Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Computers and Geotechnics 136 (2021) 104111

Contents lists available at ScienceDirect

Computers and Geotechnics


journal homepage: www.elsevier.com/locate/compgeo

Research Paper

Analytical formulation for the deformability assessment of rock masses with


filled discontinuities
Noelia Esteban *, Rubén Galindo, Alcibíades Serrano
Universidad Politécnica de Madrid (UPM), ETSI Caminos, C. y P., C/ Profesor Aranguren s/n, Madrid 28040, Spain

A R T I C L E I N F O A B S T R A C T

Keywords: The estimation of the deformation modulus of a rock mass is a fundamental factor in the design of civil works
Deformability such as tunnels and dams. However, it is one of the parameters that entails highest difficulty when evaluating the
Rock mass rock mass. In this paper, a new analytical model is presented to assess rock mass deformability, with the aim of
Analytic model
improving the estimation of the properties and their implementation in numerical models. The model presented
Filled discontinuities
is analytical, anisotropic and based on the equivalent continuum approach, and considers: the geometry of the
Equivalent continuum model
rock mass for orientation, spacing and thickness of the discontinuities, the presence of fill; and, finally, the
deformational properties of the intact rock and the discontinuities. The resultant model was adapted for the
specific case of filled discontinuities using both the linear elastic model and the Hertz-Mindlin law. The
assessment presented includes: the implementation of the model, the analysis of various theoretical scenarios, the
estimation of deformabilities for multiple cases, a numerical validation using the Distinct Element Method, and a
discussion of the scale effect. The results are compared with some empirical formulations for the deformability
assessment of rock masses. This comparison underlines the importance of separately considering all the factors
that define the rock mass, as well as the necessity of heeding the anisotropy caused by the orientation of the
discontinuity sets.

1. Introduction discontinuities and the scale effect. On the other hand, field surveys are
suitable for the evaluation of the global behaviour of the rock mass,
Deformability, along with resistance, is one of the key aspects to although careful consideration must be given to the relationship between
comprehend the behaviour of rock masses and incorporate it into the the scale of the survey and the influence area of the problem studied.
design of civil infrastructures in rock. This is of the highest importance in Furthermore, field surveys have a high cost and their interpretation can be
the design of tunnels, underground excavations, slopes and foundations. complex (Bieniawski, 1978). Both laboratory tests and field surveys are
The latter has an important application in concrete dams supported by required for empirical formulations and analytic models. While empirical
rock mass foundations where a strict control of movement is necessary in formulations have the advantage of their simplicity, their use is often
the body of the dam. limited to the local area where the data was obtained for their fitting.
It is a complex process to reach a realistic estimation of the deform­ Where this does not happen, the results usually show very high scatter.
ability of a rock mass and so is the selection of a suitable assessment Another frequent problem of this kind of formulation is the inadequacy for
method. Broadly speaking, there are four ways of analysing deformability, certain fundamental features of the analysed rock masses. On the other
which can be combined to determine the deformation modulus of a rock side, analytic models are deeply conditioned by their initial hypothesis
mass. These are: laboratory tests, field surveys, empirical formulation; and and are usually limited to very specific and simple cases.
analytic models. The four methods are complementary and are useful for The complexity in determining the deformation modulus of the rock
different stages of the design. Their simultaneous use and contrast be­ mass is, fundamentally, caused by its discontinuous nature of rock
tween their results all help to create a better and accurate assessment. Due masses. The influence of the characteristics and structure of the dis­
to the size of the samples, these are suitable in laboratory tests to estimate continuities is significant, and their behaviour is superimposed on the
the properties of the intact rock, but they cannot assess the influence of the constitutive relation of the rock matrix, resulting in a complex global

* Corresponding author.
E-mail address: noelia.erivera@alumnos.upm.es (N. Esteban).

https://doi.org/10.1016/j.compgeo.2021.104111
Received 9 October 2020; Received in revised form 15 February 2021; Accepted 28 February 2021
Available online 26 May 2021
0266-352X/© 2021 Elsevier Ltd. All rights reserved.
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

behaviour. As a result, the deformability of the rock mass is dependent Vásárhelyi & Kovács (2017) argued that the evolution of the method­
on multiple factors: the orientation, spacing and aperture of disconti­ ologies used in the in situ and laboratory testing created a discrepancy in
nuities, and the deformational behaviour of these discontinuities and the the data used to fit different formulations which explains the deviations
rock matrix. The model used to assess the deformability must account of their results. For this reason, using these old formulae in current
for all these factors and their influence, by incorporating the deforma­ projects is not good enough for the task. The same thing happens when
tion properties of discontinuities and intact rock. Considering each applying empirical relations in projects where the available rock data
aspect separately is contrary to the way empirical formulations were was obtained with different tests than the ones used in the fit of the
traditionally used, where the deformation moduli depend on the global formula. In addition, it should be noted that the results of the empirical
weighted indexes such as RMR or Q. formulae are only averaged values of the modulus of deformation,
This paper is a presentation of a mathematical model and a complete meaning that they assume an isotropic behaviour of the rock mass.
formulation that enables considering separately the structure and While this can be valid for a first approach, rock masses are vastly known
behaviour parameters of each element that were part of the rock mass. to have anisotropic behaviour, which limits the applicability of these
The result was a simulation of the global behaviour of the rock mass that formulae.
accounts for its anisotropic nature as the result of the geometric Figs. 1-3 show the classification of the empirical formulae was done
configuration of the discontinuities. In addition to the theoretical based on the parameters on which they depended on. Although the
formulation, the following were also carried out: a detailed analysis of majority of those expressions depend on classification indexes such as
the specific case of filled discontinuities using both the linear elastic RMR or RQD, there are some authors who chose not to use them to assess
model and the Hertz-Mindlin law, a numerical validation using the the damage level of the rock mass. Instead, they use other parameters,
Distinct Element Method, and a discussion of the scale effect. Finally, a such as dynamic moduli (Clerici, 1993) or velocity of propagation of
comparative assessment of the model with empirical formulations that waves (Barton, 2007). Regarding the group of models dependent on
are broadly used by practitioners was also carried out. classification indexes, further classification can be made by dis­
tinguishing those that consider the deformation modulus of the intact
2. State of the art rock (ER ), and those that do not.
Some of the most used and indexed empirical formulae are sum­
Many authors have suggested various methodologies for the esti­ marised in Table 1, where the origin of the data used for the fit of the
mation of the deformation modulus of rock masses. Two traditional expressions is also included.
paths are set out: 1) the fit of empirical relationships and 2) the theo­ Based on the data gathered from in situ and laboratory testing, there
retical development of analytic models. In the following sections, a brief are various methodologies that can be used to fit the expressions. The
review of the reach and limitations of the mentioned types of formula­ statistical regression methods are the most widely used in the published
tions is set out. In addition, a chronologic summary of the deformability research (Bieniawski, 1978; Serafim & Pereira, 1983; or Grimstad &
models is shown in Figs. 1-3 (empirical formulations based on global Barton, 1993). More recently, some formulae were adjusted by means of
classification indexes) and Fig. 4 (analytic models, classified according Neuronal Networks Algorithms (ANN) (Sonmez et al., 2006; Alemdag
to their theoretical basis). et al., 2016; Nejati et al., 2013); genetic programming (Alemdag et al.,
2016; Beiki et al., 2010); fuzzy inference of neuro-fuzzy (NF) models
(Bashari et al., 2011; Alemdag et al., 2016; Kayabasi et al., 2003); nu­
2.1. Empirical formulations merical modelling with back-analysis in work sites with auscultation
instruments (Verman et al., 1997; Reddish et al., 1997) or semi-
There are a large number of empirical formulae, some of which are theoretical models like the one by Diederichs & Kaiser (1999), based
given in Figs. 1-3 that have been appearing since the development of the on the analogy between rock strata and structural beam elements.
classicisation indexes, RQD, RMR and Q that form the basis of empirical
formulations. The main advantage of these formulae is the simplicity of
their use; they are often based on the properties of the rock matrix, 2.2. Analytic models
which can be obtained by means of laboratory tests, and the classifica­
tion indexes. However, there are various limitations to the empirical The analytic models to assess the deformability of rock masses
formulae that have to be considered. Their local area of application is began to appear at the same time as the empirical formulae (Duncan &
limited, as a consequence of the use of databases of specific rock masses Goodman, 1968). Similar to empirical expressions, there are numerous
to fit the equations. But their validity is conditioned by the quality of the analytic models and they have been appearing quite regularly. Fig. 4
data, including the equipment, test procedures and interpretation shows some analytic formulae that have been published up to the
methodologies used for the obtention of the data to fit the models. present. Their main advantage over the empirical expressions is that

Fig. 1. Empirical formulae of deformability of rock masses based on RQD (Rock Quality Designation) and the deformation modulus of the intact rock (ER). (See
above-mentioned references for further information.)

2
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Fig. 2. Empirical formulae of deformability of rock masses based on RMR (Rock Mass Rating) and the deformation modulus of the intact rock (ER). (See above-
mentioned references for further information.)

Fig. 3. Empirical formulae of deformability of rock masses based on the deformation modulus of the intact rock (ER) and indexes GSI (Geological Strength Index) or
Barton’s Q. (See above-mentioned references for further information.)

Fig. 4. Analytic models of rock mass deformability, classified based on their theoretical basis. (See above-mentioned references for further information.)

they take into consideration the specific structure of discontinuities of Yang et al (2016) suggested, in addition to their own analytic
the rock mass and, therefore, its anisotropic behaviour. Although there model, preparing a classification of the exiting formulae based on the
are some analytic models based on random distribution of disconti­ persistency of the discontinuities of the rock mass. To extend this
nuities (e.g. Budiansky & O’Connell, 1976) or those that, for the sake approach, another classification was suggested by these authors,
of simplicity, offer averaged results (Fossum, 1985). Due to the fact based on the theoretical basis of the analytic method. Thus, there are
that analytic models have a theoretical basis, they can easily be three types of models. The first and most common is the equivalent
extrapolated for different uses. However, at the same time, their strong continuum method; in this case, the mathematical development re­
dependency on initial hypothesis and assumptions limits their use. For sults are averaged properties of the rock mass, and considered as a
this reason, these models can only be applied in very specific cases. continuum. There is one sub-group of this type of model which de­
This is what happens, for example, with the models by Kulhawy & serves a mention, this is the energetic method because it is frequently
Ingraffea (1978), that valid for rock masses with three orthogonal used, and based on the variation of potential energy caused by the
discontinuity families, or Zhang et al (2011), for heavily jointed rock discontinuities of the rock mass. The third and less common type is the
masses. analysis of discrete blocks.

3
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Table 1 Table 2
Empirical formulations for the deformability of rock masses. Origin of the data. Examples of analytic models for the assessment of deformability of rock masses.
Parameters References Origin of the data used for the fit
Origin of the data.
Theoretical References Parameters Hypothesis and
Dependent on classification indexes
basis characteristics
RQD,ER (Coon & Merritt, In situ testing (geophysical, boreholes) and
1970) laboratory testing. Data from twelve work Equivalent (Duncan & Kn, Ks, s, ER and νR Rock masses with closed
sites. continuum Goodman, discontinuities (zero
(Zhang & Einstein, Technical literature (Coon y Merritt (1970), 1968) thickness). Dilatancy of
2004) Bieniawski (1978) joints is not considered
RQD (Chun et al (2006)) 61 data sets from linear projects located in (Kulhawy & Kn, Ks, s, ER and νR Orthotropic rock mass
South Korea, in various lithologies. Ingraffea, with three orthogonal
RMR,ER (Nicholson & Data from hydraulic projects in South Africa. 1978) families of joints.
Bieniawski, 1990) Boreholes, UCS* laboratory testing, load plate Dilatancy of joints is not
tests. considered
(Shen et al, 2012) Technical literature (Bieniawski, 1978; (Serrano & Orientation and Rock mass with any
Serafim and Pereira, 1983; Stephens and Soriano, thickness of joints, s, spatial distribution of
Banks, 1989) 1986) ED and νD of joints, joints. Persistent joints
RMR (Bieniawski, 1978) Data from hydraulic projects in South Africa. ER y νR and homogeneous stress
Boreholes, UCS* laboratory testing, load plate state are assumed.
tests. (Zhang, Kn, Ks, s, ER and νR Isotropic rock mass,
(Serafim & Pereira, Technical literature (Bieniawski, 1978) and 2010) heavily jointed with
1983) data sets from dam foundations. random distribution.
(Diederichs & Data from extensometers located in mines in Same spacing and
Kaiser, 1999) Ontario (Canada) constitutive properties
Q,ER (Singh & Bhasin, Technical literature. In addition, data of tests are assumed for all the
1996) in the two walls, roof and floor of tunnels were discontinuity families
considered. involved.
Q (Grimstad & Seismic profiles and boreholes in hard rock. Energetic (Budiansky & KR, ER, νR Isotropic and
Barton, 1993) Data from works in Sweden and Norway O’Connell, Crack density. homogeneous rock mass,
GSI,ER (Hoek & In situ testing in China and Taiwan, and 1976) Size of elliptic joints. heavily jointed with a
Diederichs, 2006) technical literature data (Bieniawski, 1978; Stress state random distribution of
Serafim and Pereira, 1983; Stephens and discontinuities. Non-
Banks, 1989) persistent joints, with or
GSI (Vásárhelyi & Technical literature. No tests were performed without water are
Kovács, 2017) specifically for this research. assumed. Elliptic
Rmi (Palmström & In situ testing data in Nepal, India and Bhutan roughness of the
Singh, 2001) discontinuities is used.
Composite (Nejati et al (2013)) Data from the dam site of Gotvand (Iran) Considers the interaction
Not dependent on classification indexes of joints, but not their
vP * (Jasarevic & Data from work sites in the coastal area of closure.
Kovacevic, 1996) CroatiaDependent on the propagation velocity (Yang et al., KN, KS, s, ER and νR Assumption of non-
of longitudinal waves (p-waves). 2016) persistent
vP *, UCS* (Barton, 2007) Data from metro stations in South America discontinuities. 2D
and tunnel perforations in China, United analysis
Kingdom and Norway.Dependent on the Discrete (Hart et al., Individualized block 3D analysis based on the
propagation velocity of longitudinal waves (p- blocks 1988) parameters and equations of movement
waves) and the uniaxial compression strength geometric definition of a three-dimensional
of the intact rock. system of blocks.
* Considers interaction of
Where vP is the velocity of propagation of p-waves, and UCS is the uniaxial
blocks.
compression strength of the intact rock (Shi, 1988) Individualized block Based on the equilibrium
parameters and equations of a system of
Some of the most widely used and indexed analytic models are sum­ geometric definition blocks. Large
deformations
marised in Table 2, that also includes the parameters on which they are
calculation, both static
dependent and their initial hypothesis. and dynamic.
A new equivalent continuum model is now presented in the following
Notation:
sections. The main contribution of this new model, relative to the ones
Kn, Ks: Normal and shear stiffness of discontinuities
already mentioned, is three-fold. The first is the possibility of considering
ER: Deformation modulus of intact rock
a variable number of families of joints with any orientation. This results in KR: Volumetric compressibility modulus of intact rock
a greater versatility of the geometries of rock mass structures that can be νR: Poisson coefficient of intact rock
assessed by the model. The second contribution has the possibility of s: Spacing of discontinuities
considering the influence of other geometric parameters, such as the
thickness and spacing of the discontinuities. Finally, the model presented be considered. Based on this, each family of joints has influence on the
now can be used to consider any constitutive formulation for the intact behaviour of the rock mass depending on its direction. For the model to
rock and the discontinuities, thus expanding the range of cases that it can be completed, in addition to the mathematical algorithm described in
be analysed. this section, the constitutive formulae for both intact rock and discon­
tinuities have to be adapted and implemented.
3. Assembly mathematical algorithm In this model the rock mass is considered as a macro-material consti­
tuted by different micro-materials: the intact rock and all the sets of joints.
3.1. General overview The micro-material is defined as the volume of the rock mass with the
3.1.1. Theoretical basis same behaviour, for constitutive model and parameters. In the case of
The deformability model presented in this paper is based on a families of discontinuities, those enclosed in a micro-material must also
mathematical algorithm of assemblage of matrices. With this procedure have the same opening, spacing and orientation, so that they induce
any orientation of the families, also referred to as a set of global axes, can

