Geomechanics Applied To The Petroleum Industry
Geomechanics Applied To The Petroleum Industry
GEOMECHANICS
APPLIED TO
THE PETROLEUM
https://telegram.me/Geologybooks
INDUSTRY
2011
• CO2 Capture
Technologies to Reduce Greenhouse Gas Emissions
F. LECOMTE, P. BROUTIN, E. LEBAS
• Multiphase Production
Pipeline Transport, Pumping and Metering
J. FALCIMAIGNE, S. DECARRE
• Chemical Reactors
From Design to Operation
P. TRAMBOUZE, J.P. EUZEN
tional Energy Agency) and DOE (US Department of Energy), world energy demand will
still grow in the coming years and fossil energies will contribute in the range of 60 up to
80% to the energy supply, oil and gas contributing to about 50% of the energy supply. This
has two major implications. On one hand, oil and gas will have to be produced from more
and more difficult fields; on the other hand, more CO2 will be emitted and will have to be
captured and stored.
Both fluid production and CO2 storage involve fluid flow within porous rocks. This
induces pressure changes, saturation changes, temperature changes with water injection or
thermal recovery, even fluid interaction with the rock for CO2 injection. All these phenom-
ena have an impact on stresses and strains, or on the mechanical behaviour of rocks.
Rock mechanics in the oil and gas industry was traditionally used for drilling, wellbore
stability and fracturing. Reservoir application came with the observed consequences of sub-
sidence and the associated reservoir compaction. It should be stressed here that to satisfy the
increase in energy demand, fluids will be produced from complex environments where the
mechanical behaviour of the rock will be more and more important, for instance for wellbore
stability or faults behaviour.
A new area is emerging with CO2 management and more specifically CO2 sequestration.
Massive storage will occur by injecting in deep saline aquifers. Along with the description
of the reservoir behaviour, the knowledge of the overburden behaviour from a mechanical
standpoint will be essential to assess the storage integrity.
This book covers all these aspects and aims at providing a large range of potential users
with a simple approach of a broad field of knowledge. The reader will enjoy the tables of
reference values for the needed parameters such as wave velocity, elastic properties given
for both static and dynamic properties. Furthermore, the Figures are easy to understand and
illustrate very accurately the theoretical development. The book is application oriented, thus
closed forms, or empirical formulae are provided to give first estimates of the problem. For
instance, the reader will find indicators of drilling efficiency, closed forms for hydraulic or
thermal fracturing or for subsidence prediction. The practical consideration is always
nearby. Hence, sections are devoted to lab determination of properties, in situ determination
of stresses, sand production prevention and finally geomechanical monitoring of reservoirs.
While rock mechanics is of importance in many fields of application such as drilling,
production or reservoir engineering, it is still thought of as a complex field restricted to a
few experts. I hope and even guess that, through this book, any engineer involved in those
fields will understand the basics and the practical applications of rock mechanics.
Maurice Boutéca
Director
Resources Business Unit
IFP Energies nouvelles
https://telegram.me/Geologybooks
Table of Contents
https://telegram.me/Geologybooks
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . V
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . VII
List of Authors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . XV
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . XIX
Chapter 1
ELEMENTS OF ROCK MECHANICS FUNDAMENTALS
Chapter 2
GEOMECHANICS, DRILLING AND PRODUCTION
Chapter 3
GEOMECHANICS AND RESERVOIR
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
https://telegram.me/Geologybooks
1
When the sediments are deposited in lacustrine and marine environments, they are soft
and generally water saturated. They are converted into coherent rocks under the effect of
diagenesis, a term referring to several phenomena that can occur to varying extents depend-
ing on the deposition conditions and the type of sediment concerned (compaction, cementa-
tion, recrystallisation). Sedimentary rocks are brought to the surface during the large tec-
tonic movements which take place in the earth’s crust. When the ocean or continental plates
collide, the sediments are folded, deformed and uplifted to form valleys and high mountain
belts.
During these tectonic movements, the rock may have been subjected locally to very high
pressures and temperatures.
Sedimentary rocks are classified into four groups:
– Sandstones are detrital sedimentary rocks composed of grains of sand (quartz or
silica) bound together by a siliceous, calcareous or ferruginous cement which deter-
mines the strength of the sandstone. Quartz and feldspars are the main components.
Sandstone elements belong to the granulometric class of arenites (2 mm to 0.05 mm).
– Shales are defined as detrital rocks of fine grain size distribution (grain diameter less
than 2 mm). Phyllites (chlorite, montmorillonite, muscovite and kaolinite) are the
major component minerals, but quartz, feldspar and calcite are also present. Shales
have very complex properties.
