Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

INTERNATIONAL JOURNAL FOR NUMERICAL AND ANALYTICAL METHODS IN GEOMECHANICS

Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101


Published online 22 August 2008 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/nag.735

Coupled HM analysis using zero-thickness interface elements with


double nodes. Part I: Theoretical model

J. M. Segura1, ‡ and I. Carol2, ∗, †


1 School of Earth and Environment, University of Leeds, Woodhouse Lane, LS2 9JT, Leeds, U.K.
2 ETSECCPB (School of Civil Engineering), UPC (Technical University of Catalonia), Jordi Girona 1-3,
E-08034 Barcelona, Spain

SUMMARY
In recent years, the authors have proposed a new double-node zero-thickness interface element for diffu-
sion analysis via the finite element method (FEM) (Int. J. Numer. Anal. Meth. Geomech. 2004; 28(9):
947–962). In the present paper, that formulation is combined with an existing mechanical formulation
in order to obtain a fully coupled hydro-mechanical (or HM) model applicable to fractured/fracturing
geomaterials. Each element (continuum or interface) is formulated in terms of the displacements (u)
and the fluid pressure ( p) at the nodes. After assembly, a particular expression of the traditional ‘u– p’
system of coupled equations is obtained, which is highly non-linear due to the strong dependence between
the permeability and the aperture of discontinuities. The formulation is valid for both pre-existing and
developing discontinuities by using the appropriate constitutive model that relates effective stresses to
relative displacements in the interface. The system of coupled equations is solved following two different
numerical approaches: staggered and fully coupled. In the latter, the Newton–Raphson method is used,
and it is shown that the Jacobian matrix becomes non-symmetric due to the dependence of the disconti-
nuity permeability on the aperture. In the part II companion paper (Int. J. Numer. Anal. Meth. Geomech.
2008; DOI: 10.1002/nag.730), the formulation proposed is verified and illustrated with some application
examples. Copyright q 2008 John Wiley & Sons, Ltd.

Received 16 February 2008; Revised 7 June 2008; Accepted 9 June 2008

KEY WORDS: geomechanics; reservoir engineering; discontinuities; discrete crack; zero-thickness inter-
face; HM coupling; consolidation; hydraulic fracture

∗ Correspondence to: I. Carol, ETSECCPB-UPC Jordi Girona, 1-3, E-08034 Barcelona, Spain.

E-mail: ignacio.carol@upc.edu

Formerly at ETSECCPB (School of Civil Engineering), UPC (Technical University of Catalonia), Jordi Girona 1-3,
E-08034 Barcelona, Spain.

Contract/grant sponsor: AGAUR-DURSI


Contract/grant sponsor: ETSECCPB
Contract/grant sponsor: UPC
Contract/grant sponsor: MEC; contract/grant numbers: MAT2005-24522, BIA2006-12717
Contract/grant sponsor: MFOM; contract/grant number: 80015/A04

Copyright q 2008 John Wiley & Sons, Ltd.


2084 J. M. SEGURA AND I. CAROL

1. INTRODUCTION

Geomaterials usually contain pores and other cavities that are deformable and filled with fluid
under saturated or unsaturated conditions. The pressure of the fluid in the void space (mainly
liquid water) and the deformation of the medium are intimately related in the so-called hydro-
mechanical (HM) coupling, which has been studied in depth during last decades since the classical
contributions of Terzaghi [1] and Biot [2] to consolidation analysis in geomechanics. However,
rock, concrete and other quasi-brittle materials differ from soils in that they generally contain or
develop discontinuities. A main distinction compared to a porous medium must be made here, since
a discontinuity besides opening or closing (i.e. expansion and contraction in porous media) due to
fluid pressure variations, it can also propagate, which can be thought as an additional degree of
coupling. Another factor of complexity is the fluid exchange between continuum and discontinuity
if the continuum is a porous material.
The importance of geomechanical discontinuities in civil and other branches of engineering
(petroleum, environmental, etc.) is clear since discontinuities constitute preferential planes for
the deformation and flow in the medium, therefore strongly contributing to the deformability,
strength and permeability of the system. Their presence is often crucial in engineering projects
both on the surface (e.g. safety of structures, slope stability) or underground (e.g. aquifers, oil or
gas reservoirs, tunnels, waste disposal). Moreover, new discontinuities may appear and propagate
due to the effect of fluid pressure, for instance in the upstream face of dams or in hydraulic
fracture processes in petroleum engineering. In all these problems the coupling loops between
fluid flow, fracture mechanics and geomechanics are usually strong and it becomes necessary
to integrate all the components to appropriately describe the simultaneous flow of fluid in the
void space (pores and geological discontinuities), deformation of the geomaterial, and possible
development of fractures. This article focuses on the HM numerical modelling of both pre-existing
and developing discontinuities in saturated porous materials and considers the coupling loops
between the different processes in an integrated manner, both from the physics of the problem and
the numerical viewpoints.
From a numerical perspective, the solution of the stress-flow coupled equations can be achieved
according to two different procedures: fully coupled (or monolithic) and iteratively coupled (or
staggered). The fully coupled formulation of the problem directly solves the set of coupled equations
defining the problem, and it is therefore the most straightforward procedure. The ‘u– p’ is probably
the most widespread FE formulation of this type, in which the vector of main unknowns includes the
displacements (u) and the fluid pressure ( p) at the nodes [3]. On the other hand, in the partitioned
solution approach the system of equations is reorganized in such a way that a conventional
stress and flow codes are sequentially called to reach the coupled solution by introduction of the
appropriate coupling loops between simulators. It is also common to combine a finite differences
(FD) code for the flow problem with a finite element (FE) simulator for the mechanics analysis.
Iterations between the geomechanical and the flow codes are performed at every time-step until a
certain tolerance on the solutions is attained. This method should converge to the fully (monolithic)
coupled solution of the problem if iterated to full convergence as stated in Settari and Walters [4].
Lewis and Schrefler [5] review the numerical properties of the staggered scheme. The monolithic
technique is expected to be more consistent and robust. However, if separate mechanical and flow
codes are available, the partitioned approach also exhibits some advantages; this procedure will
use two codes that although developed to solve decoupled problems, they will usually be powerful,
with many advanced capabilities and highly validated tools that can be linked in a modular fashion.