4
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

anisotropy in only one direction. Depending on the type of behaviour of Table 3


the micro-materials, the constitutive formulae fit for them will be chosen Parameters in which the deformability model of the rock mass is based.
and adapted accordingly, thus obtaining their compliance matrices. After STRUCTURE PARAMETERS OF THE ROCK MASS (For each i family)
the assemblage of all those matrices, the result is the compliance matrix of
Number of joint sets N [-]
the rock mass.
It should be pointed out that the assembly is a process between the Average dip direction angle of each set θi [◦ ]

matrices of only two micro-materials. This means that the consideration of Average dip angle of each set βi [◦ ]

a number N of families of discontinuities requires the process of assem­ Mean spacing of each set ei [m]

blage to be carried out N times, just as it is shown in Fig. 5. This successive Mean thickness of each set ai [m]

assembly process is performed on FEM models combined with damage BEHAVIOUR PARAMETERS OF JOINTS* (For each i family)
Deformation modulus of the joint fill EDi [kPa]
statistics mechanics (Rinaldi et al., 2006).
Poisson coefficient of the joint fill [-]
The full analytic model obtained by implementing this process can
νDi
BEHAVIOUR PARAMETERS OF INTACT ROCK
consider any number of families, each with its specific orientation and
Deformation modulus of the intact rock ER [kPa]
properties. To do this, various base axes are considered; the global axes of
Poisson coefficient of the intact rock νR [-]
the rock mass and one set of local axes for each family. This results in a
generalisation of the geometries that can be studied, as well as considering * In the filled joints, the deformability is governed by the fill and therefore,
the anisotropy produced by the sets of discontinuities. Hooke parameters are used. (in section 3.5 other parameters are introduced for a
more complex joint model)
3.1.2. Parameters
Aside from the geometric configuration of the rock mass and its 3.1.4. Coordinate systems
discontinuity structure, the deformability model requires the param­ The application of the assemblage process presented in this section
eters specific to the constitutive models used for the intact rock and the requires the use of a set of local coordinates, defined for each family of
joints. The characterisation of the discontinuity structure of the rock discontinuities. However, to enable the implementation of a standard
mass is made through the average orientation of the families, as well as process for any number N of joint sets with different orientations and,
their mean spacing and thickness. The parameters of the specific model therefore, different local bases, the definition of a global set of axes is
presented in this paper focused on rock masses with filled joints and needed in the first place. This global coordinate system is shown in Fig. 6
are included in Table 3. The assignment of values to the parameters of (and it is a canonical set of axes. By definition, axes 1* and 2* are within
distribution, orientation, and behaviour can be carried out, according the horizontal plane, pointing in the North and East directions respec­
to the guidelines given by the ISRM (1978). tively. Regarding axis 3*, it is vertical and pointing upwards.
For a general set of discontinuities, the configuration of the corre­
3.1.3. Hypotheses sponding local set of axes is given in Fig. 6b. The base is again canonical,
The analytic model was developed from a series of assumptions that defined so axes 2 and 3 are within the plane of the joint axes and axis 1 is
limit its application. For this reason, these assumptions have to be taken orthogonal to this plane. In this study, it has been established that axis 2
into account and are summarised below. is orthogonal to the dip direction of the joints. As can be seen in Fig. 6b
the definition of the local coordinates is given which is based on average
- For any increment in the stress state, there is a constitutive relation of the strike (θ) and the dip (β) of the joint set.
that enables the evaluation of the increment in the strain state.
- The contacts between discontinuities and intact rock are plane and 3.2. Assembly procedure for two micro-materials
parallel.
- The stress state is homogeneous in the micro-materials. As a result, When the constitutive relationship between stress and strain can be
the model works at a point-level. set to a formula, as happens with the elasticity equations, the compli­
- The tensors of stress and strain are assumed to be symmetric. ance matrix D of the material can be defined. This matrix relates the
- Joints are assumed to be persistent. fields of stresses and strains, as shown in the following equation for the
case of the rock mass:
Rockmass; ε = D⋅σ (1)
The compliance matrices corresponding to each micro-material (i.e.
intact rock and each set of discontinuities) are considered to be known,
as well as the average orientation, thickness and spacing of each joint
set. Their constitutive equations are given below:

Family of joints; εD = DD ⋅σ D (2)

Intact rock; εR = DR ⋅σR (3)


The stress and strain fields are expressed as second-order tensors.
Therefore, the constitutive relation between them is built by means of
a fourth-order tensor. Despite this issue and the need to use this
method for problems, the method can be reduced to a matrix form.
This would be done under the assumption of symmetry of the stress
and strain tensors, which can be reformulated in vector notation, as
shown in equation (4). Thus, the compliance tensors would be reduced
to 36-component matrices.

Fig. 5. Assembly sequence of the matrices for a rock mass with N + 1 families
of discontinuities.

5
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Fig. 6. Definition of the global coordinates of the rock mass (a) and local coordinates of a set of joints (b).

⎡ ⎤ ⎡ ⎤ ⎡ ⎤

σ xx σ11 εxx ε11 ⎤
⎢ τxy ⎥ ⎢ τ12 ⎥ ⎢ γxy ⎥ ⎢ γ12 ⎥
⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢ τxz ⎥ ⎢ τ13 ⎥ ⎢ γ xz ⎥ ⎢ γ13 ⎥
σ=⎢ ⎥=⎢ ⎥; ε=⎢ ⎥=⎢ (4)

⎢ σ yy ⎥ ⎢
⎥ ⎣ σ22 ⎥ ⎢
⎢ εyy ⎥⎥ ⎢ ε22 ⎥

⎣ σzz ⎦ ⎦ ⎣ εzz ⎦ ⎣ ⎦
σ33 ε33
τyz τ23 γ yz γ23

In addition to this, the engineering shear strain variable (γ ij ) is used


in this formulation, instead of the original component of the strain
tensor (εij ). Their relation would be:

γ ij = 2⋅εij i∕
=j

To assemble the compliance matrices of the two micro-materials,


matrix and set of discontinuities, it is necessary to know the relation­
ship between the volumes of both of them. For a joint family with a
certain mean spacing (s) and thickness (e), its unit volume can be
determined as per equation (5). For further clarity, this is given in Fig. 7,
which shows that, in an area of the rock mass corresponding to one
Fig. 8. Sign criterion of the stress field (analogous for the strain field).
spacing, the thickness of the joint defines the area occupied by it.

d=
AD e
= (5) 3.2.1. Equilibrium and compatibility equations
A s Prior to the definition of the equations of equilibrium and compati­
Where A is the total area of the rock mass, and AD is the sub-area bility of the compound material (rock mass) and both micro-materials, a
corresponding to the joints. distinction must be made between the components of the stress and
In the local coordinates associated to the set of joints, numbered as strain fields in two groups: those contained in the contact plane of joint
{1, 2, 3}, the equations of equilibrium and compatibility of the rock mass and matrix (shown as strains in Fig. 9) and the rest of them (Fig. 10).
and its component materials. For both stress and strain, the sign criterion To fulfil the equilibrium condition, the value of the stress compo­
is the one shown in Fig. 8, so the positive components in both fields are nents contained in the contact plane of joint and matrix (i.e. σ 11 , τ12 , τ13 )
the compressive ones. must be the same in both materials. On the other hand, the total stress
borne by the rock mass in the directions parallel to the plane of the joint
(σ 22 , σ33 , τ23 ) can be decomposed in the stress borne by the matrix ma

Fig. 7. Model of two micro-materials. Fig. 9. Strain components ε within the contact plane.

6
­
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

EC ⋅ε = EC ⋅εR = EC ⋅εD (7)

EC ⋅σ = d⋅EC ⋅σ D + (1 − d)⋅EC ⋅σR (8)


Within the contact plane

E⋅ε = d⋅E⋅εD + (1 − d)⋅E⋅εR (9)

E⋅σ = E⋅σR = E⋅σ D (10)


The mathematical procedure for the assemblage of the rock mass
compliance matrix can be obtained based on the equations presented up
to this point.

3.2.2. Relation between increments of stress and strain


The stress and strain increments in the micro-materials and the rock
mass (macro-material) can be related by means of the deduced equations
Fig. 10. Strain components ε out of the contact plane. above. In this section, is shown the procedure to obtain these expressions.
The first step is rewrite equations (2), (3), (7) and (10) in terms of
terial and the stress borne by the joint, in their corresponding unit areas. increments.
Following the notation used in equations (1) to (3) for the stress–strain ε̇D = DD ⋅σ̇D (2)
components in the rock mass and the micro-materials, and the stress
equilibrium can be expressed as follows: ε̇R = DR ⋅σ̇R (3)
Stress in the contact plane σ1,i = σ1,i R = σ1,i D i = 1, 2, 3

Out of the contact plane, in terms of force (F) and stress (σ), for an A area of contact, EC ⋅ε̇ = EC ⋅ε̇R = EC ⋅ε̇D (7)
the relations are:
Forces out of the contact plane D
Fi,j = Fi,j + FRi,j ; i = 2, 3; j = 2, 3 E⋅σ̇ = E⋅σ̇R = E⋅σ̇ D (10)
Substituting F = A⋅σ σi,j ⋅A = σDi,j ⋅AD + σRi,j ⋅AR i = 2, 3; j = 2, 3
By substituting equations (2) and (3) in equation (7), the following
expression is obtained:

Where the total area (A) is the result of adding the areas corresponding to the joint set EC ⋅DR ⋅σ̇ R = EC ⋅DD ⋅σ̇D
(AD ) and the matrix material (AR ), so that A = AD + AR . From equation (5), dividing
the latter by the total area A, the results is: This equation is then added to equation (10), with the result shown
Stresses out of the contact plane i = 2, 3; j = 2, 3
below.
σi,j = d⋅σi,j D + (1 − d)⋅σi,j R
( ) ( )
E + EC ⋅DR ⋅σ̇ R = E + EC ⋅DD ⋅σ̇ D

The same approach can be applied to the compatibility equations. The The matrices TRD and TDR and then defined, by rearranging terms in
displacements in directions parallel to the plane of the joint (ε22 , ε33 ,γ23 ), this last expression
must be the same for both micro-materials. Likewise, the group of ( )− 1 ( )
σ̇R = TRD σ̇ D TRD = E + EC ⋅DR ⋅ E + EC ⋅DD (11)
components within the contact plane is constituted by the displacement
on the direction orthogonal to the plane of the joint, alongside the ( )− ( )
(12)
1
remaining shear strains. The displacements in these latter directions are σ̇D = TDR σ̇ R TDR = E + EC ⋅DD ⋅ E + EC ⋅DR
the results of the sum of the contributions of both micro-materials, as
Taking equilibrium equation (8), the variable σ̇ D can be removed by
shown below.
substituting the matrix TDR as defined above i.e. equation (12).
Strains out of the contact plane εi,j = εi,j R = εi,j D i, j = 2, 3
( )
EC ⋅σ̇ = d⋅EC ⋅TDR + (1 − d)⋅EC ⋅σ̇R (13)
In the contact plane, the displacements (u), defined through the strains (ε) and the
analysed length (ΔL), are: u = ε⋅ΔL. They fulfil the following conditions:
Some clarification is needed at this point, regarding the auxiliary
Displacements within the contact plane u1,i = uD1,i + uR1,i i = 1, 2, 3
matrices E and EC (equation (6)). According to their definition, adding
Substitutingu = ε⋅ΔL ε1,i ⋅ΔL = εD1,i ⋅ΔLD + εR1,i ⋅ΔLR
both matrices results in the sixth order identity matrix, this is E + EC =
Again, the total length (ΔL) is the result of adding the lengths corresponding to the
discontinuities (ΔLD ) and the matrix material (ΔLR ).From its definition, equation I6 . In accordance with this, adding equations (10) and (13) and then
(5) can also be expressed as d = ΔLD /ΔL. Substituting ΔL = ΔLD + ΔLR , the result rearranging terms, the following expression is obtained.
is: ( ) ( )
Strains in the contact plane: ε1,i = d⋅ε1,i D + (1 − d)⋅ε1,i R i = 1, 2, 3 E + EC ⋅σ̇ = E + d⋅EC ⋅TDR + (1 − d)⋅EC ⋅σ̇ R
( )
σ̇ = TR⋅σ̇ R = I6 + d⋅EC ⋅(TDR − I6 ) ⋅σ̇R (14)
All these equilibrium and compatibility equations can be written in
Matrix TR is defined as explained. This matrix expresses the rela­
matrix form, if their stress and strain terms are grouped as shown in the
tionship between the fields of stresses of the global material and the
vectors of equation (4), and defining the following auxiliary matrices:
matrix micro-material.
[ ] [ ]
I3 O3 O3 O3 Following a completely analogous procedure, and eliminating vari­
E= ; EC = (6)
O3 O3 O 3 I3 able σ̇ R from equation (8) instead of σ̇D , the matrix TD expression can
then be obtained. This latter matrix relates the fields of stresses of the
Where I3 and O3 are the 3x3 identity matrix and null matrix,
global material and the discontinuity micro-material.
respectively.
Based on all the previous formula, the equilibrium and compatibility Substituting (11) in (8) : ( )
= d⋅EC + (1 − d)⋅EC ⋅TRD ⋅σ̇ D (15)
equations can be rewritten in matrix form as follows: EC ⋅σ̇
Out of the contact plane