– Carbonate rocks are composed mainly of the minerals calcite, aragonite and dolo-
mite, of origin either organic (marine organisms) or inorganic (chemical precipita-
tion). Calcite is the dominant mineral.
– Evaporites: these rocks are composed of elements precipitated from water. As the
water evaporates, the species in solution become more and more concentrated, up to
oversaturation then precipitation of the components. The main minerals precipitated
are gypsum, anhydrite and halite.
Chapter 1 • Elements of Rock Mechanics. Fundamentals 3
k ΔP
Q = −A (1.1.2)
μ L
The negative sign is needed since fluids flow from high to low pressure.
When the following units are used: Q in m3.s–1, A in m2, DP/L in Pa.m–1 and m in Poises,
the permeability k is expressed in m2 and in Darcy, with 1 Darcy = 0.97.10–12 m2. The
Darcy, denoted D, is widely used in petroleum engineering and petrophysics. The m2 is used
more by material physicists.
The hydraulic conductivity K, or permeability coefficient, is defined as the ratio of per-
meability to fluid viscosity:
k
K= (1.1.3)
μ
K is expressed in m.s–1. So, while the permeability k is a characteristic of the material
https://telegram.me/Geologybooks
that is highly dependent on the porosity, the hydraulic conductivity K is characteristic of the
conditions of flow in a given material for a given fluid. To a certain extent, the hydraulic
conductivity K takes into account the physical interactions between the fluid and the rock.
The permeability depends on the structure of the porous medium. Two rocks with the
same porosity may have quite different permeabilities. When the porous space is filled by
several fluids, the capillary effects act in addition to the dynamic head losses and, in this
case, permeabilities relative to the fluid considered can be determined.
Analysis of stresses and strains is essential whenever studying rock mechanics. Since these
notions have been studied extensively in numerous books [e.g. Jaeger and Cook, 1969], only
a few details will be given here.
1.1.2.1 Stresses
For the simple case of an axially loaded body, e.g., a cylindrical bar subjected to tension or
compression by a force passing through its centre, the stress s, or intensity of the distribu-
tion of internal forces, can be obtained by dividing the total tensile or compressive force
amplitude F by the cross-sectional area A over which it is acting. In this case the stress is a
scalar that represents an average stress savg over the area, meaning that the stress in the
cross section is uniformly distributed. We therefore obtain:
https://telegram.me/Geologybooks
F
σavg =
A
In general, however, the stress is not uniformly distributed over a cross section of a mate-
rial body, and consequently the stress at a point is different from the average stress over the
entire area. The local stress at a specific point in the body is defined as being the limiting
value of ΔF/ΔA when ΔA tends to zero.
If the force is not applied perpendicular to the cross-section A, the stress can be decom-
posed into a normal component called the normal stress and a tangential component called
the shear stress.
• Stress tensor
To provide a complete description of the stress state at a given point, we must be able to
identify the stresses relative to the surfaces in three orthogonal directions. There are 9 com-
ponents of the stress at a point, given by the tensor expression:
⎛σ τ τ ⎞
⎜ x xy xz ⎟
σ = ⎜ τ yx σ y τ yz ⎟
⎜ ⎟
⎜τ τ σ ⎟
⎝ zx zy z ⎠
• Invariants
Invariants remain unchanged during a change of coordinate system:
The mean normal stress is equal to the trace of the stress tensor matrix:
( )
σmean = σ x + σ y + σ z / 3 (1.1.4)
6 Chapter 1 • Elements of Rock Mechanics. Fundamentals
( )
I2 = − σ xσ y + σ yσ z + σ zσ x + τ xy
2 + τ2 + τ2
yz zx
(1.1.5)
I3 = σ xσ yσ z + 2τ xy τ yz τ zx − σ x τ 2yz − σ y τ zx
2 − σ τ2
z xy
• Deviatoric stress
The average stress defined previously expresses a uniform compression or extension. Dis-
tortions are expressed by the deviatoric stress tensor. Deviatoric stresses can be obtained by
subtracting the mean normal stress components from the stress tensor.
⎛s sxy sxz ⎞ ⎛ σ − σ τ xy τ xz ⎞
x
⎜ ⎟ ⎜ x mean ⎟
⎜ s s s ⎟ = ⎜τ σ − σ τ ⎟
⎜ yx y yz ⎟ ⎜ yx y mean yz ⎟
https://telegram.me/Geologybooks
⎜ s s s ⎟ ⎜τ τ zy σ z − σmean ⎟⎠
⎝ zx zy z ⎠ ⎝ zx
• Principal stresses
At any point of a body subjected to a system of forces, three planes can be defined on which
there are only normal stresses and no shear stresses. These planes are called the principal
planes and the stresses the principal stresses. In this coordinate system, the stress tensor has
a simple expression.