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2085

This modular characteristic of the method allows the programmer to include different coupling
factors depending on the needs of the analysis, and to control these coupling loops much more
easily. On the other hand, when the coupling between the two fields (displacement and fluid
pressure) is very strong, the convergence of the iterative procedure in the staggered scheme may be
difficult.
The modular or partitioned scheme can be used also to represent different degrees of coupling
depending on the degree of information exchange: in an explicitly coupled scheme no iterations
are performed, but at every time-step the last pressure or displacement solution available is used to
advance the solution in time; on the other hand, in the one-way coupling procedure (also simply
called decoupled) the coupling effect is taken in just one direction, and usually the flow problem
is solved without the influence of the displacements and the fluid pressure history obtained in this
way is used afterward by the mechanical code to calculate the evolution of the displacement field.
Details and examples on these strategies can be found in for example [4–8].
Various approaches are available for the representation of discontinuities. Basically, one can
choose between equivalent continuum [9], double continuum [10, 11], or discrete fracture HM
formulations, and in the latter case one may still use ‘thin-layer’ [12] or ‘zero-thickness’ inter-
face elements. The work described in this paper concentrates on the discrete representation of
discontinuities or cracks using “zero-thickness” interface elements, which allows the simulation of
both pre-existing discontinuities and developing cracks. In the literature, the totally coupled HM
formulation of discontinuities is not common in combination with the discrete crack approach, and
the partitioned procedure usually combines the finite element method (FEM) for the mechanical
analysis with FD for the flow problem, as in the work of Papanastasiou on hydraulic fracture
with no leak-off [13]. Boone et al. [14, 15] include some fluid exchange between the hydraulic
fracture and the porous material in their formulation. They combine the FEM for the resolu-
tion of the poroelasticity equations in the porous medium with an FD approximation of the flow
equations along the fracture. Simoni and Secchi [16] and Secchi et al. [17] present a coupled
model based on the use of a mesh generator that updates the domain and remeshes according
to fracture mechanics principles as the fractures nucleate and propagate. Poroelasticity is formu-
lated in the continuum FEs, and the exchange of fluid between the fracture and the surrounding
porous material is considered. Slowik and Saouma [18] propose a model for water flow along a
propagating crack which, combined with a fictitious crack model with a bilinear softening curve
[19], reproduces a series of experimental results obtained from fluid pressurized wedge-splitting
tests. In this case, the undamaged material is considered to be impermeable compared to the
developing crack, for which a hydraulic conductivity dependent on the aperture is adjusted to
match the experimental results. The transient diffusion model includes the interaction between the
penetrating fluid and the air in both the initially dry medium and the propagating fracture. More
recently a coupled HM model for hydraulic fracture simulation using the FD method has been also
proposed [20].
Noorishad et al. [21] and Ng and Small [22] formulate a totally coupled ‘zero-thickness’
interface model for pre-existing discontinuities. Ng and Small’s model is based on the Coulomb
criterion for the mechanical behaviour and on Darcy’s law for the flow behaviour. Although using
double-nodded joint elements, none of the models considers any hydraulic potential drop between
the two interface walls; that is they assume that the fluid pressure in two nodes at opposite sides
of the discontinuity (that in analysis often share the same coordinates) is always equal. In this
case poroelasticity theory is considered for the porous medium, as in the work of Selvadurai
and Nguyen [23], who present a joint formulated by assemblage of some traditional models

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
2086 J. M. SEGURA AND I. CAROL

for interface elements including the influence of asperity damage on the discontinuity hydraulic
conductivity. The models proposed by Baldoni and Millard [24], and Guiducci et al. [25, 26] are
totally coupled and they use triple-node ‘zero-thickness’ interface elements to reproduce the HM
behaviour of pre-existing discontinuities. Baldoni and Millard [24] consider a linear elastic joint
with constant conductivity, and the formulation in [25, 26] combines the stochastic cubic law for
the flow behaviour, with a non-linear hyperbolic function [27] that describes the typical stress-
closure relationship of pre-existing joints under confining stresses. The special ‘zero-thickness’
interface element with three nodes makes the formulation of the model a little bit special; to the
classical nodes defining the boundaries of the joint, with two (for two-dimensional, or three for
three-dimensional) degrees of freedom corresponding to the mechanical problem (displacements)
and one to the hydraulic problem (fluid pressure), one has to consider also the inner nodes of the
element that will represent the fluid flow along the discontinuity and that will have just one degree
of freedom corresponding to the fluid pressure at the centre of the fluid channel.
In the above-described context, the present paper describes a discrete crack/interface FEM-
based model for the HM coupled behaviour of discontinuities provided the use of zero-thickness
interface elements with double nodes. It combines well-established mechanical formulations for
either pre-existing discontinuities [28, 29] or developing cracks [30–32] with a more recently
proposed model for diffusion analyses that introduces a transversal potential drop across the
discontinuity [33]. Non-linearity is restricted to the discontinuities and poroelasticity is assumed
in the continuum. A fully coupled and a staggered procedures are developed, the latter through
partition of the HM coupled equations, which becomes a convenient solution strategy since the same
zero-thickness interface geometry (i.e. the same FE mesh) is used for both flow and mechanical
problems.
After this Introduction, the contents of the paper are organized as follows: Section 2 briefly shows
the FEM-based HM formulation of saturated porous materials. Section 3 contains the equations
governing the HM coupled behaviour of discontinuities, and a detailed description of its FEM
formulation using zero-thickness interface elements. In Section 4 two different numerical strategies
are presented to solve the system of coupled equations reached as an assembly of FEM equations
in Sections 2 and 3. Finally, some concluding remarks of the formulation presented are given in
Section 5. Details of the numerical implementation as well as some examples of application are
given in the Part II companion paper [34].

2. HM COUPLING IN SATURATED POROUS MEDIA

The problem of fluid flow in deforming porous media is extensively treated in literature (e.g.
[5, 35, 36]) and the corresponding governing equations and its FEM discretization are briefly
summarized here. The present formulation assumes that tension is positive and compression nega-
tive (as usual in continuum mechanics analyses), small-strain theory, isothermal equilibrium and
negligible inertial forces. The displacement of the solid phase and the fluid pressure are considered
as the main unknowns.
The mechanical behaviour of a saturated porous material is described combining the linear
momentum balance equation for the mixture solid/water, the general form of the effective stress
principle, the constitutive relationship for solid phase relating effective stresses to strains, and the
compatibility equation that links strains to displacements. The mass balance equation for water

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2087

is combined with the general form of Darcy’s law to describe the flow behaviour of the porous
medium under the influence of the solid skeleton deformation.
Following standard FEM procedures [3, 5, 35, 37], the following system of equations is reached
which describes the HM behaviour of a standard porous medium element of volume :