7
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Adding( (10) and ( ) As just mentioned, the application of the assemblage method re­
) (15) : = E + d⋅EC + (1 − d)EC ⋅TRD σ̇D quires that the constitutive matrices of the micro-materials to be in the
E + EC σ̇
local coordinates of the joint set. For this reason, each iteration begins
Rearranging terms : ( ) with the transformation of the matrices involved into the set of local
= TD⋅σ̇D TD = I6 + (1 − d)(TRD − I6 )EC (16) axes and ends with the transformation of the composed compliance
σ̇
matrix into the global axes of the rock mass.
As a result of this process, the expressions of the matrices relating the The next section includes a brief description of the procedure for the
stress fields of both micro-materials with that of the global composite base change, as needed for the implementation of the complete math­
material are known. ematical algorithm of assemblage.
σ̇ = TR⋅σ̇ R ; TR = I6 + d⋅EC ⋅(TDR − I6 ) (14)
3.3.2. Base change
σ̇ = TD⋅σ̇D ; TD = I6 + (1 − d)⋅EC ⋅(TRD − I6 ) (16) The base change transformation between the global axes of the rock
mass (base B* = {1* , 2* , 3* }) and the local axes of a joint set (base B =
The same procedure can be carried out for strains. Using the {1, 2, 3}) consists of a composition of rotations. For this reason, an in­
constitutive equations of the micro-materials (2) and (3), the equilib­
termediate basis (B’ = {1’ , 2’, 3’}) is needed. The transformation be­
rium (8), and compatibility (9) equations, the matrices relating the
tween the global and intermediate axes is made through a rotation about
strain fields can be defined as:
the global 3* axis (Fig. 12a). From the intermediate base, the local set of
ε̇R = DRD⋅ε̇D
(
DRD = EC + E⋅D− R
)− 1 ( C
E + E⋅D− D
)
(17) axes is reached by means of another rotation, this time about the 2′
intermediate axis (Fig. 12b).
( )− 1 ( C ) The rotation about the global 3* axis fulfils the following condition:
ε̇D = DDR⋅ε̇R DDR = EC + E⋅D− D E + E⋅D− R (18)
the intermediate axes 1′ and 2′ are aligned with the strike (θ) and the dip
direction (DB ) of the family, respectively. Therefore, it is a clockwise
ε̇ = DR⋅εR ; DR = (I6 + d⋅E⋅(DDR − I6 ) ) (19)
rotation (negative signed), with the following angle θ3 :
( π)
ε̇ = DD⋅ε̇D ; DD = (I6 + (1 − d)⋅E⋅(DRD − I6 ) ) (20) θ3 = − θ = − DB −
2
3.2.3. Global constitutive equation Where the values of the strike (θ) and the dip direction (DB ) of the
Up to this point, we defined the matrices that relate the fields of family are the angles of these directions with the North.
stresses and strains of both micro materials with each other and with the The second rotation, this time about axis 2′ , is made from the in­
fields of the composite. Based on these expressions, the global compli­ termediate trinomial of axes. In the final configuration of the local base,
ance matrix of the composite material can be obtained from equation axis 3 is within the discontinuity plane, while axis 1 is orthogonal to it
(1). and pointing upwards. Therefore, the value of the rotation angle θ2 is a
function of the dip angle (β) of the family:
ε̇ = D⋅σ̇
(π )
Now, the expressions of the relation between the composite material θ2 = − − β
2
and the matrix micro–material are to be substituted in the previous
equation. Again, the sign of the rotation is given by its clockwise (therefore
negative) direction.
Substituting (19) in (1) : ε̇ = DR⋅DR ⋅σ̇ R It is not straightforward to transform 3 base dimension matrices of
change into matrices adapted to the matrix formulation of the problem.
Substituting (14) in (1) : ε̇ = DR⋅DR ⋅TR− 1 ⋅σ̇ By identifying the coefficients, relating the coordinates of the stress and
strain fields in the global, intermediate, and local sets of axes, the change
The global compliance matrix : D = DR⋅DR ⋅TR− 1
(21) of basis takes the following shape:
The matrices relating the stress and strain fields of the composite σ ’ = T3 ⋅σ; ε’ = D3 ⋅ε; σ * = T2 ⋅σ ’; ε* = D2 ⋅ε’
material with the discontinuity micro-material can be used in the exact
same way to obtain another expression of the compliance matrix. Given that in each coordinate system equation (1) is fulfilled, the
D D base change expressions of the global compliance matrix D between the
Substituting (20) in (1) : ε̇ = DD⋅D ⋅σ̇

Substituting (16) in (1) : ε̇ = DD⋅DD ⋅TD− 1 ⋅σ̇

The global compliance matrix : D = DD⋅DD ⋅TD− 1


(22)
As a result of these assembly operations, we know the constitutive
expressions of the composite material: i.e. matrix material with one
family of joints. Based on this simple method for two micro-materials,
the formulation can be extended to cover a general case with any
number of joint sets.

3.3. Modification of the process for any number of joint sets


3.3.1. Procedure of successive inclusions
The method just described above can be applied to the composition
of two micro-materials, one of them acting as a matrix and one as a
discontinuity family. For the assembly of the compliance matrix of a
rock mass with N joint sets, it is necessary to compose a procedure of
successive additions of the families, as described in Fig. 8. The complete Fig. 11. Process diagram of assemblage of matrices for a number N of disconti­
process is shown with greater clarity in Fig. 11. nuity sets.

8
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Fig. 12. Transformation between the local and global coordinates through a composition of rotations about global axis 3* (a) and intermediate axis 2′ (b).

local and global sets of axes are: The definition of the behaviour of the discontinuities, however, is
( ) much more complex. The formulation developed up to this point is gen­
Global→Local; D = D2 ⋅ D3 ⋅D* ⋅T −3 1 ⋅T −2 1
eral and can be applied to any type of rock mass, regardless of the nature
( ) of its intact rock and joints, which are defined by their corresponding
Local→Global; D* = D−3 1 ⋅ D−2 1 ⋅D⋅T2 ⋅T3 constitutive formulations. At this point the number of formulae involved
The expressions of the matrices required for the change of basis are in the specific model described in this paper needs to be narrowed. The
included below, and already adapted for the matrix formulation of this resulting mathematical algorithm will be applied to the particular case of
mathematical method. The matrices of rotation about the global 3* axis rock masses with filled joints, where the fill governs the deformational
are: behaviour of the discontinuity. Further clarification on the joint behaviour
⎡ ⎤ classification is given in Fig. 13.
(cosθ3 )2 sin2θ3 0 (sinθ3 )2 0 0 Multiple options are available to define the behaviour of the dis­
⎢ − 0.5sin2θ3 cos2θ3 0 ⎥
⎢ 0 0.5sin2θ3 0 ⎥ continuities. In the case of type D joints, it was assumed to have a linear
⎢ 0 sinθ3 ⎥

T3 = ⎢
0 0 cosθ3 0 ⎥ elastic behaviour for this fill, since the constitutive law is governed by
2
⎢ (sinθ3 ) − sin2θ3 0 2
(cosθ3 ) 0 0 ⎥ ⎥ the fill of the joints. The compliance matrix of the joints was, as a
⎣ 0 0 0 0 1 0 ⎦
consequence, that of a Hookean solid. Therefore, the structure of the
0 0 − sinθ3 0 0 cosθ3
compliance matrix of both micro-materials was, in this specific case of
⎡ ⎤ study, the one shown in equation (23). It must be noted that, in the case
(cosθ3 )2 0.5sin2θ3 0 (sinθ3 )2 0 0
⎢ − sin2θ3 ⎥ of the discontinuities, this matrix as shown in equation (23) has been
cos2θ3 0 sin2θ3 0 0

⎢ 0 0 cosθ3 0 0 sinθ3

⎥ defined in their set of local axes.
D3 = ⎢
⎢ (sinθ )2 − 0.5sin2θ


⎢ 3 3 0 (cosθ3 )2 0 0 ⎥
⎣ 0 0 0 0 1 0 ⎦
0 0 − sinθ3 0 0 cosθ3
While the matrices of rotation about the intermediate 2′ axis are:
⎡ ⎤
(cosθ2 )2 0 − sin2θ2 0 (sinθ2 )2 0
⎢ 0 cosθ2 0 0 0 − sinθ2 ⎥
⎢ ⎥
⎢ 0.5sin2θ 0 cos2θ 0 − 0.5sin2θ 0 ⎥
T2 = ⎢

2 2 2 ⎥

⎢ 0 0 0 1 0 0 ⎥
⎣ (sinθ2 )2 0 sin2θ2 0 (cosθ2 ) 2
0 ⎦
0 sinθ2 0 0 0 cosθ2
⎡ ⎤
(cosθ2 )2 0 − 0.5sin2θ2 0 (sinθ2 )2 0
⎢ 0 cosθ2 0 0 0 − sinθ2 ⎥
⎢ ⎥
⎢ sin2θ 0 cos2θ2 0 − sin2θ2 0 ⎥
D2 = ⎢

2 ⎥

⎢ 0 0 0 1 0 0 ⎥
⎣ (sinθ2 )2 0 0.5sin2θ2 0 (cosθ2 )2 0 ⎦
0 sinθ2 0 0 0 cosθ2

3.4. Constitutive matrices. Model of rock mass with filled discontinuities


The obtention process for the compliance matrix of the rock mass has
already been completely described. However, the compliance matrices
of the micro-materials involved in the assessment (i.e. intact rock and
joint sets) remain undefined. This is the next and last step to have a
complete calculation algorithm that can be implemented.
The linear elastic behaviour of a Hookean solid can be assumed for
the intact rock. Therefore, the compliance matrix can be defined with
the deformation modulus ER and the Poisson coefficient νR of the intact Fig. 13. Classification of joints according to their thickness and presence of fill.
rock, as per equation (23). Source: Barton, 1973

9
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

⎡ ⎤ [
1 νi νi ( )2/3 ]
⎢ E 0 0 − − 0 ⎥ 3(2 − νD )tan(ϕB )⋅Nα Tα
⎢ i Ei Ei ⎥ uT α = 1− 1−
⎢ ⎥ 8GD (R⋅uN α )1/2 tan(ϕB )⋅Nα
⎢ ⎥
⎢ 2(1 + νi ) ⎥
⎢ 0
⎢ 0 0 0 0 ⎥
⎥ Defining the auxiliary variable f from the residual friction angle of
Ei



⎥ the joint (ϕB ) and R being the average radius (harmonic mean between
⎢ ⎥ R1 and R2).
⎢ 2(1 + νi ) ⎥
⎢ 0 ⎥
⎢ 0
Ei
0 0 0 ⎥ As a non-linear law, the mathematical treatment requires solving the
⎢ ⎥
D =⎢
i ⎥ equations incrementally, which can be obtained by differentiation:

⎢ νi 1 νi

⎥ (23)
⎢− 0 0 − 0 ⎥ dTα = KD ⋅duN α + KT ⋅duT α
⎢ Ei Ei Ei ⎥
⎢ ⎥
⎢ ⎥
⎢ ⎥ dNα = KN ⋅duN α
⎢ νi νi 1 ⎥
⎢− 0 0 − 0 ⎥
⎢ Ei ⎥
⎢ Ei Ei ⎥ The coefficients of the equations above can be determined with the
⎢ ⎥



2(1 + νi ) ⎦ following expressions:
0 0 0 0 0
Ei 2GD (R⋅uN α )1/2
KN =
Index i = R, D (1 − νD )

Where the parameters to be defined are: [ ( )1/3 ]



KD = tan(ϕB )⋅KN ⋅ 1 − 1−
tan(ϕB )⋅Nα
Ei Deformation modulus of the material (intact rock ER , or fill of
joint ED ). [ ]1/3
νi Poisson coefficient of the material (intact rock νR , or fill of joint KT =
2(1 − νD )
⋅KN ⋅ 1 −

νD ). (2 − νD ) tan(ϕB )⋅Nα
Assuming a uniform distribution of the geometric parameters of the
3.5. Constitutive matrices. Non-linear model of discontinuities roughness of the joint with a density (ρ) of contacts per unit area of the
joint and a mean spacing (s) between joint planes of the same family, a
The deformability formulation developed for the rock mass can constitutive matrix can be obtained for each family of joints in local
incorporate more general constitutive models other than linear elastic coordinates. This is analogously to the linear case of section 3.4. Hence:
ones. Thus, to make the deformational behaviour dependent on the ⎡ ⎤
0.5⋅D12 0.5⋅D12
characteristics of the joint walls, we chose to introduce in this section the ⎢ D11 0 0 0 ⎥
sinδ cosδ
Hertz-Mindlin analytical formulation. Mindlin’s formulation (1949) can ⎢



⎢ 2⋅sinδ⋅D21 2⋅D22 0 0 0 0 ⎥
include shear load to the Hertz equation (1882), which was initially ⎢ ⎥
⎢ ⎥
developed for contacts between spheres under normal load. Thus, a non- ⎢ 2⋅cosδ⋅D21 0 2⋅D22 0 0 0 ⎥
⎢ ⎥
linear elastic behaviour is simulated. ⎢ 1 ⎥
DD = ⎢ 0 0 0 0 0 ⎥
To apply the Hertz-Mindlin equations, it is necessary to characterize ⎢
⎢ ED ⎥

the contacts between the joint walls geometrically with the three vari­ ⎢
⎢ 1


⎢ 0 ⎥
ables illustrated in Fig. 14: the radii of curvature of the two roughness ⎢ 0 0 0 0
ED ⎥
⎢ ⎥
involved (R1 and R2) and the angle of inclination of the contact plane ⎢


2 ⎦
with respect to the joint median plane (α). 0 0 0 0 0
ED
Considering that the normal (N) and shear (T) load act in the lon­
gitudinal median plane of the joint (π plane) and two systems of Car­ These shown variables correspond to:
tesian axes can be defined (Fig. 14): the global axes of the discontinuity ⎡ ⎤
(x, y) and the local axes of roughness (xα , yα ). [
D11 D12
]
1 T 1 ⎣
1 0
Under these conditions, the deformational properties of the rock on D= = Q ⋅ ⋅ K KN ⎦⋅Q
D13 D14 ρ⋅s KN − T
the joint faces are: transverse elastic modulus (GD ) and Poisson’s ratio KD KT
(νD ), and the Hertz-Mindlin equations can obtain the normal (uNα ) and [ ] [ ]
transverse (uTα ) displacements from: Q=
cosα − sinα
; QT =
cosα sinα
sinα cosα − sinα cosα
4GD 3/2
Nα = ⋅R1/2 ⋅uN α
3(1 − νD )

Fig. 14. Joint geometry, cut along the longitudinal median plane (π).