⎡σ 0 0⎤
⎢ 1 ⎥
⎢ 0 σ2 0⎥
⎢ ⎥
⎢⎣ 0 0 σ3 ⎥⎦
The three principal stresses are known as the major, intermediate and minor principal
stresses, in descending order of magnitude.
• Mohr circle
In 2D, knowing the principal stresses s1 and s2 (s1 > s2), the normal and shear stresses can
be calculated in any other plane inclined at an angle q in the xy plane (Figure 1.1.1).
Solving the equilibrium equations (sum of forces equal to zero) implies:
σθ =
1
2 1
( 1
) (
σ + σ2 + σ1 − σ2 cos 2θ
2
)
(1.1.6)
1
(
τθ = − σ1 − σ2 sin22θ
2
)
Chapter 1 • Elements of Rock Mechanics. Fundamentals 7
Direction of s2
sn
q
q
s1
Direction of s1
P
tn
s2
https://telegram.me/Geologybooks
Figure 1.1.1
State of stress at a point.
The stress state (sq, tq) in a plane at an angle q to the horizontal describes a circle in the
plane (s, t), called Mohr’s circle. Figure 1.1.2 shows a semi-circle of radius
1
(σ − σ2
2 1
)
1
(
and centre at σ1 + σ2 .
2
)
s1 – s2
2
tq
q 2q
s
s2 s 1 + s2 sq s1
2
Figure 1.1.2
Mohr circle.
8 Chapter 1 • Elements of Rock Mechanics. Fundamentals
tq q
https://telegram.me/Geologybooks
E Stress path
E
D
D
C
C
B
B
A sq p
A
(a) (b)
Figure 1.1.3
Representation of successive stress states, (example when s1 increases with s2 = s3
constant), (a) Mohr circles and (b) p-q diagram [Lambe and Whitman, 1969].
1.1.2.2 Strains
The material is deformed under the effect of stresses. The strains are calculated assuming
there has been a change between two states: the initial state and the final state. The strain of
a body can be measured by calculating the change in length of a line on the body or the angle
between two lines on this body. The change in length of a line, an elongation or a contrac-
tion, can be denoted as dl. The strain e, a dimensionless ratio, is defined as the ratio of elon-
gation with respect to the original length:
d l l - l0
e� � (1.1.8)
l0 l0
Chapter 1 • Elements of Rock Mechanics. Fundamentals 9
with
e strain in the direction measured,
l0 initial length of the line,
l current length of the line.
The local strain in the given direction can be expressed as:
dl
e � lim
l
� tgg � g
l
1 ⎛ ∂ui ∂u j ⎞
εij = ⎜ + ⎟ (1.1.9)
2 ⎜⎝ ∂x j ∂xi ⎟⎠
⎛ε ε ε ⎞
⎜ xx xy xz ⎟
ε = ⎜ ε yx ε yy ε yz ⎟
⎜ ⎟
⎜ε ε ε ⎟
⎝ zx zy zz ⎠
The tensor trace εv = εxx + ε yy + εzz expresses the volume strain. This value, independent
of the coordinate system chosen, is an invariant. Other strain invariants can be defined:
( )
J2 = − εx ε y + ε y εz + εz εx + εxy
2 + ε2 + ε2
yz zx
(1.1.10)
J3 = εx ε y εz + 2εxy ε yz εzx − εx ε2yz − ε y εzx
2 − ε ε2
z xy
As for the stresses, the strain tensor can be written with respect to its principal planes.
10 Chapter 1 • Elements of Rock Mechanics. Fundamentals
q'
p' Critical state line
Hvorslev surface
Roscoe surface
Normal consolidation
line
Tension failure
Figure 1.1.4
State boundary surface [Schofield and Wroth, 1968].