BTc r d−Lc pc −Fuc = fuc (1)


duc dpc p p
Ec pc +LTc +Sc −Fc = fc (2)
dt dt
where uc and pc are the nodal element displacement and fluid pressure vectors, Bc is the strain
p p
operator, Fuc and Fc are the right-hand side vectors of force and flux, fuc and fc are the force and flux
vectors, respectively, at nodes shared with neighbouring elements, and the remaining matrices are

p
Lc = BTc c mc Nc d is the coupling matrix (3)


p Kcm p
Ec = (∇Nc )T ∇Nc d is the permeability matrix (4)
 f

pT 1 p
Sc = Nc Nc d is the compressibility matrix (5)
 Mc

In those expressions, mc is the equivalent to the identity tensor in the present vector formulation,
c is Biot–Willis’ coefficient, Mc is Biot’s modulus,  f is the fluid specific weight, Kcm is the
p
permeability matrix, and Nc is the matrix containing the shape functions that interpolate the fluid
pressure field within the FE.
Note that in Equation (1), for linear
 elasticity one may replace r = Dc e, and e = Bc u c , leading to
the traditional stiffness matrix Kc =  Bc Dc Bc d. However, in the present development the inte-
T

gral expression is preferred because it is more general, allowing inclusion of non-linear behaviour
as in the case of interface elements in Section 3.

3. HM COUPLING IN DISCONTINUITIES

The HM formulation of discontinuities is much less common in literature than the studies devoted
to the coupled analysis of continuum porous media. This section analyses more in detail the HM
governing equations and their discretization via zero-thickness interface elements with double nodes
using the FEM, which leads to a system of coupled equations in terms of the displacements and
fluid pressure at the interface nodes which structure is analogous to the popular ‘u– p’ formulation
of porous medium elements schematically introduced in previous Section 2.
In the case of HM coupling in discontinuities, flow through the discontinuity is influenced by
its deformation through changes in the permeability and the storage capacity. At the same time,
changes in fluid pressure result in changes in the effective stress distribution, and the subsequent
closure, opening or even propagation of the discontinuity. In this way, the coupling loop is closed,
which requires considering all these aspects simultaneously.

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
2088 J. M. SEGURA AND I. CAROL

Figure 1. Stresses and relative displacements at the mid-plane of a differential discontinuity element.

3.1. Governing equations


A differential section of a discontinuity is represented in Figure 1. The discontinuity is defined
by the top and bottom surfaces (or walls), and the mid-plane or axis surface. In a zero-thickness
FE-based formulation, the governing equations are first developed at the mid-plane and then
extended to the actual walls, and it is generally assumed that the three surfaces coincide (same
location). The coupled formulation is decomposed into two parts: the mechanical formulation of
a discontinuity under the influence of a fluid pressure distribution and the flow formulation of a
deformable discontinuity. The formulation is presented for the particular case of a one-dimensional
discontinuity (embedded in a two-dimensional medium), although the procedure is completely
general and its application to a two-dimensional discontinuity embedded in a three-dimensional
continuum is straightforward.

3.1.1. Mechanical formulation. Stresses and relative displacements are defined at the discontinuity
mid-plane. In the particular case of a one-dimensional discontinuity both vectors are composed of
normal and tangential components, as also represented in Figure 1.
The principle of effective stresses applied to the particular case of discontinuities considers that
the fluid pressure influences the normal stress component (i.e. as the volumetric component in
the continuum formulation), and that it is the effective stress the one that promotes the relative
movement of the discontinuity walls:

rm = rm + j mm pm (6)

where rm = [  ]T and rm = [ ]T are the mid-plane effective and total stress vectors, respec-
tively, mm = [0 1]T is a vector that introduces the influence of fluid pressure in the direction normal
to the discontinuity axis, pm is the fluid pressure at the mid-plane of the discontinuity and  j is
the discontinuity Biot’s coefficient [21] further discussed in the following subsection.
The constitutive equation links the effective stresses in the mid-plane of the discontinuity to the
relative displacements:

drm = Dm dAm (7)

where Am = [Al An ]T is the relative displacement vector, with Al the discontinuity sliding and
An its aperture, and Dm is the incremental constitutive relationship matrix, whose expression
depends on the type of discontinuity being analysed: in the case of pre-existing discontinuities
(most geological discontinuities such as joints, faults, etc.) the constitutive relationship should be

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2089

based on rock mechanics theory [28, 29], whereas in the case of developing cracks the constitutive
model will be based on fracture mechanics theory [30–32].

3.1.2. Fluid flow formulation. The fluid flow formulation of a discontinuity is developed following
Segura and Carol [33] specifications and taking the fluid pressure as the main unknown. Therefore,
differentiation is made between a fluid pressure gradient in the longitudinal direction and a fluid
pressure drop transversal to the discontinuity.
The formulation of the longitudinal flow combines the fluid mass balance equation with the
general form of Darcy’s law in the mid-plane. The mass balance equation for the differential
element in Figure 2 is
*Q l 1 * pm *An
+ + j =0 (8)
*l M j *t *t
where Q l is the local longitudinal flow rate, l is a longitudinal local coordinate, and t stands for time.
Parameters  j and M j are the discontinuity Biot’s coefficient and Biot’s modulus, respectively,
and they are understood in an analogous way as in the case of porous materials. These parameters
may depend on joint properties such as the roughness of fracture surfaces, the state of stress or
the presence of infill material [21], but for a clean discontinuity one can assume  j equal to
unity and M j equal to the inverse fluid compressibility multiplied by the discontinuity aperture
to be reasonable estimates. Very few experiments exist to determine reasonable values of these
parameters [38].
The generalized Darcy’s law is assumed to relate the flow rate to the fluid pressure gradient in
the longitudinal direction:
 
1 * pm *z m
Q l = −Tl + (9)
 f *l *l
where Tl is the longitudinal transmissivity,  f is the fluid specific weight, and z m is the mid-plane
elevation. According to the cubic law [40], the longitudinal transmissivity is
g 3
Tl = A (10)
12 n
where g is the gravity acceleration, and  is the fluid kinematic viscosity.

Figure 2. Fluid flow along and across a differential discontinuity element and detail
of the transversal fluid pressure drop.