10
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

( )
/ Table 4
δ = atan τxy Properties of the intact rock for the validation cases of the analytical model.
τxz
Intact rock deformation modulus ER [MPa] Intact rock Poisson coefficient νR [-]
Where Q is the base change matrix in the contacts and τxy , τxz are the
tangential stresses on the joint plane. 1000 0.25
Several hypotheses are required to relate the normal and shear
stresses and strains of the joint in the 3D axes of the assembly formu­
lation. Each of the 3D shear strain (εxy , εxz ) depends on the shear stress
(τxy , τxz ) in its same direction, and independent of each other. In addi­
tion, it is considered that the normal strain depends on both shear
stresses (τxy , τxz ) in the same way as just mentioned.
The non-slip equilibrium condition for the stresses (σ α , τα ) in the
contact must be fulfilled. It is reasonable to assume that the joint
deformability model imposes the equilibrium of the system as a neces­
sary condition, that is: |τα | ≤ tanϕB ⋅σα

4. Implementation and validation

The described mathematical algorithm was implemented in the Fig. 15. Validation case 1. One horizontal family of joints.
matrix calculation program MATLAB, with the deformation properties
of the micro-materials involved and the geometrical distribution of the with the notation shown in Fig. 15.
sets of joints as inputs. More specifically: the number of families, their ⎡ ⎤
strike and dip orientation, their spacing and thickness had to be defined 1
0 0 −
νHH

νVH
0 ⎥
⎢ E
for the program to function, the deformation moduli and Poisson co­ ⎢ H

EH EH ⎥

efficients of the intact rock, and all the sets of joints of the rock mass. All ⎢
⎢ 0 2(1 + νHH ) ⎥
0 0 0 0 ⎥
these properties were, for the case at hand, independent of the stress ⎢
⎢ EH ⎥

⎢ ⎥
field. ⎢ 1 ⎥
⎢ 0 0 ⎥
For the calculation, isotropic behaviour was assumed for both intact ⎢

0
GV
0 0 ⎥

rock and joint fill. The anisotropy of the rock mass due to the presence of S=⎢ ⎥ (25)
⎢ νHH 1 νVH ⎥
the discontinuities was considered. The resultant numerical calculation ⎢−
⎢ EH 0 0 − 0 ⎥

⎢ E EH ⎥
is stable and, since it does not depend on initial values, was finished in
H
⎢ ⎥
⎢ νVH ν 1 ⎥
one iteration. ⎢− 0 0 −
VH
0 ⎥
⎢ EH EH EV ⎥
The results of the program include: the compliance matrix of the rock ⎢



mass and the breakdown of the global stress state into the local stress ⎣
0 0 0 0 0
1 ⎦
fields of the micro-materials. To obtain the latter the global stress field GV
had to be among the inputs of the program. We checked that the order
The Poisson coefficients used in the anisotropic model in terms of the
the joint sets were included in the composite material did not have any
strains in the three directions of study, were defined as shown below.
significant influence in the final results.
εH2 ε2
νHH = − = −
4.1. Validation with simple theoretical cases εH1 ε1
The formulation was checked by comparing the results with those of εH ε12
simple theoretical cases with rock mass deformability parameters that νVH = −
εV
= −
ε3
can be calculated with analytical formulae. This is a case of a multi-
layered medium, where the deformability was calculated in the The properties of the family of joints considered in this case are
orthogonal direction to the discontinuity planes as per (Kulhawy & summarised in Table 5.
Ingraffea, 1978). In this case and according to the general elasticity Given all these data and with equation (24), the value of the elastic
formulae, the equivalent deformation modulus of the rock mass (EEQ ) modulus of the rock mass in the vertical direction (EV ) would be 100
can be calculated using equation (24), assuming that the stress in­ MPa. Assuming a global Poisson Coefficient of 0.25, the shear modulus
crements are the same for both intact rock and discontinuities. GV would be 40 MPa.
Regarding the program results for the analytical model the compli­
EEQ = ∑
hTOT
(24) ance matrix of the rock mass is in MPa− 1 :
i (hi /Ei ) 0.00111 0.00000 0.00000 ¡0.00028 ¡0.00028 0.00000
Where h stands for the thickness of the layers i.e. intact rock and 0.00000 0.00278 0.00000 0.00000 0.00000 0.00000
discontinuities. 0.00000 0.00000 0.02725 0.00000 0.00000 0.00000
The cases used in the validation assessment had up to three families ¡0.00028 0.00000 0.00000 0.00111 ¡0.00028 0.00000
0.00000 0.00000 0.00927 0.00000
of discontinuities in different directions, with their corresponding
¡0.00028 ¡0.00028
0.00000 0.00000 0.00000 0.00000 0.00000 0.02725
deformational properties. In all the cases studied, the properties of the
intact rock are the ones included in Table 4.
It can be seen that the structure of this matrix coincides with the
4.1.1. Case 1: one horizontal discontinuity set structure of the theoretical matrix of the anisotropic solid in equation
The first case that was analysed involves a rock mass with just one set (25) and comparing terms in both matrices, the values of the equivalent
of discontinuities in the horizontal direction (Fig. 15). Its behaviour can moduli are shown in Table 6.
be simulated by means of an elastic solid with one anisotropy direction.
In accordance with the elasticity theory, the compliance matrix of such a 4.1.2. Case 2: two horizontal joint sets
solid using the Hooke formulae is the one shown in equation (25), in a One more set of joints was then added to the previous case, to check
base of global coordinates like the one described in section 3.1.4 and the behaviour of the mathematical algorithm when assembling various

11
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Table 5
Orientation and properties of the joints present in case 1 of validation. One horizontal joint set.
Set Dip direction (◦ ) Dip β (◦ ) Spacing s (m) Thickness e (m) Fill modulus EDi [kPa] Fill Poisson coefficientνDi

1 45 54.74 1 0.1 10,000 0.25

Table 6 Table 7
Result comparison. Validation case 1. Orientation and properties of the joints present in case 2 of validation. Two
Proposed Simplified theoretical Result comparison
horizontal joint sets.
analytical formulation (equation Set Dip Dip Spacing Thickness Fill Fill Poisson
model (24)) direction β (◦ ) s (m) e (m) modulus coefficientνDi
(◦ ) EDi [kPa]
EV = 108MPa EV = 100MPa εrel = 8.0%
EH = 901MPa Modulus of intact The modulus in the direction 1 0 0.00 2 0.1 10,000 0.25
rock:EH = ER = parallel to the joint plane can be 2 0 0.00 2 0.1 1000 0.25
1000MPa compared to that of the intact rock,
with a slight reduction due to the
presence of the joint set.εrel = 9.9% (continued )
GH = 360MPa GH = 361MPa εrel = 0.28% 0.00111 0.00000 0.00000 ¡0.00028 ¡0.00028 0.00000
EV = 108MPa EV = 100MPa εrel = 8.0%
¡0.00028 0.00000 0.00000 ¡0.00028 0.04469 0.00000
0.00000 0.00000 0.00000 0.00000 0.00000 0.13351

families of discontinuities, with parallel directions (see Fig. 16) and


significantly different deformational properties. Again, the rock mass
could be simulated in terms of an elastic solid with one plane of As in case 1, the structure of the matrices coincides with the theoretical
anisotropy, with the compliance matrix shown in equation (25) and the compliance matrix (equation (25)) and comparing terms in both matrix,
notation of Fig. 16. the values of the elastic moduli are summarised in Table 8.
The characteristics of the two sets of discontinuities are summarised
in Table 7. Two assessments were carried out, changing the order of 4.1.3. Case 3: two orthogonal joint sets, horizontal and vertical
inclusion of the joint sets: order 1: set 1, set 2; order 2: set 2, set 1. This Now that the behaviour of the analytic model when assembling
was done to check the variations that could be caused by the order various families of discontinuities has been checked, the next step was to
chosen for the assemblage of the compliance matrix. analyse a case with two joint sets in different directions. Thus, a rock
In this case, the elastic modulus in the vertical direction (EV ) has a mass with two orthogonal joint families was assessed in this case. The
value of 20 MPa, according to equation (24). The shear elastic modulus, behaviour of the rock mass can be compared to that of an orthotropic
on the other hand, has a value of 8 MPa. elastic solid, with the compliance matrix shown in equation (26), in
The compliance matrix of the rock mass obtained by the analytic accordance with the notation in Fig. 17.
model in MPa− 1 , is: ⎡ ⎤
1 ν12 ν13
Order 1 ⎢ E 0 0 − − 0 ⎥
⎢ 1 E2 E3 ⎥
0.00111 0.00000 0.00000 ¡0.00028 ¡0.00028 0.00000 ⎢ ⎥
⎢ 1 ⎥
0.00000 0.00277 0.00000 0.00000 0.00000 0.00000
⎢ 0 0 0 0 0 ⎥
⎢ G ⎥
0.00000 0.00000 0.13913 0.00000 0.00000 0.00000 ⎢ 12 ⎥
⎢ ⎥
¡0.00028 0.00000 0.00000 0.00111 ¡0.00028 0.00000 ⎢ 1 ⎥
⎢ 0 0 0 0 0 ⎥
¡0.00028 0.00000 0.00000 ¡0.00028 0.04656 0.00000 ⎢ G13 ⎥
⎢ ⎥
0.00000 0.00000 0.00000 0.00000 0.00000 0.13913 S=⎢ ⎥ (26)
⎢ ν21 1 ν 23 ⎥
⎢− 0 0 − 0 ⎥
⎢ E1 E2 E3 ⎥
⎢ ⎥
Order2: ⎢
⎢ ν31 ν32 1


⎢− 0 0 − 0 ⎥
⎢ E1 E2 E3 ⎥
⎢ ⎥
0.00111 0.00000 0.00000 ¡0.00028 ¡0.00028 0.00000 ⎢ ⎥
⎣ 1 ⎦
0.00000 0.00277 0.00000 0.00000 0.00000 0.00000 0 0 0 0 0
G23
0.00000 0.00000 0.13351 0.00000 0.00000 0.00000
¡0.00028 0.00000 0.00000 0.00111 ¡0.00028 0.00000 Where the following expressions apply to the Poisson coefficient:
(continued on next column)
εj
νij = − ; i, j = 1, 2, 3
εi
νij νji
= ; i, j = 1, 2, 3
Ej Ei

The characteristics of the two joint sets present in the rock mass can
be seen in Table 9.
With all these inputs, the resulting compliance matrix when running
the program with the analytic model is in MPa− 1 :
0.00123 0.00000 0.00000 ¡0.00031 ¡0.00031 0.00000

0.00000 0.01083 0.00000 0.00000 0.00000 0.00000


0.00000 0.00000 0.02922 0.00000 0.00000 0.00000
¡0.00031 0.00000 0.00000 0.00382 ¡0.00059 0.00000
¡0.00031 0.00000 0.00000 ¡0.00059 0.00997 0.00000
0.00000 0.00000 0.00000 0.00000 0.00000 0.03286
Fig. 16. Validation case 2. Two horizontal families of joints.

12
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Table 8
Result comparison. Validation case 2.
Proposed analytic model Simplified theoretical formulation Result comparison
(equation (24))
Order 1 Order 2

EV = EV = EV,T = 20MPa εrel = 5%/10%


21MPa 22MPa
EH = EH = Modulus of the intact rock:EH,T = The modulus in the direction parallel to the joint plane can be compared to that of the intact rock, with a slight
903MPa 903MPa ER = 1000MPa reduction due to the presence of the joint setsεrel = 9.7%
GV = 7MPa GV = 7MPa GV,T = 8MPa εrel = 0%
GH = GH = GH,T = 361MPa εrel = 5%/10%
361MPa 361MPa

4.1.5. Summary of results


The assessment of the four cases described in this section leads to the
validation of the model. The analysed configurations are simple, thus
enabling the obtention of the rock mass deformation moduli in the di­
rections perpendicular to the joint sets by means of the use of the
theoretical unidimensional formulation of the elasticity for a multi-
layered medium.
Comparing the results of the theoretical formulation with those ob­
tained by means of the analytic model presented in this paper leads to
relative errors in the values of the moduli in the range of 0% to 12.5%. In
the case of the moduli in the directions parallel to the joint sets, the
reduction of the rock mass moduli, when compared with that of the
intact rock, goes from 9.7% (joint sets in one direction) to 18.6% (joint
sets in two directions).
Fig. 17. Validation case 3. Two orthogonal joint sets, horizontal and vertical.

Table 10
Result comparison. Validation case 3.
This structure coincides with that of the theoretical matrix for this type Proposed Simplified theoretical Result comparison
of solid (orthotropic), as per equation (26). Based on this context, the analytical formulation (equation
model (24))
resulting values of the elastic moduli obtained from the program and the
theoretical simplified formulation are summarised in Table 10. E3 = 100MPa E3 = 100MPa εrel = 0%
E2 = 262MPa E2 = 254MPa εrel = 3.1%
4.1.4. Case 4: three orthogonal families E1 = 814MPa E1 = ER = 1000MPa The modulus can be compared to
that of the intact rock, with a slight
The last case that was assessed consists of a rock mass with three
reduction due to the presence of the
orthogonal joint sets, shown in Fig. 18. The results of this analysis is joint setsεrel = 18.6%
shown below.
As in case 3, the behaviour of this particular rock mass can be
modelled by means of an orthotropic elastic solid. Therefore, its theo­
retical compliance matrix would be equation (26), with the notation
shown in Fig. 18. On the other hand, the characteristics of the joint sets
of this case are shown in Table 11.
The result of the programmed analytic model with these inputs is the
following compliance matrix, in MPa− 1 :
0.00533 0.00000 0.00000 ¡0.00042 ¡0.00064 0.00000

0.00000 0.02225 0.00000 0.00000 0.00000 0.00000


0.00000 0.00000 0.03880 0.00000 0.00000 0.00000
¡0.00042 0.00000 0.00000 0.00421 ¡0.00066 0.00000
¡0.00064 0.00000 0.00000 ¡0.00066 0.01083 0.00000
0.00000 0.00000 0.00000 0.00000 0.00000 0.03547

Once again, its structure is the same as that of the theoretical compliance
matrix equation (26). By substitution, the deformation moduli of the
analytic model were obtained and summarised, alongside the results of
Fig. 18. Validation case 4. Three orthogonal joint sets.
the simplified elastic equation, in Table 12.