Chapter 1 • Elements of Rock Mechanics. Fundamentals 11
“stress-volume” states from the area where “stress-volume” states cannot exist. The states
presented by the parameters (p', q, v) always lie inside or on the SBS. The response of soils
inside the SBS is purely elastic (reversible but not necessarily linear) and becomes plastic
when the soil state reaches the state boundary surface. This notion has been extended to rock
mechanics. The SBS is also called yield surface.
strains εx and ε y ( εx = ε y for a cylindrical sample). The strains in this type of experiment
are related to the stresses by the following linear relation:
σz
εz = relation known as Hooke’s Law. (1.1.12)
E
E, called the modulus of elasticity or Young’s modulus, characterises the stiffness of the
sample, in other words its resistance to an axial stress. The radial deformability can also be
characterised using the Poisson’s ratio:
εx
ν=− (1.1.13)
εz
For isotropic materials, whose response is independent of the direction of the stresses
applied, the relations between the stresses and strains are given by the generalised Hooke’s Law:
εij =
1⎡
E⎣
( )
1 + ν σij − νδijσαα ⎤
⎦
or (1.1.14)
σij = λδij εαα + 2Gεij
where l and G, the Lamé parameters, are also elastic moduli. G is called the shear modulus,
since it characterises the resistance of the sample to shear. The shear modulus is the ratio of
shear stress to shear strain.
The bulk modulus K defines the ratio of an isotropic stress to a volume strain. In an iso-
σp
tropic stress state sx = sy = sz, the other components of the stress tensor are zero: K = .
εv
K is the sample resistance to an isotropic compression. The reciprocal 1/K, called compress-
ibility, is written Cp. The various elastic constants and their definitions are given in
12 Chapter 1 • Elements of Rock Mechanics. Fundamentals
Table 1.1.1. Note that for porous media, we must introduce Kd and Ku, bulk moduli, respec-
tively in drained and undrained conditions, Ed and Eu, drained and undrained Young’s mod-
uli, and nd and nu, drained and undrained Poisson’s ratios.
K E n G
Isotropic Uniaxial Uniaxial Shear sxy
Experiment compression sp compression sz compression sz
sp sz ex s xy
Definition K� E� n�- G�
ev ez ez exy
https://telegram.me/Geologybooks
The relations between the various elastic constants K, E, G, l and n are listed in
Table 1.1.2.
Table 1.1.2 Relations between the various elastic constants of an isotropic material [Birch, 1961].
K E n G l
2 3l � 2G l
___ ___
l� G
3
G
l �G �
2 l �G �
K -l l
___ 9K
3K - l 3K - l
3
2
�
K -l � ___
9 KG 3 K - 2G 2
___ ___
3K � G �
2 3K � G � K-
3
G
EG E E - 2G
___ ___ G
�
3 3G - E � 2G
-1
� 3G - E )�
___ ___ 3K - E 3 KE 3K - E
3K
6K 9K - E 9K - E
l
1� n
l
�1 � n ��1 - 2n � ___ l
1 - 2n ___
3n n 2n
G
�
2 1� n � �
2G 1 � n � ___ ___ G
2n
�
3 1 - 2n � 1 - 2n
___ �
3 K 1 - 2n � ___ 3K
1 - 2n
2 � 2n
3K
n
1� n
E E En
___ ___
�
3 1 - 2n � 2 � 2n �
1 � n 1 - 2n�� �
Chapter 1 • Elements of Rock Mechanics. Fundamentals 13
Gdyn
https://telegram.me/Geologybooks
VS = (1.1.16)
ρ
When the P and S wave velocities are available, the following dynamic moduli of elas-
ticity can be determined:
4 2
Gdyn = ρVS2 and Kdyn = ρVP2 − ρV (1.1.17)
3 S
The dynamic values of Young’s modulus and the Poisson’s ratio are given respectively by:
Edyn =
(
ρVS2 3VP2 − 4VS2 ) and νdyn =
VP2 − 2VS2
(1.1.18)
(VP2 − VS2 ) (
2 VP2 − VS2 )
The static moduli of rocks are generally lower than the dynamic moduli, due to the
amplitudes of the strains considered. In oil exploration, seismic data, Vertical Seismic Pro-
file (VSP) data and P- and S-wave logging data provide measurements of propagation veloc-
ities for rocks around and between the wells, both in the reservoir and in the surrounding
strata. These measurements are not recorded at the same frequency (Table 1.1.3), but data of
velocity dispersion as a function of frequency is interesting for petroacousticians.
Acquisition frequency
Exploration seismic 100-300 Hz
VSP 100-300 Hz
Sonic logging 100 kHz
Measurement on samples in laboratory 0.5 to 1 MHz
14 Chapter 1 • Elements of Rock Mechanics. Fundamentals
Seismic velocities are sensitive to lithology, porosity, fluid content, temperature and
stress field. Orders of magnitudes of the velocities Vp and Vs for some rocks are provided for
information in Table 1.1.4.