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
2090 J. M. SEGURA AND I. CAROL

The foregoing assumption (9) perhaps deserves some additional comments. Physically, the two
surfaces of a geomechanical discontinuity in general will only be in contact in some areas and
separated in others due to roughness, forming an interconnected network of void space along the
mid-plane through which fluid flows. The consideration of the detailed network geometry, together
with the associated thickness variations and opening/closing effects due to normal stresses across
the discontinuity, would lead in principle to a full flow analysis analogous to that of a 2D porous
medium [40]. For practical purposes, however, at the level of observation we are interested in, the
flow along geomechanical discontinuities is generally idealized as a laminar flow between smooth
parallel plates separated by a certain ‘hydraulic’ aperture. Under those conditions, the flow may be
assumed proportional to the longitudinal fluid energy gradient by the cube of the aperture according
to the so-called cubic law [39]. The cubic law has been widely studied both experimentally and
numerically, showing its validity range for each situation being analysed [41–45].
A discontinuity may also represent an obstacle for the transversal fluid flow due to, for instance,
the presence of infill material. In fact, this can be observed in many sealing faults found in
petroleum reservoirs that act as impermeable barriers. Consequently, a transversal fluid pressure

drop p mp between the discontinuity boundaries is considered and related to a transversal flux by
a transversal conductivity coefficient K t :

qt = K t ( p bot − p top ) = K t p t (11)

where the superscripts bot and top stand for bottom and top discontinuity walls. One can estimate
reasonable values for the transversal conductivity coefficient (and also for the longitudinal one)
considering an equivalent continuum medium width as suggested in [46] and in the application
example in [33].

3.2. FEM formulation of the zero-thickness interface element


The FEM formulation is developed for a one-dimensional linear zero-thickness interface element
(i.e. with four nodes, two at each boundary as in Figure 3), although an analogous procedure would
be valid for a quadratic or higher order interface elements. Some ‘mid-points’ (MPs) located on
the mid-plane of the joint in between each pair of real nodes are considered, as well as a local
orthogonal coordinate system on the mid-plane that is related to the global coordinate system
through a rotation matrix r that transforms the relative displacements vector in the local base
(Am ) to the same vector expressed in the global coordinate system (am ). For the one-dimensional

Figure 3. Zero-thickness interface element geometry.

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2091

interface element in Figure 3, the rotation matrix r is


    
Al cos  sin  ax
Am = = = ram (12)
An − sin  cos  a y

3.2.1. Mechanical formulation. According to the FEM, the main variables defined in the mid-plane
(am and pm ) are interpolated through their values at the mid-nodes:
⎡ ⎤
am1x
    ⎢ ⎥
ax (x, y, t) u
Nm1 0 u
Nm2 0 ⎢ am1 y ⎥
am (x, y, t) = ≈ ⎢ ⎥ = Nu amn (13)
⎢ ⎥ m
a y (x, y, t) 0 u
Nm1 0 u
Nm2 (x,y) ⎣ a m2 x⎦

am2 y t
 
p p pm1 p
pm (x, y, t) ≈ [Nm1 Nm2 ](x,y) = Nm pmn (14)
pm2 t

The relative displacement (am ) is the difference of displacements between the interface boundaries.
In particular, the relative displacement at the mid-nodes amn can be expressed in terms of the
displacements at the real nodes in the form of the following compatibility equation:
⎡ ⎤ ⎡ ⎤
am1x u 3x −u 1x
⎢ ⎥ ⎢ ⎥
⎢ am1 y ⎥ ⎢ u 3 y −u 1 y ⎥

amn = ⎢ ⎥ ⎢ ⎥ = [−I4 I4 ]uj
⎥=⎢ ⎥ (15)
⎣ am2x ⎦ ⎣ u 4x −u 2x ⎦
am2 y u 4 y −u 2 y

where I4 is the (4×4) identity matrix, and uj is the nodal displacement vector for the joint element.
It is assumed that the fluid pressure at the mid-plane of the joint is the average of the fluid
pressure at the boundaries [33], which is used to simply express the fluid pressure at the mid-nodes
in terms of the fluid pressure at the element nodes as
⎡ p +p ⎤
  1 3
pm1 ⎢ 2 ⎥ 1
pmn = =⎣ ⎦ = [I2 I2 ]pj (16)
pm2 p2 + p4 2
2
where I2 is the (2×2) identity matrix and pj is the joint nodal fluid pressure vector.
The discretized equation governing the mechanical behaviour of the joint is obtained through
application of the principle of virtual work (PVW). A virtual displacement of the interface element
nodes uj and the corresponding relative displacements developed on the mid-plane of the joint
Am are considered:

u Tj fju = ATm rm dl ∀u Tj (17)
lm

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
2092 J. M. SEGURA AND I. CAROL

where fju and rm are the induced nodal forces and mid-plane stresses, respectively. Introduction of
Equations (6) and (7), together with Equations (12)–(16) leads to the following equation entirely
expressed in terms of the displacements and fluid pressure values at the interface nodes:

fj =
u
BTj rm dl −Lj pj (18)
lm

where
 
−BTm
BTj = is the interface element B-matrix (19)
BTm
 
1 −Lm −Lm
Lj = is the interface element coupling matrix (20)
2 Lm Lm

and

p
Lm = BTm  j mm Nm dl is the mid-plane coupling matrix (21)
lm

BTm = Num T rT is the mid-point B-matrix (22)

3.2.2. Flow formulation. To formulate the joint from the diffusion point of view, the relative
displacements and the fluid pressure in the mid-plane are interpolated according to expressions
(13) and (14), and the fluid pressure drop, also in the mid-plane, is interpolated as
⎡ ⎤
 p m1 p
Nm2 ](x,y) ⎣  ⎦ = Nm p mn
p p
p m (x, y, t) ≈ [Nm1 (23)
p m2
t

The fluid pressure drop is the difference of fluid pressure between the interface boundaries as
defined in Equation (11), which for the particular case of the mid-plane nodes gives the following
compatibility equation analogous to Equation (15) for the displacements:
⎡ ⎤  
 p m1 p1 − p3
p mn = ⎣ ⎦= = [I2 −I2 ]pj (24)

p m2 p2 − p4

The analogy between the mechanical and the flow problems [33] is used to apply the PVW to the
flow behaviour of the discontinuity:
  T  
p T * pm 1 * pm *An
pTj fj =  p m qt dl +  (−Q l ) dl +  pm
T
+ j dl ∀pj (25)
lm lm *l lm M j *t *t
p
where fj , qt and Q l are the induced nodal flows and mid-plane transversal and longitudinal
fluxes, respectively. Introducing Equations (8), (9) and (11), together with Equations (12)–(16)

