in situ stress and influenced by factorssuch as water flow and time. In all these subjects the geological history has played itspart, altering the rock and the applied forces.

In engineering terms, intact rock is defined as rock without significant fractures. Themost complete description of the behaviour of intact rock is the complete stress-straincurve obtained in a uniaxial compression test (Figure B.1). The uniaxial compression testis conducted by loading a sample of rock along its axis and recording the displacementproduced as the force is increased. From the stress-strain curve it is possible to definethree important parameters. The first one is the Young’s modulus, E, defined as the ratioof stress to strain, and the second one is the maximum stress that the specimen can sus-

285

286 Rock anisotropy

tain. Under the loading conditions shown in the diagram, the peak stress is the uniaxialcompressive strength, σvc. A third interesting feature is the steepness of the descendingportion of the curve, that is a measure of rock brittleness.18 Geological setting

Stress

Figure 2.8 Complete stress-strain curve illustrating the stiffness (or modulus, E), the strength, a^, and brittleness.

Ductile

Figure 2.9 Illustration of the difference between a brittle material and a ductile material.

in the form of continuing strain at the same stress level (a ductile material) and a drop in the stress level to zero at the same strain value (a brittle material). The brittleness is indicated by the steepness of the curve betw^een these tw o limits. In fact, the situation is more complicated than this because it is possible to have strain-hardening materials (a curve above the ductile line) and super-brittle materials (to the left of the brittle line). These cases will be discussed further in Chapter 6.

Possible variation in the three main factors is illustrated for a suite of rocks in Figs 2.10-2.13. The figure legends describe the features quali-tatively. As we have mentioned, the form of the complete stress-strain curve is dictated by the nature of the microstructure. For example, a high grain strength, fine grain basalt has a high stiffness, high strength and is very brittle. On the other hand, a limestone rock with a variation in the grain geometry has a medium stiffness, medium strength and a more gentle descending part of the curve caused by the gradual deterioration of the microstructure as it is progressively and increasingly damaged.

There will be variations on this theme for the variety of microstructures that exist and the influence that they have on the shape of the curve—in

Figure B.1: The complete stress-strain curve illustrating various mechanical parameters(Hudson and Harrison, 1997).

During the history of the rock mass, due to the action of geological and environmentalfactors, the rock does break and the result is a complex structure of fractures forming rockblocks. The overall geometrical configuration of the discontinuities in the rock mass istermed rock structure.

It is helpful to understand how the fractures form: there are three modes of fracture,illustrated in Figure B.2. These types of fracture lead to two fundamentally different typeof discontinuities:

• joints, discontinuity that has been simply opened;

• shear zone or faults, discontinuity which has been subjected to some lateral move-ments.

Such features exist in all rock masses and they will significantly affect the deformability,strength and failure of rock masses. Other key characteristics, such as the permeabilitycan be governed almost entirely by the rock structure configuration.

B.3 Definitions 287

Figure B.2: Type of fracturing in rock: mode I - tensile; mode II - in plane shear; mode III- out of plane shear (Hudson and Harrison, 1997).

In this way, the deformation, strength and failure properties of a rock structure aredetermined by the mechanical properties of the intact rock and the geometrical and themechanical properties of the natural fractures of the rock mass.

B.3 Definitions

When dealing with the mechanical behaviour of solids, a common assumption is thatthey are homogeneous, continuous and isotropic, but the mechanical properties of rockmass are, however, not at all isotropic. In fact, rock mass anisotropy and inhomogen-eity are two important aspects to take into account when rock engineering problems areconsidered. Such characteristics are a natural consequence of millions of years of mech-anical, chemical and thermal processes to which the rock mass has been subjected.

The term “anisotropy” means “having different properties in different directions”,whereas the word “inhomogeneity” means “having different properties at different loc-ations”. In rock mechanics, these properties may be of any type: deformability, strength,brittleness, permeability and discontinuity frequency.

In the context of rock engineering, the anisotropic and inhomogeneous behaviour isexplained using two acronyms, CHILE and DIANE (Harrison and Hudson, 2000):

• Continuous, Homogeneous, Isotropic and Linearly-Elastic (CHILE) material is onethat is most commonly assumed for the purpose of modelling; traditionally stress

288 Rock anisotropy

analysis techniques are formulated in terms of these four attributes, simply forconvenience for obtaining closed-form solutions;

• Discontinuous, Inhomogeneous, Anisotropic, Non-Elastic (DIANE) rock is the ma-terial with which the rock engineering has to deal.

The rock is discontinuous because of the bedding plane separations and any other frac-turing that may be present. It is inhomogeneous because of the existence of the differentrock type. It is anisotropic because of its sedimentary nature. It is not elastic becausethere is hysteresis and time dependency, related to bedding planes. Moreover, the rockis anisotropic because it is inhomogeneous, e.g. the deformation modulus for a suite ofstrata will be different parallel and perpendicular of the bedding planes. Also, the rockis anisotropic because it is discontinuous.

B.4 Deformability anisotropy

There are three types of isotropy commonly considered in rock mechanics: completeisotropy, transverse isotropy and orthotropy. As indicated by the sketches in FigureB.3, transverse isotropy might well represent relatively unfractured sedimentary rocks,whereas, orthotropy could well be a good representation for rocks containing three mu-tually perpendicular sets of discontinuity.

In order to define the compliance matrix, it is assumed that each component of thestrain tensor is a linear combination of all the components of the stress tensor. In the caseof the εxx component, the relation is expressed as

εxx = S11σxx + S12σyy + S13σzz + S14τxy + S15τyz + S16τzx

Because there are six independent components of the strain matrix, there will be sixequations of this type. If it is considered that the strain in the x-direction were only dueto the stress in the x-direction, the previous equation would reduce to

εxx = S11σxx

or

σxx = εxx/S11 = Eεxx

where E = 1/S11.

B.4 Deformability anisotropy 289

64 In situ stress

high shear stresses can be present and result from tectonic activity. In Fig. 4.17, we illustrate the subduction zone off the coast of Chile, the genesis of the Andes and hence the likely high horizontal normal stresses we would expect from such activity. In fact, both the largest surface and largest underground mines in the World are in Chile—and they both show strong signs of stress-related phenomena in terms of rock slope instability and rockbursts, respectively.

4.8.3 Rock anisotropy

In Section 4.6.2 we derived the expression v/(l - v) as the ratio between the horizontal and vertical stresses and noted that for v varying between 0 and 0.5 the corresponding /c-ratio varied from 0 to 1. We will be discussing anisotropy in much greater detail in Chapter 5, on strain, and in Chapter 10 on inhomogeneity and anisotropy. It is worth mentioning here, though, that there are three types of isotropy commonly considered in rock mechanics, namely: complete isotropy, transverse isotropy and orthotropy.

Our earlier calculation was for a rock with complete isotropy (having the same properties in all directions). It is possible to calculate the /c-value for a material which is transversely isotropic (having different properties in the vertical direction to the horizontal directions), and for the orthotropic case (having different properties in three perpendicular directions). The explanation, both mathematical and intuitive, for these types of isotropy and the associated elastic material constants is given in Chapter 5. The important point is that the fc-values for each case are as shown in Fig. 4.18 by the terms in parentheses.