Table 1.1.4 Order of magnitude of P and S wave velocities and density for different rock types [Lavergne, 1986].
1.1.2.7 Anisotropy
If the elastic response of the material depends on its orientation for a given stress field, the
material is said to be anisotropic. In this configuration, the moduli of elasticity vary depend-
ing on the direction in which they are evaluated.
Anisotropy is defined as heterogeneities at a smaller scale than that of the volume
observed. At large scale, the development of geological strata (foliations, lineations, etc.)
results in a certain degree of anisotropy, while at small scale this anisotropy exists due to the
presence of the anisotropic minerals forming the rocks.
For anisotropic medium, a generalised form of Hooke’s law is written:
σij = Cijkl εkl or σ = C ε (1.1.19)
C , elastic stiffness tensor, is a 9x9 tensor. The symmetry of the stresses and strains
reduces the number of independent components to 36. In addition, the existence of a unique
energy potential reduces the number of independent components to 21, the maximum
number of elastic constants that a medium can possess.
Chapter 1 • Elements of Rock Mechanics. Fundamentals 15
Most sedimentary rocks (sandstone, mudstone, shale, etc.) exhibit anisotropic character-
istics in terms of strength and deformational properties (differences between elastic moduli
measured parallel and perpendicular to bedding), and are generally considered to be trans-
versely isotropic materials, symmetric about an axis normal to a plane of isotropy. The num-
ber of independent elastic constants is reduced to 5 (2 Young’s moduli, 2 Poisson’s ratios
and one shear modulus). Some rocks may have three distinct directions of anisotropy. These
materials are orthotropic and have 9 independent elastic constants (3 Young’s moduli,
3 Poisson’s ratios and 3 shear moduli).
Apart from its structural and mineralogical variability at microscopic scale, at macroscopic
https://telegram.me/Geologybooks
scale a rock mass may exhibit pronounced stratification, cracks, faults, etc. Generally there-
fore, it is not homogeneous. The fundamental hydromechanical properties of a rock are con-
sidered to be those of a sample big enough to contain a large number of component grains,
but small enough to exclude any major structural discontinuity. More precisely, it is recom-
mended that the diameter of a sample should be more than 10 times the diameter of the
largest rock grain [Kovari et al., 1983]. Samples of a few centimetres generally correspond to the
mesoscopic scale so defined and offer the advantage of being easily tested in the laboratory.
The aim of laboratory tests is to characterise the hydromechanical properties of a rock
using relatively simple experiments. The properties determined are based on a certain num-
ber of assumptions (homogeneous stress and strain states in the sample, isotropic behaviour,
etc.). It is also essential to ensure that they are actually representative of the behaviour of the
rock, not of the experimental device. Relations between stresses and strains can generally be
deduced from the measurements taken. These relations are then extrapolated to deal with
more complex situations at a scale different from that of the tests, this scale change requiring
additional assumptions.
In situ, a rock is subjected to a triaxial stress state and to the pressure exerted by the intersti-
tial fluid(s). The stresses are generally assumed to be oriented in the vertical direction and
two orthogonal horizontal directions, the vertical stress being determined by the overburden
(see Chapter 1.4).
The time scale of the strains may vary considerably depending on the problem consid-
ered: for instance, the strain rates associated with progressive depletion of a reservoir are
much slower than those encountered when drilling a well. The strength of a rock generally
increases with strain rate. In the laboratory, strain rates between 10–6 and 10–1 s–1 are gener-
ally applied. Within this interval (not necessarily representative of the problem studied), the
properties of most rocks can be considered as globally independent of the strain rate.
16 Chapter 1 • Elements of Rock Mechanics. Fundamentals
However, with very poorly permeable rock, such as shale, the strain rate chosen must be low
enough to prevent excess pore pressure generation.
The various types of test used to study the hydromechanical characteristics of a rock, in
particular its failure properties, are mainly characterised by the specific stress states imposed
and the accessible properties. The associated cost and ease of implementation must also be
taken into consideration when selecting the tests to be conducted during a characterisation
campaign.
This chapter outlines the main types of test on rock and the properties that can be deter-
mined. We will then focus on the compressive behaviour of a rock, using the results of the
main rock mechanics test: the conventional triaxial test.
This document does not discuss hollow cylinder tests that can be used to study well wall
behaviour [Fjaer et al., 1992] [Jaeger and Cook, 1979].
35
30
25
Axial stress (MPa)
20 Increase
in pore
15 pressure
ea
er
10
0
-0.1 0.0 0.1 0.2 0.3
Strain (%)
https://telegram.me/Geologybooks
Figure 1.2.1
Unconfined compression test on a sample of Berea sandstone with 24%
porosity.