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2093

and (23)–(24), the following equation is achieved:


p dpj duj
fj = Ej pj +Sj +LTj +Qz (26)
dt dt
where
Ej = ELj +ETj is the interface element conductivity matrix (27)
 
1 ELm ELm
ELj = is the interface element longitudinal conductivity matrix (28)
4 ELm ELm
 
ETm −ETm
ETj = is the interface element transversal conductivity matrix (29)
−ETm ETm
 
1 Sm Sm
Sj = is the interface element capacity or storage matrix (30)
4 Sm Sm
 
1 −Lm Lm
T T
Lj =
T
is the transposed of the interface element coupling matrix (31)
2 −LTm LTm
 
1 QZm
Qz = introduces the effect of gravity potential on the flow system (32)
2 QZm

and
 pT p
*Nm Tl *Nm
ELm = dl is the mid-plane longitudinal conductivity matrix (33)
lm *l  f *l

pT p
ETm = Nm K̃ t Nm dl is the mid-plane transversal conductivity matrix (34)
lm

pT 1 p
Sm = Nm Nm dl is the mid-plane capacity matrix (35)
lm Mj

pT
LTm = Nm mTm  j Bm dl is the transposed of the mid-plane coupling matrix (36)
lm
 pT
*Nm *z m
QZm = Tl dl represents the effect of the gravity potential
lm *l *l
at the mid-plane (37)
Note that the final matrices (28)–(32) for the zero-thickness interface element are obtained as
a function of the matrices (33)–(37) computed at the mid-plane (Figure 3). This is due to the
extension of the formulation from the mid-plane to the actual zero-thickness interface nodes.
If the mechanical influence on the flow problem is omitted, the original flow formulation in [33]
is recovered. Here, the flow formulation has been obtained purely applying the PVW, whereas in
[33] it was obtained using balance equations at the FE nodes and considering the flow distributions

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
2094 J. M. SEGURA AND I. CAROL

and equilibrium within the joint. The formulation in [33] shows in more detail how matrix (28)
contains off-diagonal terms due to the main assumption of considering the fluid pressure and the
fluid flow at the mid-plane as the average of the corresponding values at the interface nodes. On
the other hand, matrix (29) off-diagonal terms naturally appear by considering the potential drop
between the two interfaces and also the counterpart assumption on the transversal flow.

3.2.3. Resulting ‘u– p’ formulation of the zero-thickness interface element with double nodes.
Collection of Equations (18) and (26) results in the following system of equations for the HM
coupled formulation of a zero-thickness interface element:

BTj rm  dl −Lj pj = fju (38)
lm

duj dpj p
Ej pj +LTj +Sj +Qz = fj (39)
dt dt
Equation (38) has been left in integral form, which allows the formulation to be specified for
non-linear behaviour as done in Section 4.

3.3. FEM formulation of discontinuous porous media


After assembly of the continuum (Equations (1)–(2)) and zero-thickness interface elements
(Equations (38)–(39)), the forces and flows at sharing nodes cancel and a single system of
equations in terms of the displacements and fluid pressure at the nodes is reached for the boundary
value problem (BVP) under analysis:

BT r −Lp = Fu (40)

du dp
Ep+LT +S = Fp (41)
dt dt
The generalized trapezoidal rule [37] is considered to discretize Equations (40) and (41) in time.
If the problem is to be solved between an initial time t0 and a final time t f , a partition of the
time interval [t0 , t f ] is considered, which results in a series of time increments tn+1 = tn+1 −tn . If
Equations (40) and (41) are evaluated at time n +, the following system of equations is reached:

BT rn+ −Lpn+1 = F̂un+ (42)

p
LT u n+1 +(Sn+ +tn+1 En+ )pn+1 = F̂n+ (43)
where the increment of nodal displacements un+1 = un+1 −un and fluid pressure pn+1 = pn+1 −
pn are the main unknowns:
F̂un+ = Fun+ +Lpn (44)
p p
F̂n+ = tn+1 (Fn+ −En+ pn ) (45)
The value of  (01) defines the time-marching approximation and stability conditions of the
integration procedure [5, 37].

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2095

4. NUMERICAL STRATEGIES

This section describes two different approaches to solve the system of coupled equations (42)–
(43) reached in Section 3. One approach is the staggered procedure (sometimes called partially
coupled), which makes use of two separate flow and mechanical computer codes, appropriately
linked. The other strategy is called monolithic (or sometimes also fully coupled) and it directly
solves the whole system of equations by means of a single, integrated computer code. In this case,
the overall non-linear system of equations is solved using a Newton–Raphson iterative method
with non-symmetric Jacobian matrix.
From a terminological perspective, the terms ‘fully coupled’ or ‘partially coupled’ should not
lead to misunderstanding; in this context they refer to the numerical strategy. In this way, the
monolithic procedure is a fully coupled approach in the sense that the full system of equations
is solved simultaneously. And only in this sense the staggered procedure may be considered
a ‘partially coupled approach’, because from the point of view of the system of equations the
staggered procedure is also solving the full system and it should eventually lead to the actual fully
coupled solution, that upon iteration to full convergence should be the same as the one obtained
with the monolithic procedure.
The implementation work has been based on the FE codes available within the group of
Mechanics of Materials at the Department of Geotechnical Engineering UPC (School of Civil
Engineering—Barcelona): DRAC is the main general purpose code within the group for mechanical
analysis, with some special features for rock mechanics including zero-thickness interface elements
[47]. DRACFLOW is a more recent FE code for diffusion-driven and coupled problems (heat,
moisture or water) also including double-node interface elements for diffusion analysis.

4.1. Staggered approach


As already mentioned, the staggered procedure makes use of the two separate codes for the
mechanical and flow problems, which are then appropriately linked through the corresponding
coupling loops. The procedure iterates from one code to the other until a certain tolerance on the
solution is achieved. In fact, the intensity of the coupling can be measured through the number
of iterations needed for convergence. It is an advantageous procedure in problems where the
coupling between the two coupled fields is weak, and therefore the number of iterations required
for convergence is low. However, this is not always the case in HM coupled problems in porous
materials, and even less in the analysis of fractured media due to the strong dependency of the joint
permeability on its aperture that may indeed be the cause of convergence problems in staggered
calculations.
If the displacement increment at the nodes is assumed to be known and tn+1 is the time-step
under analysis, the flow equation (43) can be reorganized to obtain the classical equation solved
by flow FE codes:
p
(Sn+ +tn+1 En+ )pn+1 = F̂n+ −LT un+1 (46)

in which, besides the influence on the permeability matrix, the deformation of the system also
influences the right-hand side vector in the sense of inducing additional fluid supply (i.e. source
term). Moving on to the mechanical equation (42), this can be reorganized if the fluid pressure
at nodes are assumed to be known in advance, in order to obtain the mechanical equation typically