As indicated by the sketches in the figure, transverse isotropy might well represent relatively unfractured sedimentary rocks, whereas, orthotropy

Transversely isotropic

(horizontal)

Orthotropic

^ = y = i r : v r ^ x = ay = ( Y^ j o, a, = (^--^-^ja,

( V + V V \

1-v V - ) " -

Figure 4.18 Relation between vertical and elastically induced horizontal stresses for the different types of isotropy. Figure B.3: Relation between vertical and elastically induced horizontal stresses for the

different type of isotropy (from Hudson and Harrison, 1997).

290 Rock anisotropy

The complete set of relations between the strain tensor and stress components writtenin matrix notation is:

|ε| = |S||σ|where

[ε] =

εxxεyyεzzεxyεyzεzx

and [σ] =

σxxσyyσzzσxyσyzσzx

and [S] =

S11 S12 S13 S14 S15 S16

S22 S23 S24 S25 S26S33 S34 S35 S36

S44 S45 S46S55 S56

S66

The [S] matrix shown above is known as the compliance matrix. In general the higher

the magnitude of a specific element in this matrix, the greater will be the contribution tothe strain. “Compliance” is a form of “flexibility”, and it is the inverse of “stiffness”.The compliance matrix is a 6 x 6 matrix containing 36 elements and it is symmetrical.In order to define the matrix it is necessary to define 21 independent elastic constants tocompletely characterize a material that follows the generalized Hooke’s law.

An orthotropic material is characterized by 9 independent elastic constants whichare three Young’s moduli and the three shear moduli and the three Poisson’s ratio, i.e.:

E1 E2 E3 G12 G23 G31 υ21 υ32 υ31

The compliance matrix could be reduced even further considering the case of trans-verse isotropy. This is represented by a rock mass with a laminated fabric or one set ofparallel discontinuities. In the case when the plane of isotropy is parallel to the planecontaining Cartesian axes 1 and 2, it is possible to say that:

E1 = E2 = E and E3 = E′

υ12 = υ21 = υ and υ13 = υ23 = υ’

G12 6= G23 and G23 = G31 = G′

The associated compliance matrix is then

1/E −υE

−υ′E 0 0 0

1/E −υ′E 0 0 0

1/E′ 0 0 02(1+υ)

E 0 01/G′ 0

sym. 1/G′

B.4 Deformability anisotropy 291

B Rock masses

In this chapter, we will concentrate on extending the ideas discussed in the previous chapter on discontinuities to provide a predictive model for the deformability and strength of rock masses. In Chapter 12, we will be discussing rock mass classification (which is a method of combining selected geometrical and mechanical parameters) to semi-quantitatively provide an overall characterization, mainly for assessing excavation support requirements.

8.1 Deformability Consider first, as an initial step in the overall development of a deforma-bility model, the deformation of a set of parallel discontinuities under the action of a normal stress, assuming linear elastic discontinuity stiffnesses. This circumstance is illustrated in Fig. 8.1. To calculate the overall modulus of deformation, the applied stress is divided by the total deformation. We will assume that the thickness of the discontinuities is negligible in com-parison to the overall length under consideration, L. Additionally, we will assume that the deformation is made up of two components: one due to

Intact rock (units of stress MPa)

'

^ -. 1

, Discontinuities (units of stress/ length) MPa/m

N discontinuities, frequency = X

Figure 8.1 The modulus of deformation of a rock mass containing a discontinuity set.

Figure B.4: The modulus of deformation of a rock mass containing a discontinuity set(Hudson and Harrison, 1997).

note that the term 2(1 + υ)/E has been substituted for 1/G12 because in the plane ofisotropy there is a relation between the shear modulus G and the Young’s modulus E andPoisson’s ratio ν. Thus the number of independent elastic constants for a transverselyisotropic material is 5:

E1 E′ G′υ υ′

The final reduction can be made to the compliance matrix of a completely isotropicmaterial, resulting in 2 elastic constants (also using the shear modulus relation):

E1 = E2 = E3 = E

υ12 = υ23 = υ31 = υ

G12 = G23 = G31 = G

Now the attention is posed on the rock mass in order to provide a model for thedeformability (Hudson and Harrison, 1997). Consider first, the deformation of a setof parallel discontinuities under the action of a normal stress, assuming linear elasticdiscontinuity stiffnesses. This circumstance is illustrated in figure B.4.

To calculate the overall modulus of deformation, the applied stress is divided by thetotal deformation. The deformation is made up of two components: one due to the de-formation of the intact rock; the other due to the deformability of the discontinuities. Inthis calculation, the assumption that the thickness of the discontinuities is negligible incomparison to the overall length, L, is taken. The contribution made by the intact rock to

292 Rock anisotropy

the deformation, δl , is σL/E (i.e. strain multiplied by length). The contribution made bya single discontinuity to the deformation, δD, is σ/ED (ED relates the stress to displace-ment directly). Assuming a discontinuity frequency of λ, there will be λL discontinuitiesin the rock mass and the total contribution made by these to the deformation will be δt

D,which is equal to σλL/ED. Hence, the total displacement, δT , is

δT =σLE

+λσLED

with the overall strain being given by

ε =δt

L=

σ

E+

λσ

ED

Finally, the overall modulus, EMASS, is given by

EMASS =σ

ε=

11E + λ

ED

GMASS =τ

γ=

11G + λ

GD

B.5 Strength anisotropy

In the same way as it has been considered the deformability of a rock mass, expressionscan be developed indicating how strength is affected by the presence of discontinuities,starting with a single discontinuity and then extending to any number of discontinuities(Hudson and Harrison, 1997).

The strength of a sample of intact rock containing a single discontinuity can be es-tablished using the “single plane weakness” theory (Jaeger, 1960). Basically, the stressapplied to the sample is resolved into the normal and shear stresses on the plane ofweakness and the Mohr-Coulomb failure criterion applied to consider the possibility ofslip. The strength of the sample depends on the orientation of the discontinuity. If thediscontinuity is, for example, parallel or perpendicular to the applied loading, it willhave no effect on the sample strength. At some angles, however, the discontinuity willsignificantly reduce the strength of the sample. This is illustrate in Figure B.5 whichshows that the lowest strength occurs when the discontinuity normal is inclined at anangle of 45°+(φw/2) to the major applied principal stress.

The formula for the reduction in strength is found by establishing the normal andshear stress on the plane passing through the specimen and substituting these into the

B.5 Strength anisotropy 293

144 Rock masses

diagram, the minimum moduli are in a direction perpendicular to the sets, and the maximum moduli are at a direction of 45° to the sets. Like the discontinuity frequency, the directions of maximum and minimum moduli are not perpendicular.

A most interesting case occurs when k = 1, i.e. the normal and shear stiffnesses are equal, and the modulus is isotropic. The significance of even this very simple case of rock mass deformability for in situ testing and numerical modelling is apparent.

8.2 Strength In the same way as we have considered the deformability of a rock mass, expressions can be developed indicating how strength is affected by the presence of discontinuities, starting with a single discontinuity and then extending to any number of discontinuities. The initial approach is via the 'single plane of weakness' theory, attributable to Jaeger, whereby the strength of a sample of intact rock containing a single discontinuity can be established. Basically, the stress applied to the sample is resolved into the normal and shear stresses on the plane of weakness and the Mohr-Coulomb failure criterion (discussed in Chapter 6) applied to consider the possibility of slip.