Table 1.2.1 Order of magnitude of the static elastic properties of rocks with different UCSs
[partially from Le Tirant and Gay, 1972 and Fjaer et al., 1992].
W
h
D
https://telegram.me/Geologybooks
Figure 1.2.2
Principle of the Brazilian test.
In practice, to estimate the tensile strength, the rule is to take a ratio of 8 to 10 with UCS.
Displacement sa Downstream
sensor
pressure pump
Cell
Upstream
pressure pump
Figure 1.2.3
Diagram of an oedometric cell.
These tests are particularly well-suited to the study of elastoplastic behaviour of very
poorly permeable rocks. Since the sample is thinner than for a triaxial test, pressures reach
equilibrium more quickly. Note that, using oedometers, it is not generally possible to apply
loads large enough to cause sample failure.
Oedometric tests can be used to determine:
– the drained uniaxial compressibility,
– the creep properties (axial strain under constant load as a function of time),
– several poromechanical parameters [Bemer et al., 2001],
– the axial permeability.
The conventional triaxial test is carried out on cylindrical rock samples subjected to an axial
load, a uniform radial stress (via a pressurised confinement fluid) and an interstitial pressure
in the interconnected porous network. The sample is surrounded by an impermeable jacket
preventing any communication between the pore fluid and the confinement fluid. The pore
pressure must always be less than the confining pressure. The imposed stress field is there-
fore axisymmetrical and defined by the axial stress sa, the confining pressure pc and the
pore pressure pp (Figure 1.2.4). This load is associated with a strain state defined by the axial
strain ea and the radial strain er.
20 Chapter 1 • Elements of Rock Mechanics. Fundamentals
sa
Downstream pore
pressure (pp)
Porous disc
Jacket
Sample pc
Upstream pore
pressure (pp)
https://telegram.me/Geologybooks
Figure 1.2.4
Stress and pressure states during a triaxial test.
The piston applying the axial stress can be loaded mechanically using a load frame or
hydraulically via an upper pressure chamber. The confining and pore pressures are imposed
by servo-controlled hydraulic pumps. The axial and radial stresses are controlled indepen-
dently. Consequently, any required stress path can be applied to the specimen mounted in
the cell.
The devices equipped with two pumps to control the pore pressure upstream and down-
stream from the sample can be used to generate a flow through the sample and therefore to
measure its vertical permeability. The porous discs placed at each end of the sample ensure
uniform distribution of the upstream and downstream pore pressures. Some laboratories
have cells that can be used to measure the horizontal permeability of the sample under triax-
ial load.
Various devices can be used to measure the axial and radial strains. A local measurement
of these two strains can be obtained using gauges stuck to the sample. Vertical displacement
sensors (LVDT or cantilever type) placed inside the confinement chamber will provide a
semi-local measurement of the axial strain (over much of the sample height). It is not recom-
mended to place these sensors outside the cell, since the global measurement carried out also
includes the cell strains, sensitive in particular to the confining pressure applied. The radial
strain can also be measured using horizontal displacement sensors (LVDT or cantilever
type) placed on the jacket half-way up the sample (local measurement) or a small chain
around the sample on a diameter (semi-local measurement). Note that the jacket strains must
be calibrated in order to use these two measuring instruments correctly.
Chapter 1 • Elements of Rock Mechanics. Fundamentals 21
The results obtained are often studied in terms of mean stress p and deviatoric stress q,
associated with the volume strain ev and the deviatoric strain ed:
σa + 2 pc
p= εv = εa + 2εr (1.2.1)
3
q = σa − pc εd =
2
(ε −ε
3 a r
) (1.2.2)
The sample can be isolated by two valves located on the upstream and downstream pore
circuits. When these valves are open and the pore pressure is kept constant via a servo-con-
trolled pump, the test is said to be “drained”. When they are closed, the volume of fluid
inside the sample remains unchanged and the test is said to be “undrained”. Note that, in this
case, the pore pressure is controlled by the stress state imposed and the rock behaviour.
The first step of a triaxial test consists in saturating the sample with the test fluid, either
https://telegram.me/Geologybooks
mineral oil (which offers the advantage of avoiding any chemical interaction with the rock
constituents), water, brine, etc. The test generally starts by applying a high vacuum in the
sample previously placed in the triaxial cell. Then, under low initial axial stress and confin-
ing pressure, an upstream pore pressure (below the confining pressure) is applied, the down-
stream valve being kept closed. At this point, we simply monitor the volume of fluid
injected; the sample is considered to be saturated when the volume reaches a constant value.