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
2096 J. M. SEGURA AND I. CAROL

solved by geomechanical codes:



BT rn+ = F̂un+ +Lpn+1 (47)

in which the fluid pressure distribution generates an additional nodal force distribution that affects
the right-hand side vector.
The solution of the coupled system is obtained by iteratively solving Equations (46) and (47)
with the fluid flow and the geomechanical simulator, respectively. This iterative process starts with
an initial estimate for the displacements (un+1 )0 that affects the permeability matrix and the right-
hand side vector of Equation (46), the solution of which gives an estimate for the pressure solution
(pn+1 )1 . This pressure increment is introduced in the mechanical equation (47), which, after
solution with the mechanical code, leads to a new estimate of the nodal displacements (un+1 )1
that can be introduced back into Equation (46) to solve the flow problem again. This sequential
procedure between both codes is repeated until the results obey a certain tolerance as schematically
shown in Figure 4.
Both codes DRAC and DRACFLOW incorporate the same interface element geometry with
double nodes (as formulated in Section 3). This allows using the same FE mesh for both problems,
with the subsequent advantages in the exchange of information handled by a master code that also
checks the convergence of the procedure. Due to the cubic dependence between the discontinuity
permeability and its aperture, it is expected that the coupling in the HM analysis of fractured or
fracturing porous media will be quite intense and therefore the number of iterations necessary
for convergence will be high, and convergence problems may occur. In the implementation and

Figure 4. Staggered procedure scheme for the transient case.

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2097

verification of the staggered procedure, it has been observed that the standard staggered strategy
of simply using the solution obtained by one code to calculate vectors and matrices needed for the
other, does not always lead to convergence, but the iterative process sometimes ends up oscillating
between two pairs of solution sets. In order to avoid these oscillations one alternative strategy is to
use instead a combination of the solutions obtained during the current and the previous iterations.
According to this, the following expression is considered for the tentative solution to be used in
the iteration k +1 of the staggered procedure:

k
xk+1 = (1− )k+1 x0 + (1− )i x̂k+1−i (48)
i=0

where 0 1, x0 is an initial estimate, x̂ j is the solution given by the codes in iteration j, and
xk+1 is the estimate introduced in the staggered procedure for iteration k +1. In the particular case
of considering = 1, the codes will directly use the last solution obtained by the other code. On
the other hand, considering = 0 no iteration would be made between both codes and the first
estimate at each instant would be directly the solution (i.e. explicit iterative coupling technique).
Different values of for the nodal displacements ( u ) and for the nodal pressure ( p ) can be used.
Saetta et al. [48] cite different predictors that can be obtained as combination of previous solutions
during the iterative procedure.

4.2. Fully coupled approach


The monolithic or simultaneous solution scheme is the most straightforward procedure numerically
speaking, and it directly solves the whole system of equations. The displacements and fluid pressure
at the nodes are obtained simultaneously at the end of each time-step.
As already described in Section 3, the transient problem is solved at a series of consecutive
time steps. If tn+1 is the time-step under analysis, the main unknowns will be the increment
of displacements and fluid pressure at nodes, un+1 and pn+1 , and the non-linear system of
equations to be solved (42)–(43) may be expressed through an operator w as a function of the
unknowns vector xTn+1 = [un+1 pn+1 ]:
⎡  ⎤
 u  T 
wn+ (xn+1 ) ⎢ B r n+ −Lp n+1 − F̂ u
n+ ⎥
wn+ (xn+1 ) = p =⎣ ⎦=0 (49)
wn+ (xn+1 ) p
LT un+1 +(Sn+ +tn+1 En+ )pn+1 − F̂n+

The Newton–Raphson method is the algorithm used to find the root of the operator wn+ , which
is equivalent to finding the solution of Equations (42)–(43), leading to the following system:

Jn+ (xkn+1 )xk+1


n+1 = −wn+ (xn+1 )
k
(50)

where xk+1 k k+1


n+1 = xn+1 +xn+1 , and Jn+ (xn+1 ) = [*wn+ /*x]xk
k is the Jacobian matrix:
n+1
⎡ ⎤
*wun+ *wun+
⎢ *u *p ⎥
⎢ ⎥
Jn+ (xkn+1 ) = ⎢ p p ⎥ (51)
⎣ *w *w ⎦
n+ n+
*u *p xkn+1

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
2098 J. M. SEGURA AND I. CAROL

Therefore, provided the use of an initial estimate (x0n+1 ), the root-finding algorithm consists
of solving Equation (50) until a certain tolerance is attained in xk+1 k+1
n+1 = xn+1 +xn+1 and/or
k

wn+ (xn+1 ).
k
p
Taking into account Equation (49) for wun+ and wn+ , Equation (44) for F̂u , and Equation (45)
for F̂p , the partial derivatives obtained in the calculation of the Jacobian matrix depend on whether
they involve a continuum medium or a zero-thickness interface element.
Poroelasticity is assumed in the continuum elements, for which the usual Jacobian matrix in
standard ‘u– p’ formulations is obtained:
 
Kc −Lc
Jn+ (xn+1 ) =
k
(52)
LTc Scn+ +tn+1 Ecn+ xk
n+1

that becomes symmetric if the upper equation is divided by (−).


Dependency between the hydraulic conductivity of discontinuities and their aperture introduces
non-symmetry in the Jacobian matrix for the zero-thickness interface elements. If the cubic law
(Equation (10)) is considered:
 
Juu1 −Jup1
Jn+ (xn+1 ) =
k
(53)
Jpu1 +Jpu2 −Jpu3 Jpp1 +Jpp2 xk
n+1

where
 
*
Juu1 = BTj rm n+ dl = KjT (54)
*u lm

*
Jup1 = (Lj pn+1 ) = Lj (55)
*p
*
Jpu1 = (LT un+1 ) = LTj (56)
*u j
*
Jpu2 = (tn+1 Ejn+ [pn +pn+1 ])
*u
 I  p T T  p
1 2 *Nm ln+ *Nm
= tn+1 [I2 I2 ][pn +pn+1 ]mTj Bj dl (57)
4 lm I2 *l  f *l
 I  p T T 
* p 1 2 *Nm ln+ *z m T
Jpu3 = (tn+1 Fn+ ) = tn+1 m Bj dl (58)
*u 2 lm I2 *l  f *l j

*
Jpp1 = (Sjn+ pn+1 ) = Sjn+ (59)
*p
*
Jpp2 = (tn+1 Ejn+ pn+1 ) = tn+1 Ejn+ (60)
*p

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2099

and
f
Tln+ = A2n (61)
4 f

The Jacobian matrix is non-symmetric if all the previous sub-matrices are taken into account.
However, one can use an approximation that results in a symmetric matrix with the subsequent
economy in computer memory, which is compensated with an expected higher number of iterations
until convergence. If Jpu2 and Jpu3 are neglected and the upper equation is divided by (−), a
symmetric approximation to the Jacobian matrix analogous to that for the continuum elements (52)
may be obtained, which would correspond to the actual Jacobian matrix for interface elements
with constant transmissivity.