The strength of the sample depends on the orientation of the discon-tinuity. If the discontinuity is, for example, parallel or perpendicular to the applied loading, it will have no effect on the sample strength. At some angles, however, the discontinuity will significantly reduce the strength of the sample. This is illustrated in Fig. 8.4 which shows that the lowest strength occurs when the discontinuity normal is inclined at an angle of 45° + (072) to the major applied principal stress. The formula for the reduction in strength is found by establishing the normal and shear stress on the plane passing through the specimen and substituting these into the Mohr-Coulomb failure criterion.

Given the geometry of the applied loading conditions in Fig. 8.4,

|T| = y2 (cTi - 0-3) sin 2/J^

(a, - a.)

45 + ^^ 2"

Figure 8.4 Effect of a discontinuity on the strength of a rock sample

^^K

Figure B.5: Effect of a discontinuity on the strength of a rock sample according withMohr-Coulomb criterion (Hudson and Harrison, 1997).

Mohr-Coulomb failure criterion. Given the geometry of the applied loading conditionsin Figure B.5,

|τ| = 12(σ1 − σ3) sin 2βw

σn =12(σ1 + σ3) +

12(σ1 − σ3) cos 2βw

Substituting these into the Mohr-Coulomb criterion, |τn| = cw + σn tan φw, and rearran-ging,

(σ1 − σ3) =2 (cw + σ3 tan φw)

(1− cot βw tan φw) sin 2βw

where cw and φw are the cohesion and the angle of friction for the discontinuity (i.e. planeof weakness), and βw is illustrated in Figure B.5. The plot of the equation in figure B.5shows the minimum strength and also the angles at which the sample strength becomesless than that of the intact rock. Outside of the region, failure occurs more easily on aplane other than the plane of weakness: the failure occurs at Mohr-Coulomb stress andthe predicted failure plane is oriented at 45°-(φw/2), independently of the orientation ofthe plane of weakness.

294 Rock anisotropy

Strength 145

(Jn = y 2 ( a i + (J3) + y 2 ( a i - G^) COS 2j3^

Substituting these into the Mohr-Coulomb criterion, | T| = CW + CTn tan 0^ and rearranging gives

( ^ 1 - ^ 3 ) = 2(c^+o-3tan0v^)

(1 - cot /3w tan 0vv) sin 2/J^

where Cw and 0vv are the cohesion and the angle of friction for the discon-tinuity (i.e. plane of weakness), and j3w is illustrated in Fig. 8.4. The plot of the equation in Fig. 8.4 shows the minimum strength and also the angles at which the sample strength becomes less than that of the intact rock.

An alternative presentation is via the Mohr's circle representation, as shown in Fig. 8.5. The Mohr-Coulomb failure loci for both the intact rock and the discontinuity are shown. We also show three Mohr's circles, A, B and C, representing the lowest strength, an intermediate case and the highest strength.

• Circle A represents the case when the failure locus for the discontinuity is just reached, i.e. for a discontinuity at the angle 2/3w° = 90° + 0 .°.

• Circle B is a case when failure can occur along the discontinuity for a range of angles, as indicated in the figure.

• Circle C represents the case where the circle touches the intact rock failure locus, i.e. where failure will occur in the intact rock if it has not already done so along the discontinuity.

The importance of these different failure mechanisms will be made clear in later chapters, when we consider the stresses around excavations in rock containing discontinuities. According to the circumstances, failure can either occur along the discontinuities or through the intact rock, depending on the relative orientations of the principal stresses and the discontinuities.

T = c + a,, tancj) Intact rock

Discontinuity T = c^ + a^ tan()^

2P^ = 90 + (1) Range of angles 2p^ = 90 + cj)

Figure 8.5 Mohr's circle representation of the possible modes of failure for rock containing a single plane of weakness. Figure B.6: Mohr’s circle representation of the possible modes of failure for rock contain-

ing a single plane of weakness (Harrison and Hudson, 2000).

An alternative presentation is via the Mohr’s circle representation, as shown in FigureB.6. The Mohr-Coulomb failure loci for both the intact rock and the discontinuity areshown. It is also shown three Mohr’s circle, A, B and C, representing the lowest strength,an intermediate case and the highest strength.

• Circle A represents the case when the failure locus for the discontinuity is justreached, i.e. for a discontinuity at the angle 2βw = 90° + φw.

• Circle B is a case when failure can occur along the discontinuity for a range ofangles, as indicated in figure.

• Circle C represents the case where the circle touches the intact rock failure locus,i.e. where the failure will occur in the intact rock if it has not already done so alongthe discontinuity.

Relations identical to the above have been derived from the equivalent Griffith crackmodel in which it is assumed that there are two population of cracks, long cracks lyingparallel to a given plane, the plane of weakness, and short cracks randomly oriented,and the cracks are closed and subjected to friction in the relative sliding of their surfaces(Paterson and Wong, 2005).

It is possible to consider, on the basis of the single plane of weakness theory, whatwould happen if there were two or more discontinuities at different orientations present

B.5 Strength anisotropy 295146 Rock masses

((7, - (TV)

^3

Figure 8.6 Strength of a rock mass containing multiple discontinuity sets.

We can consider, on the basis of this single plane of weakness theory, what would happen if there were two or more discontinuities at different orientations present in the rock sample. Each discontinuity would weaken the sample as indicated in Fig. 8.4, but the angular position of the strength minima would not coincide. As a result the rock is weakened in several different directions simultaneously as shown in Fig. 8.6. The material tends to become isotropic in strength, like a granular soil. When plotting the superimposed curves, care should be taken in interpreting the magnitude of p^ correctly for each of the discontinuities concerned.

The main advantage of the single plane of weakness theory is its simplicity and helpfulness in the interpretation of rock mass failure. We have presented here the two-dimensional case (applicable to plane stress) and one can imagine an extension to the general three-dimensional loading case in which none of the principal stresses is perpendicular to the discon-tinuity normal. In reality, the situation is rather complicated because the stresses will not be transmitted directly through the discontinuity. However, despite these shortcomings the authors feel that the advantages gained by understanding such idealized models do help in interpreting the far more complex behaviour of real rock masses.

We have already discussed the Hoek-Brown failure criterion in Chapter 6 in terms of its application to intact rock. The criterion is especially powerful in its application to rock masses due to the constants m and s being able to take on values which permit prediction of the strengths of a wide range of rock masses. Recent publications, i.e. Hoek and Brown (1988), Hoek (1990) and Hoek et al (1992), provide an update of the failure criterion as it applies to rock masses, together with the relation between the Hoek-Brown and Mohr-Coulomb criteria.

From the first of these publications, we have included Table 8.1 which indicates the relation between rock mass quality and the m and s material constants. The table also provides a guide to the relation between these constants and two commonly used rock mass characterization values, i.e. the CSIR and NGI classification ratings (the latter being described in Chapter 12)

Figure B.7: Strength of a rock mass containing multiple discontinuity sets (Harrison andHudson, 2000).

in the rock sample (Hudson and Harrison, 1997). Each discontinuity would weakenthe sample as indicated in Figure B.7, but the angular position of the strength wouldnot coincide. As a result the rock is weakened in several different directions simultan-eously (Figure B.7). The material tends to become isotropic in strength, like a granularsoil. When plotting the superimposed curves, care should be taken in interpreting themagnitude of βw correctly for each of the discontinuities concerned.

Figure B.8 demonstrates the anisotropy of compressive strength recorded for a seriesof test performed on a slate (Hudson and Harrison, 1997). This is an example of aniso-tropy characterised through application of the single plane of weakness theory, whichhave directionally built into its formulation.