Table 9
Orientation and properties of the joints present in case 3 of validation. Two orthogonal joint sets.
Set Dip direction (◦ ) Dip β (◦ ) Spacing s (m) Thickness e (m) Fill modulus EDi [kPa] Fill Poisson coefficientνDi

1 0 0.00 1 0.1 10,000 0.25


2 0 90.00 1 0.1 30,000 0.25

13
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Table 11
Orientation and properties of the joints present in case 4 of validation. Three orthogonal joint sets.
Set Dip direction (◦ ) Dip β (◦ ) Spacing s (m) Thickness e (m) Fill modulus EDi [kPa] Fill Poisson coefficientνDi

1 0 0.00 1 0.1 10,000 0.25


2 0 90.00 1 0.1 30,000 0.25
3 90 90.00 1 0.1 20,000 0.25

The elastic properties of the blocks are characterized in Table 13,


Table 12
while the deformability of the discontinuities is defined from their
Result comparison. Validation case 4.
normal (Kni ) and shear (Ksi ) stiffnesses. In this case of filled disconti­
Proposed analytical Simplified theoretical formulation Result nuities of elastic behaviour and thickness e, the stiffnesses can be
model (equation (24)) comparison
calculated using the following expressions:
E3 = 92MPa E3 = 92MPa εrel = 0%
EDi (1 − νDi ) EDi
E2 = 237MPa E2 = 233MPa εrel = 1.7% Kni = ; Ksi =
E1 = 188MPa E1 = 168MPa εrel = 11.9%
t(1 − 2νDi )(1 + νDi ) 2t(1 + νDi )
To calculate the deformability of the set of blocks in each direction, a
uniaxial test was induced in each of the three directions. This consisted
4.2. Numerical validation of imposing a constant velocity on the boundary in which the load was
applied and recording the reactions generated both in this boundary and
4.2.1. Analysis case in the opposite one in which movement is inhibited. These two bound­
For a more general validation, it is necessary to use specially devel­ aries simulated the effect of the plates of a press, while the other two
oped numerical models to evaluate the deformability of fractured rock boundaries were left unrestricted. The sum of the reactions in each
masses. Thus, it is possible to analyse a rock mass crossed by three families boundary divided by its area provides the average stress generated in
arranged according to randomly chosen orientations and properties in each load increment.
space. The properties of the 3 families of joints and of the intact rock The model stresses are zero when the simulation of a test is started.
considered in this case are summarized in Table 13. This strategy, therefore, requires that the velocity be low enough to
With all these inputs, the resulting compliance matrix when running avoid possible numerical errors due to abnormally large stress–strain
the program with the proposed analytic model is in MPa− 1 : increases, especially at the beginning of the simulation. The problem is
0.00220 0.00115 0.00147 ¡0.00038 ¡0.00040 ¡0.00045 that with low strain velocity, longer calculation time are expected, so in
0.00115 0.00888 0.00046 ¡0.00101 ¡0.00021 ¡0.00089 the present model an additional strategy was established to increase the
0.00146 0.00046 0.01070 ¡0.00029 ¡0.00112 ¡0.00102 strain velocity of the boundary as the mean stress increases in the model.
0.00159 0.00041
In this way, and for the high number of blocks and size of the model, it
¡0.00038 ¡0.00101 ¡0.00029 ¡0.00049
¡0.00040 ¡0.00021 ¡0.00112 ¡0.00049 0.00202 0.00077
¡0.00455 ¡0.00089 ¡0.00102 0.00041 0.00077 0.00522 was possible to fit the simulation time of each uniaxial test in one work
day.

4.2.3. Results of the numerical simulation


4.2.2. Numerical model description The final X-horizontal displacement distribution in the model is
A numerical model was built with the Itasca code 3DEC 7.0 (Itasca, shown in Fig. 20a. The right lateral boundary was 11.26 cm, which
2020), which uses the Distinct Element Method to simulate the me­ implies a horizontal-X strain of 0.113% in a 100 m side model. Fig. 20b
chanical behaviour of a fractured medium; that is, by a set of deformable shows the stress–strain curve as recorded in the model. The linear
blocks that interact with each other through the discontinuity planes. regression on the points of Fig. 20b provides a value of modulus of
The orientation and spacing of the 3 families of joints object of deformability equal to 483 MPa.
simulation were introduced in the model (Table 13). The model is pre­ The resulting values of the elastic moduli obtained in all directions
sented in Fig. 19a. The details of the blocks corresponding for each from the proposed analytical model and the simulation with 3DEC are
family of discontinuities is shown in: Fig. 19b family 1, with 0.5 m summarized in Table 14.
spacing; in Fig. 19c family 2, with 4 m spacing; and in Fig. 19d family 3, Comparing the results of the 3DEC numerical simulation with those
with 20 m spacing. from the analytical model leads to relative errors in the values of the
The size of the model is a critical aspect in computing capacity and is moduli lower than 6% thus validating the analytical model for general
conditioned by the family with the largest spacing, which is family 3. To cases when it is not possible to find a simplified theoretical solution.
limit computational times, dimensions of 100x100x100 m3 have been
adopted, to include at least 5 discontinuity planes of family 3 and thus 5. Comparative study with empirical formulations
obtain a representative simulation of the study problem. Furthermore,
since the blocks are deformable it is necessary to discretize them by In addition to the theoretical validation explained in section 4, the
means of a finite difference mesh, for which are used tetrahedral ele­ results of the analytic model were compared with the values obtained by
ments with a maximum side size of 2 m. means of some of the most widely used empirical formulations.

Table 13
Orientation and properties of the joints and intact rock present in the general case of validation.
Intact Rock Joints

Deformation modulus ER Poisson coefficient νR Set Dip direction Dip β Spacing s Thickness e Fill modulus EDi Fill Poisson
[MPa] [-] (◦ ) (◦ ) (m) (m) [kPa] coefficientνDi

2000 0.25 1 20 55 0.5 0.01 20,000 0.3


2 85 10 4 0.02 5000 0.35
3 280 80 20 0.05 1000 0.4

14
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Fig. 19. 3DEC model: a) General view; b) Blocks delimited by family 1; c) Blocks delimited by family 2; d) Blocks delimited by family 3.

Fig.20. Uniaxial test in the X direction: a) Final distribution of displacements; b) Stress–strain curve.

The empirical models considered in this assessment are all depen­


Table 14
dent on classification indexes, since their use is more common. More
Result comparison of deformability. Numerical validation in a general case.
specifically, an expression dependent on the RQD index (Zhang & Ein­
Analysed Deformation modulus E [MPa] Relative error εrel stein, 2004) was chosen, alongside three models dependent on the RMR
direction [%]
(Bieniawski, 1978; Serafim & Pereira, 1983; and Hoek et al., 2002). The
Proposed analytical 3DEC numerical Result
latter formulations can be applied to different rock masses, according to
model simulation comparison
their weathering degree and their strength. All of the equations and
Horizontal (x) 454 483 6.0% conditions are summarised in Table 15.
Ex
Horizontal (y) 628 668 6.0%
Ey 5.1. Assessed cases
Vertical (z) Ez 494 513 3.7%
Volumetric 305 322 5.3% Twenty-five cases were analysed to ensure that a more complete
assessment was obtained. These 25 cases consisted of rock masses with

15
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

different conditions; either one or two sets of joints, and different values joint fill. To cover a wider range of rock masses, each of the 25 cases
of the parameters defined under the classification by Bieniawski (1989). defined were studied with four different types of joint fill. These four
The analytic model presented in this paper was specifically defined for types of fill were characterised by different levels of stiffness, with the
the assessment of rock masses with filled joints (“type D”). For this reason, properties included in Table 21.
all 25 cases were completely weathered and filled discontinuities. At this point, all the parameters needed for the calculation of the
Considering this condition, the cases defined have RMR values of global deformability of the rock masses have been defined, thus enabling
either 10, 30, 50 or 70. No cases could be defined with higher values for the estimation of their behaviour with the analytic model, as well as
the RMR, due to their incompatibility with the condition of filled joints. with the empirical formulations.
All 25 cases are summarised in Tables 16-19, including their character­
istics and the corresponding RMR scores, according to the classification by 5.2. Result discussion
Bieniawski (1989). The deformation moduli obtained by means of the empirical
The orientation of the joint sets of each case can be derived based on formulae is summarised in Table 22 for all the 25 cases assessed. The
the spacing and RQD values included in Tables 16-19. This can be done models included in Table 13 have been considered, thus obtaining
by means of the formulation developed by Priest & Hudson (1976), moduli dependent on the classification indexes RMR (ERMR ) and RQD
which is given below. (ERQD ). Given the nature of the indexes involved, the former moduli
0.1λθ were averaged for the rock mass, while the latter were estimated in the
RQD(%) = 100⋅e− ⋅(0.1λθ + 1)
direction of determination of the RQD.
One set of joints : λθ = λ⋅cosθ; λ = 1/s However, the analytical formulation presented in this paper enables
the obtention of deformation moduli in three different directions. Thus,
Two sets of joints : λθ = λ1 ⋅cosθ + λ2 ⋅|cos(ϕ − θ) | the volumetric deformation modulus (Evol ) has to be estimated based on
λi = 1/si ; i = 1, 2 these results, in order to have a value comparable to the moduli obtained
from the empirical equations i.e. averaged moduli.
Where:
First, a rock mass with isotropic behaviour was considered, with
RQD: Rock Mass Rating (%) calculated in the direction of the bore­
deformation modulus E and Poisson coefficient ν. When bearing an
hole, defined by angle θ.
isotropic load p, this rock mass had a volumetric deformation (εv ):
λi : Inverse value of the mean spacing of a certain joint set, in the
direction i orthogonal to the family orientation. 3p(1 − 2ν)
εv = (27)
λθ : Inverse value of the mean spacing of the rock mass joint sets in the E
direction of the borehole. On the other hand, an orthotropic rock mass with moduli E1 , E2 , E3
si : Mean spacing of the i joint set, in meters (m). had a different response to the volumetric deformation (εv ), which is:
θ: angle between the direction of the borehole and the direction ( )
orthogonal to the planes of joint set 1. εv = p(1 − 2ν)
1
+ +
1 1
(28)
ϕ: angle between the directions of the two sets of joints, when E1 E2 E3
present. Therefore, considering the volumetric deformations obtained from
The geometrical distribution of all the joint sets and directions equations (27) and (28) to be equal, the equivalent deformation modulus
involved in this calculation are shown in Fig. 21. According to this of an anisotropic rock mass (Evol ) can be obtained. The result is the har­
definition, the orientation of the joint sets can be calculated for the 25 monic mean of the deformation moduli in each of the three directions:
cases.
Based on the formulation by Priest & Hudson (1976), the informa­ 3 1 1
= + +
1
tion of the joint structure of the rock masses was obtained and sum­ Evol E1 E2 E3
marised in Table 20. The parameter d included in the table, was The equivalent moduli Evol thus obtained for the assessment cases are
calculated as per its definition in equation (5), and determined by summarised in Table 23, obtained as a mean of the result of the analytic
assuming a joint thickness of e = 10 mm in all cases. model. The four different types of joint fill defined in Table 19 were
The deformation properties of the intact rock was assumed to be a considered, with their corresponding fill deformation moduli (ED ).
Poisson coefficient of 0.25 and a deformation modulus of 100 GPa. The When analysing the deformation moduli estimated by means of the
latter corresponds to the value obtained considering a value of the RMR analytic model, it becomes apparent that there is a huge difference when
of 100 (intact rock) in the empirical formulation of Bieniawski (1978). considering the various types of joint fill. This underlines the great
Only one factor remains undefined: the deformation properties of the importance of this specific aspect when assessing the global deforma­
tional behaviour of the rock mass.
The moduli obtained from the empirical formulations (ERMR , ERQD )
Table 15 and the mean harmonic of the analytic values (Evol ) are compared in
Empirical formulations used in the comparative study with the deformation Fig. 22. This comparison highlights the already mentioned scatter
moduli obtained by the analytic model. depending on the type of fill, but also the great difference between the
Citation Formula Conditions values estimated by means of the RQD and the rest of the results. The
Bieniawski (1978) ERM (GPa) = 2RMR − 100 RMR > 50 reason behind this difference has already been mentioned; the RQD is an
Serafim & Pereira ERM (GPa) = 10((RMR− 10)/40 ) RMR < 50 index with a strong directional component. As a result, an accurate
(1983) analysis requires the distinction of two different types of results: the
Hoek et al (2002)
ERM (GPa) =
RMR < 50 mean moduli (to be compared with the empirical expressions dependent
UCS < 100
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ on the RMR) and the original directional moduli of the analytic model
RCS(MPa) (RMR− 10)/40 MPa
(1 − D/2) ⋅10 (to be compared with the empirical moduli based on the RQD).
100
Zhang & Einstein ERM (GPa) = ER ⋅10(0.0186RQD− 1.91) RQD, ER
(2004) 5.2.1. Comparison with the deformation moduli obtained from the RMR
ERM is the deformation modulus of the rock mass The deformation moduli obtained by means of the empirical for­
UCS is the uniaxial compressive strength of the intact rock mulations based on the RMR (ERMR ) and the moduli resultant of aver­
ER is the deformation modulus of the intact rock aging the values calculated by the analytical model (Evol ) are compared
D is the Hoek weathering index in Fig. 23. The ordinate axis is presented in logarithmic scale to make the

16
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Table 16
Summary of conditions of the rock mass and RMR scores according to the classification by Bieniawski (1989). Cases with RMR = 10.
Case UCS [MPa] RQD [%] Spacing [mm] State of the joint Water RMR

Joint set 1 Joint set 2

1 <1 35 60 60 Flow
RMR score 0 5 5 5 0 0 10
2 <1 35 20 200 Flow
RMR score 0 5 2 8 0 0 10
3 <1 50 20 Flow
RMR score 0 8 2 0 0 10
4 <1 50 20 20 Flow
RMR score 0 8 2 2 0 0 10
5 50 25 40 40 Flow
RMR score 4 3 3 3 0 0 10
6 50 25 50 20 Flow
RMR score 4 3 4 2 0 0 10
7 50 15 40 60 Flow
RMR score 4 2 3 5 0 0 10