5. CONCLUDING REMARKS

The formulation presented for HM coupled problems in geomaterials with discontinuities has
several advantages. One is that spatial discretization is made using the same technique (the FEM)
for both continuum and discontinuities, as opposite to traditional approaches that combined the
FEM for mechanical behaviour with finite differences for the fluid flow. Also, among the various
possibilities, the zero-thickness interface elements for flow have been formulated using the double-
node formulation proposed recently by the same authors [33], which makes it possible to use the
same mesh for both problems, mechanical and flow, while still considering a transversal potential
drop across the discontinuity. From the viewpoint of the scheme for the solution of the HM
coupling, the two main options available, staggered and monolithic, have been described in a
unified context and notation, and have been both implemented within the same computer code
environment. In both cases, the iterative strategies implemented have been discussed, including
the Jacobian structure in the monolithic Newton–Raphson scheme. All these developments set the
scene for verification and comparison examples which are presented in the companion paper [34].

ACKNOWLEDGEMENTS
The first author wishes to acknowledge the FI doctoral fellowship from AGAUR-DURSI (Generalitat de
Catalunya, Barcelona), and support from ETSECCPB and UPC for completing his doctoral thesis. Partial
support was also obtained from grant MAT2005-24522 and research project BIA2006-12717, funded by
MEC (Madrid), as well as grant 80015/A04 funded by MFOM (Madrid). Early stages of this research
were also supported by Schlumberger.

REFERENCES
1. Terzaghi K. Erdbaumechanik auf Bodenphysikalischer Grundlage (Soil Mechanics based on Soil Physics). Franz
Deuticke: Leipzig, 1925.
2. Biot MA. A general theory of three dimensional consolidation. Journal of Applied Physics 1941; 12:155–164.
3. Zienkiewicz OC, Humpheson C, Lewis RW. A unified approach to soil mechanics problems (including plasticity
and visco-plasticity). In Finite Elements in Geomechanics, Gudehus G (ed.). Wiley: New York, 1977; 151–179.
4. Settari A, Walters DA. Advances in coupled geomechanical and reservoir modeling with applications to reservoir
compaction. SPE 51927. 1999 SPE Reservoir Simulation Symposium, Houston, TX, U.S.A., 1999.
5. Lewis RW, Schrefler BA. The Finite Element Method in the Static and Dynamic Deformation and Consolidation
of Porous Media. Wiley: Chichester, 1998.

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
2100 J. M. SEGURA AND I. CAROL

6. Chin LY, Thomas LK, Sylte JE, Pierson RG. Iterative coupled analysis of geomechanics and fluid flow for rock
compaction in reservoir simulation. Oil and Gas Science and Technology 2002; 57(5):485–497.
7. Mainguy M, Longuemare P, Wozniak JM. Coupled reservoir and geomechanics simulations to assess the risk of
well plug failure after reservoir abandonment. In Poromechanics III—Biot Centennial (1905–2005), Proceedings
of the 3rd Biot Conference on Poromechanics, Abouleisman YN, Cheng AH-D, Ulm FJ (eds). Taylor & Francis,
University of Oklahoma: OK, U.S.A., 2005; 647–652.
8. Liu J, Shen JC, Zhu WC, Liu JX. Numerical investigation of coupled multiphysics for enhanced oil recovery. In
Poromechanics III—Biot Centennial (1905–2005), Proceedings of the 3rd Biot Conference on Poromechanics,
Abouleisman YN, Cheng AH-D, Ulm FJ (eds). Taylor & Francis, University of Oklahoma: OK, U.S.A., 2005;
359–364.
9. Duarte Azevedo IC, Vaz LE, Vargas EA. A numerical procedure for the analysis of the hydromechanical coupling
in fractured rock masses. International Journal for Numerical and Analytical Methods in Geomechanics 1998;
22:867–901.
10. Sánchez M. Thermal-hydro-mechanical coupled analysis in low permeability media. Ph.D. Thesis, Technical
University of Catalonia (UPC), Barcelona, Spain, 2004.
11. Nair R, Ekbote S, Mody FK. Multiphase dual-porosity poromechanics. In Poromechanics III—Biot Centennial
(1905–2005), Proceedings of the 3rd Biot Conference on Poromechanics, Abouleisman YN, Cheng AH-D, Ulm
FJ (eds). Taylor & Francis, University of Oklahoma: OK, U.S.A., 2005; 141–148.
12. Nguyen TS, Selvadurai APS. Coupled thermal–mechanical–hydrological behaviour of sparsely fractured rock:
implications for nuclear fuel waste disposal. International Journal of Rock Mechanics and Mining Science and
Geomechanics Abstracts 1995; 32(5):465–479.
13. Papanastasiou P. The influence of plasticity in hydraulic fracturing. International Journal of Fracture 1997;
84:61–79.
14. Boone TJ, Ingraffea AR. A numerical procedure for simulation of hydraulic-driven fracture propagation in
poroelastic media. International Journal for Numerical and Analytical Methods in Geomechanics 1990; 14:27–47.
15. Boone TJ, Ingraffea AR, Roegiers JC. Simulation of hydraulic fracture propagation in poroelastic rock with
application to stress measurement techniques. International Journal of Rock Mechanics and Mining Science and
Geomechanics Abstracts 1991; 28(1):1–14.
16. Simoni L, Secchi S. Cohesive fracture mechanics for a multi-phase porous medium. Engineering Computations:
International Journal for Computer-Aided Engineering 2003; 20(5):675–698.
17. Secchi S, Simoni L, Schrefler BA. Mesh adaptation and transfer schemes for discrete fracture propagation in porous
materials. International Journal for Numerical and Analytical Methods in Geomechanics 2007; 31:331–345.
18. Slowik V, Saouma VE. Water pressure in propagating concrete cracks. ASCE Journal of Structural Engineering
2000; 126(2):235–242.
19. Červenka J, Chandra JM, Saouma V. Mixed ode fracture of cementitious bimaterial interfaces: part II. Numerical
simulation. Engineering Fracture Mechanics 1998; 60:95–107.
20. Murdoch LC, Germanovich LN. Analysis of a deformable fracture in permeable material. International Journal
for Numerical and Analytical Methods in Geomechanics 2006; 30:529–561.
21. Noorishad J, Ayatollahi MS, Witherspoon PA. A finite-element method for coupled stress and fluid flow analysis
in fractured rock masses. International Journal of Rock Mechanics and Mining Science and Geomechanics
Abstracts 1982; 19:185–193.
22. Ng KLA, Small JC. Behavior of joints and interfaces subjected to water pressure. Computers and Geotechnics
1997; 20(1):71–93.
23. Selvadurai APS, Nguyen TS. Mechanics and fluid transport in a degradable discontinuity. Engineering Geology
1999; 53:243–249.
24. Baldoni F, Millard A. A finite element formulation for the coupled hydro-mechanical behaviour of porous rock
joints. In Poromechanics: A Tribute to Maurice A Biot. Proceedings of the 1st Biot Conference, Thimus JF,
Abousleiman Y, Cheng A, Coussy O, Detournay E (eds). Taylor & Francis: Louvain-la-Neuve, Belgium, 1998;
339–344.
25. Guiducci C, Pellegrino A, Radu JP, Collin F, Charlier R. Numerical modeling of hydro-mechanical fracture
behavior. In Numerical Models in Geomechanics-NUMOG VIII, Pande GN, Pietruszczak S (eds). Swets &
Zeitlinger: Lisse, 2002; 293–299.
26. Guiducci C, Pellegrino A, Radu JP, Collin F, Charlier R. Hydro-mechanical behavior of fractures: 2D, F.E.M.
modeling. In Poromechanics II: Proceedings of the 2nd Biot Conference on Poromechanics, Auriault JL, Geindreau
C, Royer P, Bloch GF (eds). Taylor & Francis: Grenoble, France, 2002; 217–224.