296 Rock anisotropy

24(#

3 = 0

P -15

160t-

120

\

Inhomogeneity 167

Pi 1 P = 30 3 = 60

P -45

Le<iend • 0 •

• n • •

A V

Lowe

(7^ =

(T^ =

Confining pressure a^ MPa

1.7 2.7 3.6 3.2 6.0 7.0 8.0

10.0 12.0

r bound trends

0 10

P = 90^

P = 75

• A«

/ /

/ * / / * •

/ /

Angle between major principal stress and normal to cleavage - p° J_

10 20 30 40 J_ _1_ !S0 60 70 80 90

Figure 10.2 Compressive strength anisotropy in dark grey slate (after Brown et al, 1977).

values of a rock property. For example, in choosing the type of cutters for a tunnel boring machine, not only the mean compressive strength would be required but also the range within which the top, say, 10% of strengths occur.

In Fig. 10.3, we illustrate both the standard statistical procedure for characterizing a parameter through the full probability density curve which does not explicitly take the distances between sample locations into account) and the semi-variogram of geostatistics which does take these distances into account).

Where we encounter rock properties varying with location within a rock mass, there are three main approaches to the characterization procedure:

Figure B.8: Anisotropy of compressive strength in slate rock (Harrison and Hudson,2000).

B.6 Numerical modelling of anisotropic rock 297

B.6 Numerical modelling of anisotropic rock

Starting from the CHILE and DIANE assumptions, it will be helpful to consider aniso-tropy and inhomogeneity in the modelling procedures. According to the circumstances,failure can either occur along the discontinuities or through the intact rock, dependingon the relative orientations of the principal stresses and the discontinuities.

When the rock mass initially reaches its peak strength due to the application of stresses,either the intact rock or a fracture will begin to fail. Hence, the properties that governthe failure of rock masses are the failure properties of the intact rock and the fractures.The orientation of the applied stresses relative to the fractures governs how failure oc-curs, bearing in mind that the failure of a rock mass is a complex structural breakdownprocess following a similar curve to that in Figure B.1 for the intact rock.

The development of computer based numerical solutions helps to deal with morecomplex geometry and material properties. These techniques include finite difference,finite element, boundary element and distinct element formulations, providing the cap-ability of incorporating discontinuousness, anisotropy, inhomogeneity and more com-plex constitutive behaviour.

All the numerical techniques can accommodate wide variations in problem geometryand the presence of discontinuities. This is not the case for the inhomogeneity, aniso-tropy and constitutive behaviour relating to volume properties, because the individualelements in these numerical formulations should not be assigned a single value relatingto a volume property which may varying on a scale commensurate with the elementsthemselves. Developments aimed at overcoming these difficulties are the use of hybridnumerical formulations which recognize the advantages of the continuum and discon-tinuum component methods.

B.7 FDEM modelling of laboratory tests on anisotropic rock

The scope of this study is to simulate the behaviour of a layered rock specimen undera standard laboratory test. Data collected from uniaxial and triaxial laboratory tests ofLuserna gneiss (Barla et al., 1989) and Martinsburg slate (Donath, 1964) were comparedwith numerical results obtained using the Y2D code. Note that, in the following numer-ical analyses, only strength anisotropy will be analysed because the Y2D code takes noaccount of the deformability anisotropy.

298 Rock anisotropy

B.7.1 Numerical modelling of uniaxial tests of gneiss

The rock specimen to be considered is a cylinder of 54 mm in diameter with a ratioheight/diameter of 2.5:1. In a uniaxial test, the specimens are loaded axially up to fail-ure whereby the specimen is deformed and the axial and the radial deformation can bemeasured.

Several numerical simulations have been carried out by considering uniaxial com-pression tests of Luserna Gneiss specimens obtained from a site in Northwest Alps, inItaly (Barla et al., 1989). Each specimen was characterised by a different angle betweenthe direction of loading and the orientation of the joints planes. The spacing between thejoints was 30 mm (Figure B.9).

Figure B.9: Schematic of the rock specimen; σv1 is the axial stress, σv3 is the confiningpressure and β is the inclination of joints.

The properties of intact rock and joints are shown in Table B.1. A Mohr-Coulombmaterial with maximum tensile strength cut-off was used both for intact rock and joints.

B.7 FDEM modelling of laboratory tests on anisotropic rock 299

Rock sample Joints Loading platens

Elastic modulus, E (GPa) 30 - 193Poisson’s Ratio, ν (-) 0.3 - 0.29Density, ρ (kg/m3) 2700 - 8030Internal cohesion, ci (MPa) 18.30 13.97 -Internal friction angle, φi (°) 59.52 48.43 -Tensile strength, σvi (MPa) 17 4 -Fracture energy release rate, Gf (J/m2) 200 50 -

Table B.1: Material properties of gneiss, joints and loading platens.

An unstructured mesh with average element size of 1 mm was generated with Phase²and used in numerical analysis with the Y2D code. Figure B.10 shows the geometry(load, mesh and joints) of the seven specimens analysed.

To simulate a uniaxial test, a constant vertical velocity (0.25 mm/s) was applied tothe loading platens. Usually, in laboratory tests a quasi-static displacement rate (e.g.0.01mm/s) is applied. Considering the size of the time-step needed to satisfy the sta-bility condition of the explicit time integration scheme, a simulation which reproducesthe velocity applied in a laboratory test would have resulted in a very long computa-tion time. Tests carried out on the velocity applied to the loading platens demonstratethat a peak strength convergent to the quasi-static condition value can been obtainedby using a relatively high displacement rate (i.e. 0.25 mm/s). During the simulation,the reaction force at the nodes of the upper and lower platens, the vertical stress, they-displacement of a point of the specimen next to the platens, and the x-displacement ofthe lateral boundary of the sample have been recorded.

The stress-strain curves obtained from numerical modelling of uniaxial tests on gneissare plotted in Figure B.11. Table B.2 compares the values of the uniaxial compressivestrength obtained from laboratory tests, numerical modelling, and the theoretical valuesaccording to Mohr-Coulomb criterion (modified). This comparison is plotted on FigureB.12.

The effect of anisotropy on rock strength can be considered according to the theoret-ical model described in the first part of this appendix. As shown in Figure B.12 the loweststrength occurs when the discontinuity normal is inclined at an angle of 45°-(φw/2) (i.e.20.78°) to the major applied principal stress.

The fracture pattern obtained in the numerical modelling with the Y2D code, whenthe joints are vertical (i.e β=0°) is shown in Figure B.13. Fractures develop along thebedding planes first, then shear bands form. The splitting process can be considered

300 Rock anisotropy

Figure B.10: Jointed specimens of Luserna gneiss: geometry and mesh.

0° 15° 30° 45° 60° 75° 90°

UX laboratory test 116.21 101.68 92.46 75.68 116.68 127.96 152.44Mohr-Coulomb 134 80 92 134 134 134 134

Numerical 134 72 126 112 111 112 111

Table B.2: Comparison among laboratory, numerical and theoretical values of compress-ive strength from uniaxial test of Luserna gneiss.

B.7 FDEM modelling of laboratory tests on anisotropic rock 301

0

20

40

60

80

100

120

140

160

180

200

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

σ 1(M

Pa)

axial strain (%)

0° 15° 30° 45°

60° 75° 90°

Figure B.11: Uniaxial tests of Luserna gneiss: stress-strain curve for specimens charac-terised by a different inclination of joints (β). The curves have been obtained from theY2D numerical modelling of uniaxial tests.