Note that with a reservoir rock, this procedure assumes that the sample has first been cleaned
by removing all the in situ fluids. To date, no thorough studies have been conducted to deter-
mine the possible impact of this cleaning on the hydromechanical properties of the rock.
Note, however, that the rock wettability properties may be modified by the chemical treat-
ment applied.
The standard loading path is a drained path with constant pore pressure. The stresses are
increased first isotropically (hydrostatic loading) up to the required confining pressure, then
only the axial stress is further increased up to sample failure.
The axial stress is increased at constant axial strain rate (velocity control) or at constant
axial load rate dσa dt (force control). Unloading/reloading cycles can be carried out while
loading to estimate the elastic properties.
More complex loading paths can be performed, in particular to simulate the effects of
depletion on the behaviour of a reservoir (see § 3.1.3). An uniaxial strain test can be carried
out if the radial strain rate dεr dt is maintained at zero via a servo-control system. A varia-
tion of the confining pressure proportional to that of the axial pressure can also be applied.
When a constant axial load is imposed, a creep test is carried out; when a constant axial
strain is imposed, a relaxation test is carried out.
By definition, the deviatoric stress is “effective”, it does not depend on the pore pressure:
( )(
q ' = σa '− pc ' = σa − α pp − pc − α pp = σa − pc = q )
We will now consider the results of a standard drained triaxial test carried out at constant
pore pressure. The end of hydrostatic loading is characterised by a mean stress pco , zero
deviatoric stress and a pore pressure pop . This point is generally chosen as reference state
when calculating strains. The stress-strain relations then concern the stress increments:
σa + 2 pco 3 pco q
q = σa − pco and Δp = p − pco = − = (1.2.3)
3 3 3
Since the pore pressure is constant, irrespective of the expression of the effective stress
considered, we obtain:
Δp ' = Δp
https://telegram.me/Geologybooks
220 220
200 200
180 180
q (MPa)
q (MPa)
160 160
140 140
pp = 1 MPa
pp = 10 MPa
120 pp = 20 MPa 120 pp = 1 MPa
Failure line (pp = 1 MPa) pp = 10 MPa
100 Failure line (pp = 10 MPa) 100 pp = 20 MPa
Failure line (pp = 20 MPa) Failure line
80 80
20 40 60 80 100 120 140 20 40 60 80 100 120 140
p (MPa) p' = p–pp (MPa)
Figure 1.2.5
Deviatoric stress at failure as a function of mean stress and mean effective
stress on a limestone with 10% porosity [Vincké et al., 1998].
Chapter 1 • Elements of Rock Mechanics. Fundamentals 23
stress, q = deviatoric stress), the various curves lie on top of each other. In other words,
increasing the pore pressure by Δpp will produce the same reduction in axial stress at failure
as reducing the confining pressure by Δpp [Vincké et al., 1998].
The Mohr-Coulomb failure criterion traditionally used to describe brittle failure of rocks
can therefore be written in the form:
q = A + B p ' where p ' = p − pp (1.2.5)
parameters A and B being related to the (drained) cohesion c ' and the (drained) internal
angle of friction j ' of the rock by the following relations:
obtained during a triaxial test. For a loading path with pore pressure variations, different
effective stresses may need to be considered, including Biot’s effective stress which is
related to the rock elastic strains:
σ ' = σ − b pp 1 (1.2.6)
where b is Biot’s coefficient.
Tangent
modulus
Secant
modulus
https://telegram.me/Geologybooks
0 ev 0 ev
Figure 1.2.6
Stress-strain curves representative of elastic behaviour.
Above a certain level of loading some rocks, including crystalline rocks, sandstones of
low to medium porosity, compact limestones, chalks, etc. exhibit linear behaviour. The iso-
tropic elastic parameters of the rock are then easily determined using the stress-strain curves
obtained at constant pore pressure (drained test).
The drained bulk modulus Kd is defined by the slope of the mean effective stress-volume
strain curve, and the shear modulus G by one third of the slope of the deviatoric stress-devi-
atoric strain curve:
Kd = dp ' dεv (1.2.7)
3G = dq dεd (1.2.8)
The drained Young’s modulus Ed and the drained Poisson’s ratio nd are deduced from
the following relations:
9 Kd G 3 Kd − 2G
Ed = and νd = (1.2.9)
3 Kd + G 6 Kd + 2G
In the special case of hydrostatic loading (i.e. isotropic), only the bulk modulus can be
determined.