Table 17
Summary of conditions of the rock mass and RMR scores according to the classification by Bieniawski (1989). Cases with RMR = 30.
Case UCS [MPa] RQD [%] Spacing [mm] State of the joint Water RMR

Joint set 1 Joint set 2

8 25 100 200 Flow


RMR score 2 20 8 0 0 30
9 25 100 200 200 Flow
RMR score 2 20 8 8 0 0 30
10 100 100 40 Flow
RMR score 7 20 3 0 0 30
11 100 90 100 Flow
RMR score 7 17 6 0 0 30
12 100 90 100 100 Flow
RMR score 7 17 6 6 0 0 30
13 100 75 2000 60 Flow
RMR score 7 13 15 5 0 0 30

Table 18
Summary of conditions of the rock mass and RMR scores according to the classification by Bieniawski (1989). Cases with RMR = 50.
Case UCS [MPa] RQD [%] Spacing [mm] State of the joint Water RMR

Joint set 1 Joint set 2

14 100 90 2.5 20 Dry


RMR score 7 17 20 2 0 15 50
15 100 100 200 Dry
RMR score 7 20 8 0 15 50
16 100 100 200 200 Dry
RMR score 7 20 8 8 0 15 50
17 250 60 2000 60 Dry
RMR score 15 10 15 5 0 15 50
18 250 80 60 Dry
RMR score 15 15 5 0 15 50
19 250 80 60 60 Dry
RMR score 15 15 5 5 0 15 50
20*1 250 100 2000 Flow
RMR score 15 20 15 0 0 50
21*2 250 100 2000 Flow
RMR score 15 20 15 0 0 50
22*3 250 100 2000 Flow
RMR score 15 20 15 0 0 50

*1 One set of joints. Dipβ = 0


*2 One set of joints. Dipβ = 45


*3 One set of joints. Dipβ = 90


results clearer. The analytic results are shown alongside their expo­ Likewise, the scatter of the analytic results regarding the mathe­
nential fit and the empirical results, in four different charts corre­ matical fit line was observed to increase for higher values of the RMR.
sponding to the four different types of joint fill. The only exception to this trend was in the points with RMR = 70, and
The results summarised in Fig. 23 lead to a number of conclusions. the reason was the low number of analysed cases. In addition to this,
First, the laws that relate the deformation moduli and the RMR show an there was the fact that the scatter happens independently of the stiffness
exponential tendency, depicted as linear in the semi-logarithmic charts. of the type of joint fill.

17
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Table 19
Summary of conditions of the rock mass and RMR scores according to the classification by Bieniawski (1989). Cases with RMR = 70.
Case UCS [MPa] RQD [%] Spacing [mm] State of the joint Water RMR

Joint set 1 Joint set 2

23*1 > 250 100 > 2000 Dry


RMR score 15 20 20 0 15 70
24*2 > 250 100 > 2000 Dry
RMR score 15 20 20 0 15 70
25*3 > 250 100 > 2000 Dry
RMR score 15 20 20 0 15 70

*1 One set of joints. Dipβ = 0


*2 One set of joints. Dipβ = 45


*3 One set of joints. Dipβ = 90


(slope of the trend line) is practically the same in all cases and is similar
to the one of the empirical formulations.
All these results bring under the spotlight the immense influence of
the joint stiffness in the global behaviour of the rock mass. When
observing Fig. 23d it is clear that the empirical formulation is closest to
the analytic results in the cases where the joint fill is stiffest. In order to
understand this, it must be noted that the empirical formulations that
depend on the RMR are meant to be used in rock masses with non-filled
joints. This reflects the fact that their expressions are fitted with data
obtained from tests done in rock specimens in which the joints were not
filled and were, as a result, very stiff.
Another matter already mentioned is that these empirical formulae
are limited to the assessment of average deformability, since they cannot
distinguish different directions of study.
When considering the aforementioned issues together they demon

Fig. 21. Projection of the spacing of joint sets into the direction of the borehole
(RQD direction). Table 21
Deformability parameters of the joint fill for the 25 cases assessed.

On the other hand, it can be seen that the mean values obtained from Fill Deformation modulus ED (MPa) Poisson Coefficient νD (-)
the analytic model are generally lower than the ones derived from the Very soft 1.0 0.25
empirical laws, independently of the stiffness of the joint fill. However, Soft 10.0 0.25
it must be pointed out that the analytic moduli increase for higher values Medium 50.0 0.25
of the joint fill stiffness and that the exponential increase with the RMR Hard 100.0 0.25

Table 20
Orientation parameters of the discontinuities defined for the 25 cases of analysis.
Case N of joint sets RMR Spacing (m) Joint set 1 Joint set 2 d

s1 (m) s2 (m) Dip direction (◦ ) Dip (◦ ) Dip direction (◦ ) Dip (◦ )

1 2 10 0.06 0.06 0 25.28 180 64.7 0.333


2 2 10 0.02 0.2 0 69.51 180 20.5 0.55
3 1 10 0.02 – 0 70.39 – – 0.50
4 2 10 0.02 0.02 0 87.75 0 72.8 1.00
5 2 10 0.04 0.04 0 5 180 85 0.50
6 2 10 0.05 0.02 0 8.2 180 81.8 0.70
7 2 10 0.04 0.06 0 46.95 0 1.9 0.417
8 1 30 0.2 – 0 89.59 – – 0.05
9 2 30 0.2 0.2 0 90 0 90 0.10
10 1 30 0.04 – 0 90 – – 0.25
11 1 30 0.1 – 0 60 – – 0.10
12 2 30 0.1 0.1 0 89.05 0 59.1 0.20
13 2 30 2 0.06 0 33.49 180 56.5 0.172
14 2 50 2.5 0.02 0 5.65 180 84.4 0.504
15 1 50 0.2 – 0 89.63 – – 0.05
16 2 50 0.2 0.2 0 90 0 90 0.10
17 2 50 2 0.06 0 53.9 180 36.1 0.172
18 1 50 0.06 – 0 60.35 – – 0.167
19 2 50 0.06 0.06 0 75.68 0 75.7 0.333
20 1 50 2 – 0 0 – – 0.005
21 1 50 2 – 0 45 – – 0.005
22 1 50 2 – 0 90 – – 0.005
23 1 70 2.5 – 0 0 – – 0.004
24 1 70 2.5 – 0 45 – – 0.004
25 1 70 2.5 – 0 90 – – 0.004

18
­
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Table 22 Table 23
Deformation moduli obtained by means of empirical formulae for the 25 Equivalent deformation moduli (Evol) obtained from the presented analytic
assessed cases. formulation for the 25 assessed cases and different types of fill (ED).
Case RMR RQD UCS RMR formulation ERMR ERQD Equivalent deformation modulus, Evol [MPa]
[MPa] [GPa] [GPa]
Case Number of RMR Very soft Soft fill Medium fill Hard fill
1 10 35 0.5 Hoek et al (2002) 0.07 5.50 joint sets fill ED = ED = 10 ED = 50 ED = 100
2 10 35 0.5 Hoek et al (2002) 0.07 5.50 1 MPa MPa MPa MPa
3 10 50 0.5 Hoek et al (2002) 0.07 10.5
1 2 10 6.9 68.5 341.5 680.6
4 10 50 0.5 Hoek et al (2002) 0.07 10.5
2 2 10 4.6 45.8 228.4 455.3
5 10 25 50 Hoek et al (2002) 0.70 3.60
3 1 10 5.1 51.4 256.3 510.8
6 10 25 50 Hoek et al (2002) 0.70 3.60
4 2 10 3.9 38.7 192.5 383.1
7 10 15 50 Hoek et al (2002) 0.70 2.30
5 2 10 6.1 61.3 305.7 608.9
8 30 100 25 Hoek et al (2002) 1.60 89.1
6 2 10 3.9 39.3 196.2 391.2
9 30 100 25 Hoek et al (2002) 1.60 89.1
7 2 10 5.3 52.9 263.8 526.1
10 30 100 100 Serafim and Pereira 3.16 89.1
8 1 30 71.2 706.9 3436.4 6637.2
(1983)
9 2 30 36.5 363.9 1792.1 3517.0
11 30 90 100 Serafim and Pereira 3.16 58.1
10 1 30 14.2 142.2 706.5 1401.9
(1983)
11 1 30 20.6 205.3 1018.0 2014.8
12 30 90 100 Serafim and Pereira 3.16 58.1
12 2 30 13.4 134.0 666.1 1323.3
(1983)
13 2 30 11.4 113.6 565.5 1124.4
13 30 75 100 Serafim and Pereira 3.16 30.5
14 2 50 6.8 68.3 339.9 676.3
(1983)
15 1 50 71.2 706.9 3436.4 6637.2
14 50 90 100 Serafim and Pereira 10.0 58.1
16 2 50 36.5 363.9 1792.1 3517.0
(1983)
17 2 50 11.0 110.1 548.2 1090.5
15 50 100 100 Serafim and Pereira 10.0 89.1
18 1 50 12.4 124.1 617.5 1227.5
(1983)
19 2 50 9.6 95.7 476.6 948.2
16 50 100 100 Serafim and Pereira 10.0 89.1
20 1 50 707.2 6651.9 26087.0 41666.7
(1983)
21 1 50 358.7 3472.2 15306.1 26315.8
17 50 60 250 Serafim and Pereira 10.0 16.1
22 1 50 707.2 6651.9 26087.0 41666.7
(1983)
23 1 70 882.6 8174.4 30612.2 47619.0
18 50 80 250 Serafim and Pereira 10.0 37.8
24 1 70 448.0 4310.3 18292.7 31250.0
(1983)
25 1 70 882.6 8174.4 30612.2 47619.0
19 50 80 250 Serafim and Pereira 10.0 37.8
(1983)
20 50 100 250 Serafim and Pereira 10.0 89.1
(1983) While there are significant differences in the moduli shown in Fig. 26, the
21 50 100 250 Serafim and Pereira 10.0 89.1 tendency for the behaviour is similar for both empirical and analytical
(1983) values. In addition and similar to the assessment done in section 5.2.1, it
22 50 100 250 Serafim and Pereira 10.0 89.1 should be noted that the analytical deformation moduli increase with
(1983)
23 70 100 300 Bieniawski (1978) 40.0 89.1
stiffer joint fills, and are closest to the RQD-based empirical moduli for the
24 70 100 300 Bieniawski (1978) 40.0 89.1 hardest fill.
25 70 100 300 Bieniawski (1978) 40.0 89.1 An analogous assessment was done in Fig. 27, which includes the
deformation moduli based on the empirical formulae (ERQD ) and those
obtained with the analytical model, this time, in the direction orthog­
strate the great influence of the parameter d in equation (5), in the global
onal to the borehole (E90+θ ). The results in Fig. 27 show that these values
deformability of the rock mass. This joint parameter was calculated for
were clearly not suitable to be compared, given their orientation was
all 25 assessed cases, as can be seen in Table 20. The results obtained
completely different.
from the analytic model are shown in Fig. 24, which shows the evolution
The conclusion of this assessment was oriented towards the limita­
of the mean deformation moduli (Evol ) for variations of parameter d. The
tions of the Zhang & Einstein (2004) formula; their RQD-based formu­
correlations shown in Fig. 24 correspond to an exponential fit, and their
lation was clearly limited to the estimation of the deformation properties
correlation index is high at over 98%.
in the direction in which the RQD was calculated, and not being able to
The mean of the resulting analytic moduli can also be shown alongside
assess other orientations. In addition to this, the formulation was only
the deformation moduli of the empirical expressions. This is done in
suitable for the analysis of rock masses in which the joints were not
Fig. 25, where it can be seen that the exponential fit of the empirical
filled.
moduli is very similar to that of the analytic results for the cases with
hardest joint fill, thus corroborating the notes on the limitation of the
6. Scale effect
empirical formulae regarding the joint stiffness.
The study of the scale effect on deformability can be introduced in
5.2.2. Comparison with the deformation moduli obtained from the RQD
the analysis, by considering: the geometric scale, associated with the size
The fundamental difference between the deformation moduli ob­
of the application problem, and the intrinsic scale, associated with the
tained using the RMR and the RQD indexes has already been mentioned
roughness profile according to the type of joints of the rock mass.
at the beginning of section 5.2; the former have an averaged nature,
The scale of application for problems is illustrated in Fig. 28 (Wyllie
while the latter are strongly influenced by the direction. This explains
and Mah, 2004), where it is easy to see that increasing areas of influence
why the deformation moduli resultant of the analytic model, oriented in
involve new families according to their spacing. For deformability
the three directions of study, are closer to the RQD-dependent values
models with filled joints that can be characterized with linear elastic
than to those based on the RMR.
models, an analysis of the influence of the geometric scale was incor­
The deformation moduli obtained from the RQD-based empirical
porated for some case studies: developed in section 6.1.
formulation (ERQD ) and those from the analytic model in the direction of
Alternatively, to study the effect of the intrinsic scale, it is necessary
evaluation of the RQD (Eθ ) are compared in Fig. 26 for the four types of
to use constitutive models of the joint that include the morphology that
joint fill considered as per Table 21. It was assumed that the RQD was
define them, such as the Hertz-Mindlin model included in section 3.5.
calculated in the direction of the borehole, which is generally vertical.
Thus, in joints with little or no filler thickness (type A, B and C, Fig. 13),

19
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Fig. 22. Comparison of the equivalent deformation moduli obtained from both analytic and empiric models.

Fig. 23. Comparison of the equivalent deformation moduli obtained by means of the analytic model and the empirical formulations dependent on the RMR for a)
very soft fill (ED = 1 MPa); b) Soft fill (ED = 10 MPa); c) Medium fill (ED = 50 MPa); d) Hard fill (ED = 100 MPa).

the effect of the intrinsic scale can be introduced, as indicated in a simple different spacing, so their influence will be conditioned by the dimensions
example in section 6.2. of the construction. Thus, for large dimensions, such as dams, the three
families will condition the deformability of the rock mass; for smaller
6.1. Size scale dimensions, such as bridge foundations, the deformability of families 1
and 2 should be considered, with family 3 not being a determining factor
To analyze the influence on the deformability of the size of the rock due to their great spacing. Finally, local actions will only be influenced by
mass, we studied the rock mass described in Table 13 of section 4.2, discontinuities of family 1, with 0.5 m spacing.
composed of three families of filled discontinuities, of different spatial Fig. 29 shows the influence of the different scales of the problem on the
configuration, and elastic characteristics. Each of the three families has deformability of the rock mass. These represent the analytical results of

20
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Fig. 24. Equivalent deformation moduli obtained by the analytic model. Evolution with parameter “d” for all four types of fill.