0

20

40

60

80

100

120

140

160

180

200

0 15 30 45 60 75 90

σ 1(M

Pa)

joint inclination angle (°)

experimental

numerical

Mohr-Coulomb criterion (modified)

45 - φ/2

Figure B.12: Influence of anisotropy on rock strength (uniaxial tests of Luserna gneiss):comparison among numerical, theoretical and laboratory data.

302 Rock anisotropy

predominant. For the sample with β=15°, it was found that pure shear fractures developalong the bedding planes (Figure B.14), whereas when the fracture are horizontal (β=90°),inclined shear fractures are observed (Figure B.15).

Figure B.13: Uniaxial test of a Luserna gneiss specimen (β=0°): fracture patterns simu-lated with the Y2D code. Colours represent vertical stresses (red: compressive).

B.7 FDEM modelling of laboratory tests on anisotropic rock 303

Figure B.14: Uniaxial test of a Luserna gneiss specimen (β=15°): fracture patterns simu-lated with the Y2D code. Colours represent vertical stresses (red: compressive).

Figure B.15: Uniaxial test of a Luserna gneiss specimen (β=90°): fracture patterns simu-lated with the Y2D code. Colours represent vertical stresses (red: compressive).

304 Rock anisotropy

B.7.2 Numerical modelling of uniaxial and triaxial tests of slate

The specimens under investigation is a rock cylinder of 54 mm in diameter with a ratioheight/diameter of 2.5:1 for uniaxial test, and of 2:1 for triaxial test. In this numericalstudy, each rock specimen has been characterised by joints with a different inclination.Actually, the angle between the direction of loading and the orientation of the joints hasbeen varied from 0° to 90° with an increment of 15°. The spacing between the jointsplanes is 30 mm.

The properties of intact rock and joints are listed in Table B.3. The intact rock and thejoints behave according to the Mohr-Coulomb criterion with a maximum tensile strengthcut-off. Each specimen was discretised with an unstructured mesh characterised by av-erage element size of 1 mm. The mesh was generated in Phase², then imported in Y2Dcode. A triaxial test has been simulated for each rock specimen shown in Figure B.16,whereas the uniaxial tests have been simulated for rock specimen with the same geo-metry used in the numerical study on gneiss (Figure B.10). The boundary conditionsapplied to the loading platens are a constant vertical velocity of 0.25 mm/s and, in thetriaxial tests, a confinement pressure of 10.5 MPa.

Rock sample Joints Loading platens

Elastic modulus, E (GPa) 30 - 193Poisson’s Ratio, ν (-) 0.3 - 0.29Density, ρ (kg/m3) 2600 - 8030Internal cohesion, ci (MPa) 38.34 7.33 -Internal friction angle, φi (°) 37.84 20.18 -Tensile strength, σvi (MPa) 19 1 -Fracture energy release rate, Gf (J/m2) 200 50 -

Table B.3: Material properties of Martinsburg slate, joints and loading platens.

The stress-strain curves resulting from the simulation of uniaxial tests and triaxialtests with the Y2D code are shown respectively in Figure B.17 and Figure B.18, whereasthe values obtained from laboratory tests and numerical analyses are indicated in TableB.4. Using a confinement pressure, the peak load increases, and also the residual valueafter the peak increases.

As described in the first part of this appendix, the effect of anisotropy on rock strengthcan be considered with the relationships based on the Mohr Coulomb strength criterion,modified to take into account the presence of joints. The strength of the specimen de-pends on the orientation of the principal stresses relative to fracture orientation. Assum-

B.7 FDEM modelling of laboratory tests on anisotropic rock 305

Figure B.16: Triaxial tests of Martinsburg slate. Multiple jointed specimens: mesh andjoints orientation.

0° 15° 30° 45° 60° 75° 90°

Mohr-Coulomb 157 33 21 23 47 157 157Numerical 163 31 24 43 147 130 149

TX laboratory test 169 79 45 53 103 164 250Mohr-Coulomb 200 60 43 46 85 200 200

Numerical 222 76 62 99 183 192 192

Table B.4: Comparison among laboratory, numerical and theoretical values of compress-ive strength from uniaxial tests and triaxial tests of Martinsburg slate.

306 Rock anisotropy

-20.0

.0

20.0

40.0

60.0

80.0

100.0

120.0

140.0

160.0

180.0

200.0

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

σ 1(M

Pa)

axial strain (%)

0° 15° 30° 45°

60° 75° 90°

Figure B.17: Stress-strain curves for uniaxial tests of Martinsburg slate specimens withdifferent inclination of joints (β).

-20.0

.0

20.0

40.0

60.0

80.0

100.0

120.0

140.0

160.0

180.0

200.0

220.0

240.0

0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 1.80 2.00 2.20 2.40

σ 1(M

Pa)

axial strain (%)

0° 15° 30° 45°

60° 75° 90°

Figure B.18: Stress-strain curves for triaxial tests of Martinsburg slate specimens withdifferent inclination of joints (β).

B.7 FDEM modelling of laboratory tests on anisotropic rock 307

ing that failure is induced when the normal and shear stress components on the fracturesatisfy the Mohr-Coulomb failure criterion, it is possible to develop an expression for thespecimen strength as a function of the angle between the maximum principal stress andthe normal to the fracture. As shown in Figure B.19 the lowest strength occurs when thediscontinuity normal is inclined at an angle of 45°-(φw/2) (i.e. 34.92°) to the maximumapplied principal stress.

Figures B.20÷B.22 show some screenshots of uniaxial tests on slate, simulated withthe Y2D code. As shown in Figure B.20, the rock specimen exhibit vertical tensile split-ting along the joints at the beginning of fracturing, then shear fractures occur. Whenthe angle between the joints and the specimen axis is equal to 30°, the fractures occurmainly on joints, as shown in Figure B.21. In Figure B.22 the uniaxial test on slate withhorizontal joints (β=90°) is shown; the failure occurs inside the material as shear bandsare created, due to the coalescence of single micro-cracks.

Figures B.23÷B.25 show the fracture patterns for the triaxial tests. It is interestingto note that the application of a confining pressure inhibits cracking thus making thesamples more ductile. When β=30° failure occurs along joints, whereas for β=0° andβ=90° shear bands are created inside the material and lead the specimen to failure.

308 Rock anisotropy

0

50

100

150

200

250

300

350

400

0 15 30 45 60 75 90

σ 1(M

Pa)

β (°)

UX test - numerical simulation

Mohr-Coulomb (modified)

TX test - experimental results form Donath 1964

TX test - numerical simulation

Mohr-Coulomb (modified)

Figure B.19: Influence of anisotropy on rock strength. Comparison among the theoret-ical curve of Mohr Coulomb modified criterion, laboratory data and the values of thepeak principal stress obtained in the numerical simulation of uniaxial and triaxial testsof Martinsburg slate.

B.7 FDEM modelling of laboratory tests on anisotropic rock 309

Figure B.20: Uniaxial test of a Martinsburg slate specimen (β=0°): fracture patterns sim-ulated with the Y2D code. Colours represent vertical stresses (red: compressive).

Figure B.21: Uniaxial test of a slate specimen (β=30°): fracture patterns simulated withthe Y2D code. Colours represent vertical stresses (red: compressive).