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag
COUPLED HM ANALYSIS USING ZERO-THICKNESS INTERFACE ELEMENTS 2101

27. Bart M. Contribution a la modlisation du comportement hydromcanique des massifs rocheux avec fracture. Ph.D.
Thesis, Univ. des Sciences et Technologies de Lille, Lille, France, 2000.
28. Carol I, Gens A, Alonso E. A three dimensional elastoplastic joint element. In Fundamentals of Rock Joints,
Stephansson O (ed.). Centek Publishers: Lulea, 1985; 441–451.
29. Gens A, Carol I, Alonso EE. A constitutive model for rock joints: formulation and numerical implementation.
Computers and Geotechnics 1990; 9:3–20.
30. Carol I, Prat P, López CM. A normal/shear cracking model. Application to discrete crack analysis. ASCE Journal
of Engineering Mechanics 1997; 123(8):765–773.
31. Carol I, López CM, Roa O. Micromechanical analysis of quasi-brittle materials using fracture-based interface
elements. International Journal for Numerical Methods in Engineering 2001; 52:193–215.
32. Caballero A, López CM, Carol I. 3D meso-structural analysis of concrete specimens under uniaxial tension.
Computer Methods in Applied Mechanics and Engineering 2006; 195(52):7182–7195.
33. Segura JM, Carol I. On zero-thickness interface elements for diffusion problems. International Journal for
Numerical and Analytical Methods in Geomechanics 2004; 28(9):947–962.
34. Segura JM, Carol I. Coupled HM analysis using zero-thickness interface elements with double nodes. Part II:
verification and application. International Journal for Numerical and Analytical Methods in Geomechanics 2008;
DOI: 10.1002/nag.730.
35. Zienkiewicz OC, Chan AHC, Pastor M, Paul DK, Shiomi T. Static and dynamic behaviour of soils: a rational
approach to quantitative solutions. I. Fully saturated problems. Proceedings of the Royal Society of London,
Series A—Mathematical Physical and Engineering Sciences 1990; 429(1877):285–309.
36. Coussy O. Mechanics of Porous Continua. Wiley: Chichester, 1995.
37. Zienkiewicz OC. El Método de los Elementos Finitos. Editiorial Reverté: Barcelona, 1980.
38. Rutqvist J, Stephansson O. The role of hydromechanical coupling in fractured rock engineering. Hydrogeology
Journal 2003; 11:7–40.
39. Snow D. A parallel plate model of fractured permeable media. Ph.D. Dissertation, University of California,
Berkeley, 1965.
40. National Research Council. Rock Fractures and Fluid Flow: Contemporary Understanding and Applications.
National Academy Press: Washington, DC, U.S.A., 1996.
41. Witherspoon PA, Wang JCY, Iway K, Gale JE. Validity of the cubic law for fluid flow in a deformable rock
fracture. Water Resources Research 1980; 16(6):1016–1024.
42. Tsang YW, Tsang CF. Channel model of flow through fractured media. Water Resources Research 1987;
23:467–479.
43. Brown SR. Fluid flow through rock joints: the effects of surface roughness. Journal of Geophysical Research
1987; 92(B):1337–1347.
44. Thompson ME, Brown SR. The effect of anisotropic surface roughness on flow and transport in fractures. Journal
of Geophysical Research 1991; 96(B):21923–21932.
45. Oron AP, Berkowitz B. Flow in rock fractures: the local cubic law assumption reexamined. Water Resources
Research 1998; 34:2811–2825.
46. Segura JM. Coupled HM analysis using zero-thickness interface elements with double nodes. Ph.D. Thesis,
Technical University of Catalonia (UPC), Barcelona, Spain, 2007.
47. Prat P, Gens A, Carol I, Ledesma A, Gili JA. DRAC: a computer software for the analysis of rock mechanics
problems. In International Symposium on Application of Computer Methods in Rock Mechanics, vol. 2, Liu H
(ed.), Xian, China, 1993; 1361–1368.
48. Saetta A, Schrefler BA, Vitaliani R. Estrategias de solución para análisis de consolidación acoplados en medios
porosos. Revista Internacional de Métodos Numéricos para Cálculo y Diseño en Ingenierı́a 1991; 7(1):55–66.

Copyright q 2008 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2008; 32:2083–2101
DOI: 10.1002/nag

You might also like