Fig. 25. Comparison between the deformation moduli obtained with the empirical formulae (ERMR) and the equivalent moduli of the analytic model. Evolution with
parameter “d”.

Fig. 26. Comparison between the deformation moduli of empirical formulations (ERQD) and the results of the analytic model in the direction of the RQD. Results of
all 25 assessed cases and all types of fill.

21
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Fig. 27. Comparison between the deformation moduli of empirical formulations (ERQD) and the results of the analytic model in the direction orthogonal to the RQD.
Results of all 25 assessed cases and all types of fill.

families. The figure shows that for the three families together, large vol­
umes of rock masses will be necessary (100x100x100 m3) as calculated in
section 4.2, where a volumetric compressibility modulus equal to 305
MPa was obtained in the analytical model (Table 14).
Fig. 29 also represents the analytical results considering the maximum
compressibility of the fill for the three families of the rock mass (1 MPa)
and the results considering the minimum compressibility for all families
(20 MPa). The scale effect curves in both extreme cases obviously make it
possible to obtain the upper deformability envelope (for the case of more
rigid joints) and the lower envelope (for more deformable joints). It is
remarkable how both envelopes stabilize the modulus of compressibility
from a certain block size (between 1 and 100 m3) although the original
rock mass manifests a continuous decrease in all scales. This fact shows the
importance of a particularized study of the conditions of applicability of
geometrically simplified deformation models due, for example, to high
Fig. 28. Size effects in rock mass deformability (from Wyllie and Mah (2004)). computational costs in numerical models.

the proposed model according to the value obtained from the volumetric 6.2. Intrinsic scale
compressibility modulus considering family 1 (local scale), family 1 and 2
(bridge scale) or the three families (dam scale). The deformability is The different mineralogical characteristics of the rocks and the
considerably reduced as a function of a larger size of the rock mass geological conditions induce different morphologies in the walls of the
considered in the analysis, from the size of the laboratory corresponding to discontinuities. In this way, the different roughnesses and amplitudes of
the intact rock to that corresponding to the spacing of the different the contacts that are measured at the joints are different manifestations
of the scale effect that rock masses have inherently. In this regard, it is
possible to use empirical formulas for the correction of the scale effect
(Barton and Bandis, 1982) and graphs (Barton, 1982) where, depending
on the amplitude of the roughness, different geomechanic parameters
are associated as a function of the size of the joint.
To consider the intrinsic scale, it is necessary to account for the
geometric parameters that define the profile of the discontinuities at the
scale of the problem. Thus, this consideration is not possible in a linear
elastic model, which is more representative of joints with thick fill, but it
can be introduced in the Hertz-Mindlin model, as developed in section
3.5.
To demonstrate the importance of the scale, it is enough to perform
the analytical calculations with a simple model of a single family of
horizontal joints, as shown in Table 24.
To introduce the Hertz-Mindlin deformability model, we need to
identify the morphology of the roughness of the joint from a mean radius
of roughness (R) and a density of contacts per unit surface (ρ). By varying
both parameters with the values of laboratory and field profilometers (or
Fig. 29. Size effects in rock mass deformability. Calculation of the volumetric large-scale topographic survey), the proposed analytical model can be
compressibility modulus using the analytical model proposed for the rock mass applied, and the volumetric compressibility module obtained to study the
in Table 13, considering different stiffnesses of the filled joints: the original ones influence of the intrinsic scale. Table 25 includes values of mean radius
(generic rock mass), low compressibility of all the joints (20 MPa), and high and contact density, which were introduced to perform the analytical
compressibility for all joints (1 MPa). calculations. The density was chosen so that the proportion of the area

22
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

occupied by roughness was maintained, or the same as ρ⋅π⋅R2 = cte. For Table 25
this simple application case we adopted constant values of the geometric Characterization of joint roughness for Hertz-Mindlin deformability models.
parameters of the contacts, but a more detailed study can be made from Mean radius of Roughness R Density of contacts per surface unit ρ ρ⋅π⋅R2
distribution functions. (mm) (contacts/m2)
Fig. 30 represents the results obtained by means of the analytical 0.2 106 0.126
formulation with the Hertz-Mindlin model, which shows the deform­ 2 104 0.126
ability of the rock mass as a function of the size of the roughnesses. Since a 20 102 0.126
non-linear model is being used, the stress state influences the deform­ 200 1 0.126

ability of the model and the volumetric compressibility moduli was


calculated for 100 KPa and 1000 KPa. The curves in Fig. 30 both show the
markedly decreasing tendency of the deformability with an increasing size
of the contacts (with the same proportion of roughness area) which shows
that, under normal stress, the rock masses are extremely deformable the
more the joints are open. Also, as expected, a greater modulus of volu­
metric deformability was obtained at a higher stress level and with higher
values of roughness size.

7. Conclusions

The research carried out in the present paper confirms the adequate
capacities of the analytic model presented here. This new formulation
makes it possible to evaluate the deformation behaviour of a rock mass
based on its fundamental parameters, commonly obtained by means of
geomechanical field assessments.
The formulation presented here should be used for general analyses
regarding both the number of joint sets considered and the constitutive
Fig. 30. Intrinsic scale in rock mass deformability. Calculation of the volu­
equations considered for the definition of the stress–strain behaviour of
metric compressibility modulus using the analytical model proposed for the
intact rock and discontinuities. The specific scenario included in this rock mass in Table 24, considering different mean radius of roughness for a
paper involves only the cases with filled joints, in which the fill governs roughness area ratio in the joint of 12.6%.
the constitutive behaviour using both the linear elastic model and the
Hertz-Mindlin law.
- In the specific case of RMR-based empirical expressions, by comparing
The model was validated by means of a comparison with simple
the results with those of the analytic model show the necessity to
theoretical cases and a numerical method. In addition the results were
average the resulting moduli, thus calculating a volumetric deforma­
also compared to those of the most widely used empirical formulations
tion modulus. Otherwise, the results of both formulations would not be
within the framework of Rock Mechanics. Some of the conclusions
suitable for comparison. The reason behind this lies in the fact that the
drawn from this assessment are offered below:
empirical RMR formulations do not give oriented results, but average
moduli. These expressions are thus not valid for the study of rock
- The obtention of a general analytic constitutive formulation is
masses with anisotropic behaviour induced by their discontinuities.
possible. It can be done by means of the assemblage of the consti­
- Regarding the empirical formulations based on the RQD index, their
tutive matrices of the materials involved, based on the parameters
results are appropriately fitted when compared to those of the ana­
estimated in geomechanical field assessments.
lytic model, as long as the analytic moduli chosen for the analysis are
- The formulation is developed and some results are presented for both
oriented in the direction in which the RQD was calculated. The
linear elastic and Hertz-Mindlin models, and thus able to consider
second condition for the comparison to show good results is that the
the behaviour of the discontinuities governed by the filled material
rock mass must have stiff discontinuities. In conclusion, it must be
or by the contacts of the joint walls.
noted that the RQD-dependent formulations are not suitable for the
- The developed analytic model was validated and shown to give
obtention of volumetric deformation moduli.
suitable results both with simple theoretical formulations and with
- The mentioned limitations of the empirical formulations put under
the Distinct Element Numerical Method.
the spotlight the necessity to complement the deformability assess­
- Based on the proposed analytical model, a discussion is presented
ment by means of both RQD-based and RMR-dependent formula­
regarding the scale effect due to both the size of the application and
tions. It also underlines the need that other models can and have to
the intrinsic composition of the rock mass.
be applied, with the formulation presented in this paper as one
- Through comparison with empirical formulae, the model was shown
possible choice.
to give a good fit for stiff discontinuities. From this fact, it can be
ascertained that the most common empirical models lose accuracy
when analysing rock masses in which the discontinuities are not stiff Declaration of Competing Interest
or have no fill. The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.

Table 24
Orientation and properties of one horizontal joint set and intact rock to study the influence of the intrinsic scale in the deformability.
Intact Rock Joints

Deformation modulus ER Poisson coefficient Set Dip direction Dip β Spacing s Thickness e Elactic modulus, joint wall Poisson coefficient, joint
[MPa] νR [-] (◦ ) (◦ ) (m) (m) EDi [kPa] wall νDi