310 Rock anisotropy

Figure B.22: Uniaxial test of a Martinsburg slate specimen (β=90°): fracture patternssimulated with the Y2D code. Colours represent vertical stresses (red: compressive).

Figure B.23: Triaxial test of a Martinsburg slate specimen (β=0°): fracture patterns simu-lated with the Y2D code. Colours represent vertical stresses (red: compressive).

B.7 FDEM modelling of laboratory tests on anisotropic rock 311

Figure B.24: Triaxial test of a Martinsburg slate specimen (β=30°): fracture patterns sim-ulated with the Y2D code. Colours represent vertical stresses (red: compressive).

Figure B.25: Triaxial test of a Martinsburg slate specimen (β=90°): fracture patterns sim-ulated with the Y2D code. Colours represent vertical stresses (red: compressive).

312 Rock anisotropy

B.8 Summary

Anisotropy is a key aspect of rock engineering problems. In a rock mass, the discontinu-ity frequency can vary very significantly with direction and properties such as the rockmass deformability and strength will be functions of the discontinuity frequency.

For deformability, the architecture of the compliance matrix takes into account thelinking between stresses and strains and hence also explicitly quantifies some aniso-tropy. However, not all rock properties have anisotropy incorporated into their charac-terization. For example, compressive strength is usually assumed to be a scalar value,which is by definition directionless, but measurement of compressive strength should bequalified with information on the direction of loading relative to the rock structure.

The effect of a single fracture on rock strength can be studied using the single planeof weakness theory which state that the strength of the specimen will then depend onthe orientation of the principal stresses relative to fracture orientation.

The numerical simulations presented in this appendix demonstrate the capability ofFDEM for quantitatively modelling the failure of a layered rock specimen under uni-axial and triaxial stress conditions. The results obtained from numerical analysis matchwell with the experimental results and the behaviour of jointed rock mass have beenaccurately predict.

References

Barla, G., Chiappone, A., and Scavia, C. (1989). “Anisotropia di resistenza delle roccein condizioni di compressione triassiale.” Rivista Italiana di Geotecnica, 3: 105–120.

Donath, F. (1964). “Strength variation and deformational behaviour in anisotropicrocks.” In W. R. Judd, editor, “State of Stress in the Earth’s Crust,” American ElsevierPublishing Co., New York.

Harrison, J. P. and Hudson, J. A. (2000). Engineering Rock Mechanics Part II. IllustrativeWorked Examples. Elsevier Ltd.

Hudson, J. A. and Harrison, J. P. (1997). Engineering Rock Mechanics. An Introduction tothe Principles. Elsevier Ltd.

Jaeger, J. (1960). “Shear failure of anisotropic rocks.” Geological Magazine, 97: 65–72.

Paterson, M. and Wong, T.-f. (2005). Experimental Rock Deformation - The Brittle Field.Springer.

313

314 Rock anisotropy

List of Figures

2.1 Kinematically admissible and inadmissible landslides (Hutchinson, 2006) 132.2 Rock creep typologies after Ter-Stepanian (1966) . . . . . . . . . . . . . . . 162.3 Typical profiles of sackung from Zischinsky (1966) . . . . . . . . . . . . . . 172.4 Examples of creep types from Nemcok et al. (1972) . . . . . . . . . . . . . . 192.5 Cross section through the ridge Sivy mountain (Nemcok and Baliak, 1977) 202.6 Cross section of Contact Mountain, Montana, according to Radbruch-Hall

et al. (1976) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.7 Geological cross section at the Beauregard dam site showing the accumu-

lation of fluvial and glaciofluvial deposits (Desio, 1973) . . . . . . . . . . . 232.8 Morpho-structural features typical of a DSGSD (Agliardi et al.,2001) . . . 252.9 Cross section of the Chabenec Southern slope (from Mahr, 1977) . . . . . . 262.10 Development of slope failure according to Feda (1973) . . . . . . . . . . . . 272.11 Deformation occurring in a sackung (Sorriso-Valvo and Dramis, 1994) . . 282.12 Scheme of gravitational disequilibrium in the culmination part of the fly-

sch Carpathians (Hradechy and Panek, 2008) . . . . . . . . . . . . . . . . . 30

3.1 Typical problems, critical parameters, methods of analysis and acceptab-ility criteria for slope (Hoek, 1991) . . . . . . . . . . . . . . . . . . . . . . . 39

3.2 Flowchart illustrating three levels of landslide analysis (Stead et al., 2006) 413.3 Prandtl’s prism transition zone (Stead et al., 2006) . . . . . . . . . . . . . . 423.4 Distinct-element strain-softening model (Eberhardt et al., 2002) . . . . . . 493.5 Simulation of a stepped-path failure (Stead et al., 2006) . . . . . . . . . . . 513.6 Simulation of a 50 m high rock slope with a bi-planar failure surface (Stead

et al., 2006) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 523.7 Stages of a “total slope failure” analysis with the appropriate Level I, II and

III analysis codes (Stead et al. 2006, modified) . . . . . . . . . . . . . . . . . 54

315

316 LIST OF FIGURES

3.8 Massive rock slope sliding mechanisms as controlled by progressive shearplane development and internal rock mass damage (Eberhardt et al., 2003) 57

3.9 Geological section and geometry of the 1991 Randa rockslides (Stead andCoggan, 2006) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.10 Combined finite-discrete element analysis of the Randa rockslide (Steadand Coggan, 2006) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.11 The Elm rockslide (Heim, 1932) . . . . . . . . . . . . . . . . . . . . . . . . . 603.12 Stages in the combined finite-discrete element simulation of the Elm slide

(Stead and Coggan, 2006) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 613.13 Combined finite-discrete element “total slope failure” analysis of the Delabole

slate quarry failure (Stead and Coggan, 2006) . . . . . . . . . . . . . . . . . 623.14 The Alpetto mine rockslide (Barla et al., 2011) . . . . . . . . . . . . . . . . . 643.15 Final topographic profile obtained with numerical simulations (Barla et

al., 2011) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.1 Suitability of different numerical methods for an excavation in a rock mass 744.2 Node A penetrates through point B and exits through point C, thus nu-

merically creating energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 784.3 Distributed and concentrated contact force approach . . . . . . . . . . . . 794.4 Contact force due to an infinitesimal overlap around points Pc and Pt . . . 814.5 Distribution of contact force between the target triangle and an edge of

contactor triangle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 834.6 Equivalent nodal forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 844.7 Contact between two finite elements . . . . . . . . . . . . . . . . . . . . . . 864.8 Space subdivided into identical square cell for the NBS contact detection

algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 904.9 Contact checking mask . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 904.10 Frames of reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 934.11 Physical meaning of deformation gradient . . . . . . . . . . . . . . . . . . . 934.12 Cauchy stress tensor components in deformed configuration . . . . . . . . 964.13 The central difference time integration scheme . . . . . . . . . . . . . . . . 1004.14 A typical stress-displacement diagram for rock under uniaxial tension . . 1034.15 Typical strain softening curve defined in terms of strains . . . . . . . . . . 1054.16 Strain softening defined in terms of displacements . . . . . . . . . . . . . . 1064.17 Single crack model with bonding stress . . . . . . . . . . . . . . . . . . . . 1064.18 A singular stress field approximated by constant stress finite elements . . 109

5.1 Screenshot of Y-GUI interface . . . . . . . . . . . . . . . . . . . . . . . . . . 1165.2 Model configuration for planar failure analysis . . . . . . . . . . . . . . . . 119

Bibliography 317

5.3 Planar failure simulation modelled using Y2D code . . . . . . . . . . . . . 1225.4 Planar failure simulation modelled using Y2D code (joint friction angle φ

= 44°) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1235.5 Model configuration for bi-planar failure analysis . . . . . . . . . . . . . . 1245.6 Bi-planar failure mechanism simulation with Y2D code . . . . . . . . . . . 1265.7 Mesh configuration for multi-step path failure . . . . . . . . . . . . . . . . 1285.8 Numerical simulation of multi step-path failure using Y2D code . . . . . . 1305.9 FEM simulation: maximum shear strain and deformed mesh . . . . . . . . 1305.10 Multi-step path failure simulation with Y2D . . . . . . . . . . . . . . . . . 1315.11 Block toppling (Goodman and Bray,1976) . . . . . . . . . . . . . . . . . . . 1325.12 Geometric scheme for limit equilibrium analysis of toppling on a stepped

base . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1335.13 Conditions for toppling and for sliding of the nth block (Goodman and

Bray, 1976) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1345.14 Geometry of a toppling slope (Goodman and Bray, 1976) . . . . . . . . . . 1375.15 Basic slope geometry used in the numerical modelling . . . . . . . . . . . 1395.16 Block toppling model with elastic blocks (UDEC) . . . . . . . . . . . . . . . 1405.17 Block toppling model with elastoplastic blocks (UDEC) . . . . . . . . . . . 1425.18 Finite element mesh used for FDEM analysis of block toppling . . . . . . . 1435.19 Total kinetic energy during the first stage of FDEM analysis . . . . . . . . 1445.20 Case 1. Velocity distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . 1445.21 Case 2. Velocity distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . 1455.22 Case 1. Y2D analysis of block toppling. Selected output . . . . . . . . . . . 1475.23 Case 2. Y2D analysis of block toppling. Selected output . . . . . . . . . . . 1485.24 Geometry of the slope problem . . . . . . . . . . . . . . . . . . . . . . . . . 1495.25 Global minimum sliding surface with the factor of safety computed by

LEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1505.26 Finite Element mesh and boundary conditions . . . . . . . . . . . . . . . . 1515.27 Maximum shear strains and deformed mesh when sliding occurs in the

model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1525.28 FDEM model with different excavation layers of 5 m and 2 m thickness . . 1535.29 Total kinetic energy during Stage 1 . . . . . . . . . . . . . . . . . . . . . . . 1545.30 Monitored points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1555.31 Comparison between computed with FDEM and expected vertical stresses 1555.32 Displacement vectors and deformed assembly when instability is triggered

in the FDEM model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1575.33 Subsequent screenshots of the slope instability . . . . . . . . . . . . . . . . 1585.34 Factor of safety of the slope . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

318 LIST OF FIGURES

6.1 Area of interest and view of the dam site (Barla et al., 2006) . . . . . . . . . 1646.2 Geological cross section at the Beauregard dam site (Desio, 1973, modified) 1656.3 Geological and morpho-structural map of the Beauregard DSGSD (Barla

et al., 2010a) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1676.4 Map showing the investigated area and three zones characterised by dif-

ferent geomorphological features (Barla et al., 2006) . . . . . . . . . . . . . 1686.5 Monitoring system installed at the Beauregard site (Barla et al., 2010) . . . 1696.6 Slope cumulated displacements at the top of plumb lines and snow height

(a); snow height and water level in piezometers (b) (Barla et al., 2010a) . . 1716.7 K1, K4, K13 and K19 total station targets: comparison between predicted

and measured displacements (Barla et al., 2010a) . . . . . . . . . . . . . . . 1726.8 Map showing displacements vectors measured on the landslide monitor-

ing network (Barla et al., 2010a) . . . . . . . . . . . . . . . . . . . . . . . . . 1736.9 Key instabilities developed in Beauregard massive landslide . . . . . . . . 1746.10 Stereograms and rose diagrams of main discontinuity sets (Barla et al.

2010a, modified) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1766.11 Map showing the cross sections and the orientation of the schistosity planes1776.12 Geological sections of Bochat Ridge . . . . . . . . . . . . . . . . . . . . . . 1796.13 Arp Vieille geological section . . . . . . . . . . . . . . . . . . . . . . . . . . 1816.14 Bochat geological section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1836.15 Site investigations results in the Bochat area (Barla et al. 2010a, modified) 1846.16 Geological section E–E’ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

7.1 The main geological features, the constitutive laws, and the correspondinggeotechnical parameters adopted in the numerical modelling (Barla et al.,2006) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

7.2 Basic slope geometry for the K4-schistosity model . . . . . . . . . . . . . . 1957.3 FEM simulation (K4-shistosity model) . . . . . . . . . . . . . . . . . . . . . . 1977.4 Basic slope geometry for the K2-K3 model . . . . . . . . . . . . . . . . . . 1987.5 FEM simulation (K2-K3 model) . . . . . . . . . . . . . . . . . . . . . . . . . 2007.6 Total kinetic energy during Stage 1 (K4-schistosity model) . . . . . . . . . . 2037.7 FDEM simulation of instability (K4-schistosity model) for σvt,rock mass=7 MPa 2067.8 FDEM simulation of instability (K4-schistosity model) for σvt,rock mass=3 MPa 2077.9 FDEM simulation of instability (K2-K3 model) for σvt,rock mass=3 MPa . . . . 2087.10 View of the Bochat Ridge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2097.11 Photo of the thick talus located at the toe of the Bochat Ridge . . . . . . . . 2107.12 Trenches in the upper part of the Bochat Ridge . . . . . . . . . . . . . . . . 2117.13 Particular of the photo of the Bochat Ridge . . . . . . . . . . . . . . . . . . 212

Bibliography 319

7.14 The main discontinuity sets sets at the Bochat Ridge . . . . . . . . . . . . . 2137.15 Sliding and toppling instability of a block on an inclined plane . . . . . . . 2147.16 Basic slope geometry used in the FEM analysis (section B-B’) . . . . . . . . 2157.17 Total displacements, yielded elements and joints at the end of the FEM

analysis (section B-B’) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2177.18 FEM model of section B-B’ with a lower value of spacing for schistosity . . 2177.19 Total kinetic energy during Stage 1 (section B-B’) . . . . . . . . . . . . . . . 2187.20 FDEM simulation of block toppling (section B-B’, σvt,rock mass=3 MPa) . . . 2197.21 FDEM simulation of block toppling (section B-B’, σvt,rock mass=7 MPa) . . . 2207.22 View of block toppling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2227.23 Basic slope geometry for the FEM analysis (K1-schistosity model)) . . . . . 2247.24 FEM simulation (K1-schistosity model) . . . . . . . . . . . . . . . . . . . . . 2267.25 Total kinetic energy during Stage 1 (K1-schistosity model) . . . . . . . . . . 2287.26 FDEM simulation of instability (K1-schistosity model) for σvt,rock mass=3 MPa 2297.27 FDEM simulation of instability (K1-schistosity model) for σvt,rock mass=7 MPa 2307.28 Details of the main scarp: vertical cliff where blocks are defined by the

main joint sets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2317.29 View of the wide trenches (K1 and K4 joint sets) located above the crest . . 2317.30 Basic slope geometry for the two models of Bochat section (D-D’) . . . . . 2347.31 Phase² model (FEM) with near limit equilibrium parameters along the

basal shear zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2367.32 Total kinetic energy during Stage 1 (circular surface model) . . . . . . . . . 2387.33 FDEM simulation of instability (rigid-block model, friction angle of 19°) . . 2407.34 FDEM simulation of instability (circular surface model, friction angle of 22°) 241