1000 0.25 1 45 0 0.5 0.05 10,000 0.25

23
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

References Gerrard, C., 1982c. Joint compliances as a basis for rock mass properties and the design
of supports. International Journal of Rock Mechanics and Mining Sciences &
Geomechanics Abstracts 19 (6), 285–305.
Agharazi, A., Martin, C., Tannant, D., 2012. A three-dimensional equivalent continuum
Ghamgosar, M., Fahimifar, A., & Rasouli, V. (2010). Estimation of rock mass deformation
constitutive model for jointed rock masses containing up to three random joint sets.
modulus from laboratory experiments in Karun dam. Eurock 2010.
Geomechanics and Geoengineering 7 (4), 227–238.
Gokceoglu, C., Sonmez, H., Kayabasi, A., 2003. Predicting the deformation moduli of
Ajalloeian, R., Mohammadi, M., 2013. Estimation of limestone rock mass deformation
rock masses. International Journal of Rock Mechanics & Mining Sciences 40,
modulus using empirical equations. Bulletin of Engineering Geology and the
701–710.
Environment 73 (2), 541–550.
Goodman, R., 1976. Methods of geological engineering in discontinuous rocks. West Pub.
Alemdag, S., Gurocak, Z., Gokceoglu, C., 2015. A simple regression based approach to
Co.
estimate deformation modulus of rock masses. Journal of African Earth Sciences 110,
Grimstad, E., Barton, N., 1993. Updating of the Q-System for NMT. Proceedings of
75–80.
International Symposium on Sprayed Concrete - Modern Use of Wet Mix Sprayed Concrete
Alemdag, S., Gurocak, Z., Cevik, A., Cabalar, A.F., Gokceoglu, C., 2016. Modeling
for Underground Support. Norwegian Concrete Association, Fagernes.
deformation modulus of a stratified sedimentary rock mass using neural network,
Hart, R.D., Cundall, P.A., Lemos, J., 1988. Formulation of a three-dimensional distinct
fuzzy inference and genetic programming. Engineering Geology 203, 70–82.
element model - Part II. Mechanical calculations for motion and interaction of a
Amadei, B. (1983). Rock anisotropy and the theory of stress measurements. Lecture notes
system composed of many polyhedral blocks. International Journal of Rock
in engineering, C. A. Brebbia and S. A. Orszag. Berlin: Springer.
Mechanics and Mining Sciences & Geomechanics Abstracts 25 (3), 117–125.
Amadei, B., & Goodman, R. E. (1981). A 3-D constitutive relation for fractured rock
Hertz, H., 1882. Über die Berührung fester elastischer Körper. Journal für die reine und
masses. Proceedings of the international symposium on mechanical behavior of
angewandte Mathematik 92, 156–171.
structured media, (págs. 249–268). Ottawa.
Hoek, E., Brown, E., 1997. Practical estimates of rock mass strength. International
Amadei, B., & Savage, W. Z. (1993). Effect of joints on rock mass strength and
Journal of Rock Mechanics and Mining Sciences 34 (8), 1165–1186.
deformability. En Comprehensive rock engineering— Principle, practice and projects
Hoek, E., Brown, E.T., 1980. Empirical strength criterion for rock masses. Journal of
vol. 1 (págs. 331–365). Pergamon, London: J. A. Hudson.
Geotechnical and Geoenvironmental Engineering 106, 1013–1035.
Aydan, Ö., Ulusay, R., Kawamoto, T., 1997. Assessment of rock mass strength for
Hoek, E., Diederichs, M.S., 2006. Empirical estimation of rock mass modulus.
underground excavations. International Journal of Rock Mechanics and Mining
International Journal of Rock Mechanics & Mining Sciences 43, 203–215.
Sciences 34 (3–4), 18.e1–18.e17.
Hoek, E., Carranza-Torres, C., & Corkum, B. (2002). Hoek-Brown failure criterion - 2002
Barton, N., 1973. A review of the shear strength of filled discontinuities in rock.
Edition. Proceedings NARMS-TAC conference, Toronto, 1, 267-273.
Bergmekanikkdag, Oslo.
Hönisch, K. (1993). Keynote lecture: Conclusions from 100 constructed power caverns
Barton, N. (1982). Modelling rock joint behaviour from in situ block tests: implications
for future planning. Proceedings - Safety and Environmental Issues in Rock
for nuclear waste repository design. Office of Nuclear Waste Isolation, Columbus,
Engineering (págs. 1013-1027.). Lisbon, Portugal: Eurock 93.
OH, p 96, ONWI-308.
Hu, K., Huang, Y., 1993. Estimation of the elastic properties of fractured rock masses.
Barton, N., 1983. Application of Q system and index tests to estimate shear strength and
International Journal of Rock Mechanics and Mining Sciences & Geomechanics
deformability of rock masses. International Symposium on Engineering Geology and
Abstracts 30 (4), 381–394.
Underground Construction, Lisbon, Portugal, pp. 51–70.
Huang, T.H., Chang, C., Yang, Z.Y., 1995. Elastic moduli for fractured rock mass. Rock
Barton, N., 2007. Near-surface gradients of rock quality, deformation modulus, Vp and
Mechanics and Rock Engineering 28 (3), 135–144.
Qp to 1 km depth. First Break 25, 53–60.
Isik, N., Ulusay, R., Doyuran, V., 2008. Deformation modulus of heavily jointed-sheared
Barton, N., Bandis, S., 1982. In: Effects of block size on the shear behavior of jointed
and blocky greywackes by pressuremeter tests: Numerical, experimental and
rock. American Rock Mechanics Association:, Richardson, TX, USA, pp. 739–760.
empirical assessments. Engineering Geology 101 (3–4), 269–282.
Bashari, A., Beiki, M., Talebinejad, A., 2011. Estimation of deformation modulus of rock
Itasca. (2020). Advanced, three dimensional discrete element modeling for geotechnical
masses by using fuzzy clustering-based modeling. International Journal of Rock
analysis of rock (3DEC) (Version version 7.0). Minneapolis (USA): Itasca Consulting
Mechanics & Mining Sciences 48, 1224–1234.
Group.
Beiki, M., Bashari, A., Majdi, A., 2010. Genetic programming approach for estimating the
ISRM (International Society for Rock Mechanics) (1978). Suggested Methods for the
deformation modulus of rock mass using sensitivity analysis by neural network.
Quantitative Description of Discontinuities in Rock Masses. International Journal of
International Journal of Rock Mechanics & Mining Sciences 47, 1091–1103.
Rock Mechanics and Mining Sciences & Geomechanics Abstracts, 15, 319-368.
Bieniawski. (1978). Determining Rock Mass Deformability: Experience from Case
Jasarevic, I., & Kovacevic, M. S. (1996). Analyzing applicability of existing classification
Histories. International Journal of Rock Mechanics, Mining Science & Geomechanics
for hard carbonate rock in mediterranean area. Proceedings of EUROCK’96, Turin,
Abstracts, Vol 15, 237-247.
Italia, 811-818.
Bieniawski, 1989. Engineering Rock Mass Classifications: A Complete Manual for
Jiang, Q., Cui, J., Feng, X.-T., Zhang, Y.-H., Zhong, S., Ran, S.-G., 2017. Demonstration of
Engineers and Geologists in Mining, Civil and Petroleum Engineering. John Wiley &
spatial anisotropic deformation properties for jointed rock mass by an analytical
Sons, New York.
deformation tensor. Computers and Geotechnics 88, 111–128.
Boyd, R.D., 1993. Elastic properties of jointed rock masses with regard to their rock mass
Jiang, X.-W., Wan, L., Wang, X.-S., Wu, X., Zhang, X., 2009. Estimation of rock mass
rating value. Engineering Geology of Weak Rock 8, 329–336.
deformation modulus using variations in transmissivity and RQD with depth.
Budiansky, B., O’Connell, R.J., 1976. Elastic moduli of a cracked solid. International
International Journal of Rock Mechanics and Mining Sciences 46 (8), 1370–1377.
journal of Solids and Structures 12 (2), 81–97.
Kachanov, M., 1982. A microcrack model of rock inelasticity part I: Frictional sliding on
Carvalho, J., 2004. Estimation of rock mass modulus. Personal communication.
microcracks. Mechanics of Materials 1, 19–27.
Chun, B., Ryu, W., Sagong, M., Do, J., 2009. Indirect estimation of the rock deformation
Kang, S., Kim, H., Jang, B., 2013. Correlation of in situ modulus of deformation with
modulus based on polynomial and multiple regression analyses of the RMR system.
degree of weathering. RMR and Q-system. Environmental Earth Sciences 69 (8),
International Journal of Rock Mechanics and Mining Sciences 46 (3), 649–658.
2671–2678.
Chun, B.-S., Jung, S.-H., Lee, Y.-J., Ahn, K.-C., Shin, J.-K., 2006. An Evaluation of
Kayabasi, A., Gokceoglu, C., Ercanoglu, M., 2003. Estimating the deformation modulus
Empirical Prediction Equation for Deformation Modulus of Rock Masses by Field
of rock masses: a comparative study. International Journal of Rock Mechanics &
Measurements. Tunnel & Underground Space 3, vol. 16, 251–258.
Mining Sciences 40, 55–63.
Clerici. (1993). Indirect determination of the modulus of deformation of rock masses -
Kemeny, J., Cook, N., 1986. Effective moduli, non-linear deformation and strength of a
Case histories. Eurock 93, Lisboa, Portugal - Proceedings: Safety and Environmental
cracked elastic solid. International Journal of Rock Mechanics and Mining Sciences
Issues in Rock Engineering Vol 1, 509-517.
& Geomechanics Abstracts 23 (2), 107–118.
Coon, R.F., Merritt, A.H., 1970. Predicting in situ modulus of deformation using rock
Kulatilake, P., Wang, S., Stephansson, O., 1993. Effect of finite size joints on the
quality indices. Determination of the in-situ Modulus of Deformation of Rock. ASTM
deformability of jointed rock in three dimensions. International Journal of Rock
STP 477, 154–173.
Mechanics and Mining Sciences and 30 (5), 479–501.
Cui, J., Jiang, Q., Feng, X., Li, S., Gao, H., Li, S., 2016. Equivalent elastic compliance
Kulhawy, F.H., Ingraffea, A.R., 1978. Geomechanical Model for Settlement of Long Dams
tensor for rock mass with multiple persistent joint sets: Exact derivation via modified
on Discontinuous Rock Masses. ISRM International Symposium. International Society
crack tensor. Journal of Central South University 23 (6), 1486–1507.
for Rock Mechanics and Rock Engineering, Rio de Janeiro.
Diederichs, M.S., Kaiser, P.K., 1999. Stability of large excavations in laminated hard rock
Li, C., 2001. Technical note: A method for graphically presenting the deformation
masses: the voussoir analogue revisited. International Journal of Rock Mechanics
modulus of jointed rock masses. Rock Mechanics and Rock Engineering 34 (1),
and Mining Sciences 36, 97–117.
67–75.
Duncan, J.M., Goodman, R.E., 1968. Finite element analyses of slopes in jointed rock.
Lyakhovsky, V., Reches, Z., Weinberger, R., Scott, T., 1997. Non-linear elastic behaviour
Washington D.C.: U.S. Army Corps of Engineers Rep. TR No, 1–68.
of damaged rocks. Geophysical Journal International 130 (1), 157–166.
Fossum, A.F., 1985. Effective elastic properties for a randomly jointed rock mass.
Mehrotra, V.K., 1992. Estimation of engineering parameters of rock mass. University of
International journal of Rock Mechanics and Mining Sciences 22 (6), 467–470.
Roorkee, Roorkee, India.
Galera, J. M., Álvarez, M., & Bieniawski, Z. T. (2005). Evaluation of the deformation
Mindlin, R.D., 1949. Compliance of Elastic Bodies in Contact. Applied Mechanics Trans.
modulus of rock masses: comparison of pressuremeter and dilatometer tests with
ASME 16, 259–268.
RMR prediction. ISP5-PRESSIO 2005 International Symposium, (págs. 1-25). Paris.
Mitri, H. S., Edrissi, R., & Henning, J. (1994). Finite element modelling of cable-bolted
Gerrard, C., 1982a. Elastic models of rock masses having one, two and three sets of joints.
stopes in hard rock ground mines. SME Annual Meeting, (págs. 94–116).
International Journal of Rock Mechanics and Mining Sciences & Geomechanics
Alburquerque.
Abstracts 19 (1), 15–23.
Morland, L., 1976. Elastic anisotropy of regularly jointed media. Rock mechanics 8 (1),
Gerrard, C., 1982b. Equivalent elastic moduli of a rock mass consisting of orthorhombic
35–48.
layers. International Journal of Rock Mechanics and Mining Sciences &
Muñiz Menéndez, M., Linares, H., Pardo, F., & Montero, N. (2013). Evaluation of the
Geomechanics Abstracts 19 (1), 9–14.
Deformation Modulus of Algeciras Flysch Unit by means of Pressuremeter Tests:

24
N. Esteban et al. Computers and Geotechnics 136 (2021) 104111

Correlation with RMR. 18th International Conference on Soil Mechanics and Singh, B., Villadkar, M., Samadhiya, N., Mehrotra, V., 1997. Rock mass strength
Geotechnical Engineering. Parallel session ISP-6. Paris. parameters mobilised in tunnels. Tunnelling and Underground Space Technology 12
Nassir, M., Settari, A., Wan, R., 2013. Joint Stiffness and Deformation Behaviour of (1), 47–54.
Discontinuous Rock. Journal of Canadian Petroleum Technology 49. Singh, R., & Bhasin, R. (1996). Q-System and deformability of rock masses. Proceedings
Nejati, H., Ghazvinian, A., Moosavi, A., Sarfarazi, V., 2013. On the use of the RMR system of conference on recent advances in tunnelling technology, New Delhi, India, 57-67.
for estimation of rock mass deformation modulus. Bulletin of Engineering Geology Sonmez, H., Gokceoglu, C., Ulusay, R., 2004. Indirect determination of the modulus of
and the Environment 73 (2), 531–540. deformation of rock masses based on the GSI system. International Journal of Rock
Nicholson, G.A., Bieniawski, Z.T., 1990. A nonlinear deformation modulus based on rock Mechanics and Mining Sciences 41 (5), 849–857.
mass classification. International Journal of Mining Geological Engineering 8, Sonmez, H., Gokceoglu, C., Nefeslioglu, H.A., Kayabasi, A., 2006. Estimation of rock
181–202. modulus: For intact rocks with an artificial neural network and for rock masses with
Oda, M., 1988. An experimental study of the elasticity of mylonite rock with random a new empirical equation. International Journal of ock Mechanics & Mining Sciences
cracks. International Journal of Rock Mechanics and Mining Sciences & 43, 224–235.
Geomechanics Abstracts 25, 59–69. Trunk, U., & Floss, R. (1991). Verwendung von vorinformationen bei der planung von
Oda, M., 1993. Modern developments in rock structure characterization. Comprehensive krafthauskavernen im fels. Proceedings of 7th International Congress in Rock
rock engineering. 1, 185–200. Mechanics, (págs. 1223-1226). Aachen: A.A. Balkema.
Oda, M., Kenichiro, S., Maeshibu, T., 1984. Elastic compliance for rock-like materials Ván, P., & Vasarhelyi, B. (2010). Relation of rock mass characterization and damage.
with cracks. Soils and Foundations 24 (3), 27–40. Rock Engineering in Difficult Ground Conditions (Soft Rock and Karst) (págs. 399-
O’Neill, M., Reese, L., 1999. Drilled shafts: construction procedures and design methods 404). Cavtat: Taylor & Francis Group.
FHWA-IF-99-025. Federal Highway Administration, U.S. Department of Stephens, R.E., Banks, D.C., 1989. Moduli for deformation studies of the foundation and
Transportation, Washington DC. abutments of the Portugues Dam—Puerto Rico. In: Rock Mechanics as a Guide for
Palmström, A., Singh, R., 2001. The deformation modulus of rock masses - comparisons Efficient Utilization of Natural Resources: Proceedings of the 30th US Symposium,
between in situ tests and indirect estimates. Tunnelling and Underground Space Morgantown. Balkema, Rotterdam, pp. 31–38.
Technology 3 vol 16, 115–131. Vásárhelyi, B., Kovács, D., 2017. Empirical methods of calculating the mechanical
Priest, S.D., Hudson, J.A., 1976. Discontinuity spacings in rock. International Journal of parameters of the rock mass. Periodica Politechnica Civil Engineering 61 (1), 39–50.
Rock Mechanics and Mining Sciences & Geomechanics Abstracts 13 (5), 135–148. Verman, M., Singh, B., Viladkar, M.N., Jethwa, J.L., 1997. Effect of tunnel depth on
Ramamurthy, T., 2004. A geo-engineering classification for rocks and rock masses. modulus of defomation of rock mass. Rock Mechanics and Rock Engineering 30 (3),
International Journal of Rock Mechanics and Mining Sciences 41 (1), 89–101. 121–127.
Read, S., Perrin, N. D., & Richards, L. R. (1999). Applicability of the Hoek - Brown Vychytil, J., & Horii, H. (1998). Micromechanics-based continuum model for hydraulic
Failure Criterion to New Zealand greywacke rocks. Rock Mechanics: Proceedings of fracturing of jointed rock masses during HDR stimulation.
the 9th International Congress. Paris, France. Wyllie, D.C., Mah, C.W. (2004). Rock Slope Engineering: Fourth Edition. The Institute of
Reddish, D.J., Mohammad, N., Stace, R., 1997. The Relation between in-situ and Mining and Metallurgy, Abingdon, Oxford, England OX14 4RN.
laboratory Rock Properties used in Numerical Modelling (Technical Note). Yang, J.P., Chen, W., Yang, D., Tian, H.M., 2016. Estimation of Elastic Moduli of Non-
International Journal of Rock Mechanics and Mining Sciences 34 (2), 289–297. persistent Fractured Rock Masses. Rock Mechanics and Rock Engineering 49 (5),
Rinaldi, A., Krajcinovic, D., Mastilovic, S., 2006. Statistical damage mechanics- 1977–1983.
constitutive relations. Journal of theoretical and applied mechanics 44, 585–602. Yang, K., 2006. Analysis of laterally loaded drilled shafts in rock. University of Akron.
Salamon, M., 1968. Elastic moduli of a stratified rock mass. International Journal of Rock Yoshinaka, R., Yamabe, T., 1986. Joint stiffness and the deformation behaviour of
Mechanics and Mining Sciences & Geomechanics Abstracts 5 (6), 519–527. discontinuous rock. International Journal of Rock Mechanics and Mining Sciences &
Sanei, M., Rahmati, A., Faramarzi, L., Goli, S., & Mehinrad, A. (2013). Estimation of rock Geomechanics Abstracts 23 (1), 19–28.
mass deformation modulus in Bakhtiary Dam Project in Iran. Rock Characterisation, Zhang, L., 2010. Method for Estimating the Deformability of Heavily Jointed Rock
Modelling and Engineering Design Methods. Proceedings of the 3rd ISRM Masses. Journal of Geotechnical and Geoenvironmental Engineering 136 (3).
SINOROCK 2013 Symposium. Zhang, L., Einstein, H.H., 2004. Using RQD to estimate the deformation modulus of rock
Serafim, J. L., & Pereira, J. P. (1983). Considerations on the geomechanical classification masses. International Journal of Rock Mechanics and mining Sciences 46,
of Bieniawski. Proceedings of the Symposium on Engineering Geology and 1370–1377.
Underground Openings, Lisboa, Portugal, 1133-1144. Zhang, Y., He, J., Wei, Y., & Nie, D. (2011). Prediction Research of Deformation Modulus
Serrano, A. A., & Soriano, A. (1986). Numerical model for jointed media. 2nd of Weak Rock ZOne under In-situ conditions. Science Press and Institute of Mountain
International Symposium on Numerical Models in Geomechanics. Ghent. Hazards and Environment.
Shen, J., Karaku, M., Xu, C., 2012. A comparative study for empirical equations in Zienkiewicz, O., Pande, G., 1977. Time-dependent multilaminate model of rocks—a
estimating deformation modulus of rock masses. Tunelling and Underground Space numerical study of deformation and failure of rock masses. International Journal for
Technology. Numerical and Analytical Methods in Geomechanics 1 (3), 219–247.
Shi, G.-H., 1988. Discontinuous deformation analysis - A new model for the static and
dynamics of block systems. Ph.D. Thesis. University of California, Berkeley.
Singh, B., 1973. Continuum characterization of jointed rock masses: Part II—Significance Further reading
of low shear modulus. International Journal of Rock Mechanics and Mining Sciences
& Geomechanics Abstracts 10 (4), 337–349. L’A.F.T.E.S. (2003). En Recommandations relatives à la caractérisation des massifs
rocheux utile à l’étude et à la réalisation des ouvrages souterrains. Annexe 15.
Tunnels et ouvrages souterrains 177.

25

You might also like