A.1 Regular quadrilateral grid for the FDM and irregular quadrilateral gridfor the FVM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255

A.2 Discretization of blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264A.3 Mechanical representation of contacts in the 2D DEM . . . . . . . . . . . . 266A.4 Representation of rock fractures for the flow equation solution . . . . . . . 270A.5 Hybrid model for a rock mass . . . . . . . . . . . . . . . . . . . . . . . . . . 272A.6 Fracture elements in FEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275A.7 Illustrative meshes for fracture analysis with BEM . . . . . . . . . . . . . . 276

B.1 The complete stress-strain curve illustrating various mechanical paramet-ers (Hudson and Harrison, 1997) . . . . . . . . . . . . . . . . . . . . . . . . 286

B.2 Type of fracturing in rock (Hudson and Harrison, 1997) . . . . . . . . . . . 287B.3 Relation between vertical and elastically induced horizontal stresses for

the different type of isotropy (Hudson and Harrison, 1997) . . . . . . . . . 289

320 LIST OF FIGURES

B.4 The modulus of deformation of a rock mass containing a discontinuity set(Hudson and Harrison, 1997) . . . . . . . . . . . . . . . . . . . . . . . . . . 291

B.5 Effect of a discontinuity on the strength of a rock sample according withMohr-Coulomb criterion (Hudson and Harrison, 1997) . . . . . . . . . . . 293

B.6 Mohr’s circle representation of the possible modes of failure for rock con-taining a single plane of weakness (Hudson and Harrison, 1997) . . . . . . 294

B.7 Strength of a rock mass containing multiple discontinuity sets (Hudsonand Harrison, 1997) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295

B.8 Anisotropy of compressive strength in slate rock (Hudson and Harrison,1997) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296

B.9 Schematic of the rock sample . . . . . . . . . . . . . . . . . . . . . . . . . . 298B.10 Jointed specimens of Luserna gneiss: geometry and mesh . . . . . . . . . 300B.11 Uniaxial tests of Luserna gneiss: stress-strain curves from the Y2D numer-

ical modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301B.12 Influence of anisotropy on rock strength (uniaxial tests of Luserna gneiss) 301B.13 Uniaxial test of a Luserna gneiss specimen (β=0°) . . . . . . . . . . . . . . . 302B.14 Uniaxial test of a Luserna gneiss specimen (β=15°) . . . . . . . . . . . . . . 303B.15 Uniaxial test of a Luserna gneiss specimen (β=90°) . . . . . . . . . . . . . . 303B.16 Triaxial tests on Martinsburg slate. Multiple jointed specimens . . . . . . . 305B.17 Stress-strain curve for uniaxial tests of Martinsburg slate specimens . . . . 306B.18 Stress-strain curve for triaxial tests of Martinsburg slate specimens . . . . 306B.19 Influence of anisotropy on rock strength – uniaxial and triaxial test of Mar-

tinsburg slate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308B.20 Uniaxial test of a Martinsburg slate specimen (β=0°) . . . . . . . . . . . . . 309B.21 Uniaxial test of a Martinsburg slate specimen (β=30°) . . . . . . . . . . . . 309B.22 Uniaxial test of a Martinsburg slate specimen (β=90°) . . . . . . . . . . . . 310B.23 Triaxial test of a Martinsburg slate specimen (β=0°) . . . . . . . . . . . . . . 310B.24 Triaxial test of a Martinsburg slate specimen (β=30°) . . . . . . . . . . . . . 311B.25 Triaxial test of a Martinsburg slate specimen (β=90°) . . . . . . . . . . . . . 311

List of Tables

3.1 Conventional methods of analysis (Stead et al., 2006) . . . . . . . . . . . . 443.2 Numerical methods of rock slope analysis (Stead et al., 2006) . . . . . . . . 463.3 Advanced hybrid numerical methods of analysis (Stead et al., 2006) . . . . 50

5.1 Data structure of Y2D code (from Mahabadi et al., 2010) . . . . . . . . . . . 1155.2 Rock mass and discontinuity properties for planar failure analysis . . . . . 1215.3 Rock mass and discontinuity properties for bi-planar failure analysis . . . 1255.4 Rock mass and discontinuity properties (multi-step path failure) . . . . . . 1295.5 Example from Goodman & Bray (1976) . . . . . . . . . . . . . . . . . . . . 1375.6 Analytical solution of block toppling . . . . . . . . . . . . . . . . . . . . . . 1385.7 Rock mass and joints properties (block toppling - UDEC) . . . . . . . . . . 1405.8 Rock mass properties (block toppling - elastoplastic blocks - UDEC) . . . . 1415.9 Ground properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1505.10 Computed factor of safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

6.1 Mean spacing of discontinuity sets . . . . . . . . . . . . . . . . . . . . . . . 1776.2 Range of properties of joint sets on Bochat Ridge . . . . . . . . . . . . . . . 1786.3 Range of properties of joint sets on Arp Vieille area . . . . . . . . . . . . . 1816.4 Range of properties of joint sets on Bochat area . . . . . . . . . . . . . . . . 183

7.1 Range of dip and dip direction of main joint sets (Bochat Ridge) . . . . . . 1937.2 Rock mass deformability and strength properties . . . . . . . . . . . . . . . 1957.3 Main joint sets and schistosity strength properties for FEM models . . . . 1967.4 Joint network setting for the K4-schistosity model . . . . . . . . . . . . . . 1967.5 Joint network setting for the K2-K3 model . . . . . . . . . . . . . . . . . . . 1997.6 Rock mass properties and numerical parameters used in the FDEM study 2017.7 Main joint sets and schistosity strength properties for FDEM models . . . 202

321

322 LIST OF TABLES

7.8 Section B-B’: joint network setting . . . . . . . . . . . . . . . . . . . . . . . 2167.9 Natural joints and schistosity strength properties (FEM analyses) . . . . . 2167.10 Range of dip and dip direction of main joint sets (Arp Vieille) . . . . . . . 2237.11 Joint network setting for section C-C’ . . . . . . . . . . . . . . . . . . . . . 2257.12 Rock mass deformability and strength properties . . . . . . . . . . . . . . . 2337.13 Computed Strength Reduction Factor in the FEM analyses . . . . . . . . . 2347.14 Rock mass and shear zone mechanical properties used in the FDEM study 2377.15 Trenches and faults strength properties for FDEM models . . . . . . . . . . 237

A.1 Types of contacts for polygons and polyhedral . . . . . . . . . . . . . . . . 265

B.1 Material properties of gneiss, joints and loading platens . . . . . . . . . . . 299B.2 Comparison among laboratory, numerical and theoretical values of com-

pressive strength from uniaxial test of Luserna gneiss . . . . . . . . . . . . 300B.3 Material properties of Martinsburg slate, joints and loading platens . . . . 304B.4 Comparison among laboratory, numerical and theoretical values of com-

pressive strength from uniaxial test and triaxial tests of Martinsburg slate 305