For purely axial loading, the drained Young’s modulus is defined by the slope of the
effective axial stress-axial strain curve: Ed = dσa ' dεv , and the drained Poisson’s ratio by
the opposite of the slope of the radial strain-axial strain curve: νd = − dεr dεa .
Chapter 1 • Elements of Rock Mechanics. Fundamentals 25
Biot’s coefficient can be determined by varying the pore pressure, keeping the axial
stress and the confining pressure constant. The ratio Kd b , defined as the opposite of the
slope of the pore pressure-volume strain curve, is then measured: Kd b = − dpp dεv .
Knowing Kd, we can deduce b.
Assuming undrained loading, we can calculate the undrained bulk modulus
Ku = dp dεv , the shear modulus 3G = dq dεd and Biot’s modulus, provided that Biot’s
coefficient Mb = dpp dεv has been determined.
Numerous rocks have nonlinear stress-strain curves in the elastic range, reflecting an
increase in bulk modulus as the material is compressed (Figure 1.2.7). This behaviour is
observed in particular with rocks whose structure can be schematised by an assembly of
cemented “grains”. The grains themselves exhibit linear elastic behaviour. The increase in
the contact surfaces introduces nonlinearities, however.
https://telegram.me/Geologybooks
60
50
40
p’ (MPa)
30
20
10
0
–0.010 –0.008 –0.006 –0.004 –0.002 0
Dev
Figure 1.2.7
Drained hydrostatic compression test on a reservoir sandstone with 19.8%
porosity.
In practice, nonlinear elasticity curves are interpreted in two ways. Depending on the extent of
the stress range associated with the phenomenon studied, the true behaviour may be represented
by a linear approximation by determining tangent or secant moduli of elasticity (Figure 1.2.6).
The nonlinear nature of the behaviour can also be taken into account via different models: Biot’s
semilinear model [Bemer et al., 2001] or Hertzian contact model [Vincké, 1994].
– At low mean effective stresses, brittle failure through the formation of a shear band
crossing the sample (shear localization) is observed. The result is a rapid drop in the
ability of the rock to withstand a load.
– At high mean effective stresses, failure occurs under the effect of homogeneous
microcracking (cataclastic flow). In this case we speak of ductile failure, with the rock
retaining an ability to withstand a load as the strain increases. It is important to note
here that this strain is due to slipping along a multitude of intersecting shear planes
and is therefore not true plastic strain.
Figure 1.2.8 schematises the corresponding failure surface in the q-p ' plane (deviatoric
stress-Terzaghi’s mean effective stress).
Brittle failure
"Shear localisation"
q (Pr)
Ductile failure
https://telegram.me/Geologybooks
"Cataclastic flow"
"Damage"
(C’)
pt p* p’
Figure 1.2.8
Failure surface of a sandstone.
Brittle and ductile failure modes are characterised by distinct critical stresses corre-
sponding to different steps in the development of the damage, which is associated with
acoustic activity of the rocks: microcracking, slipping with friction and pore collapse, etc.
These critical stresses can be identified by measuring the acoustic emissions during the test.
Figure 1.2.9 schematises the change in porosity (ΔΦ) as the mean effective stress (p ')
increases depending on the loading path followed. Note that in this case we are considering
the variation in Lagrangian porosity, i.e. calculated with respect to the initial volume of the
sample. The various curves show a tightening phase followed by an elastic phase (assumed
linear on the figure), these two phases being associated with a reduction in the permeability.
The third phase is nonlinear and occurs after crossing a critical effective pressure which
depends on the loading path (C ', C* or p*). The phenomena involved in this phase, known
as the post-elastic phase, are complex and may lead to a decrease or an increase in the per-
meability reflecting the competition under the effect of progressive microcracking of the
rock between firstly the greater connectivity and secondly the greater tortuosity and/or
reduction in the size of the pore throats.
Chapter 1 • Elements of Rock Mechanics. Fundamentals 27
p*
Dilatancy –
Shear Pc
localisation Ps C* Shear-enhanced
compaction –
Start of linear C’ Cataclastic flow
phase
0 DF
Dilatancy Contractancy
Figure 1.2.9
Behaviour of a sandstone subjected to compression.
https://telegram.me/Geologybooks
Hydrostatic loading
The curve Δφ − p ' presents a point of inflection at the critical effective pressure p*, associ-
ated with the onset of grain crushing and pore collapse induced by the growth of transgranu-
lar cracks and is accompanied by an increase in acoustic activity indicating a change of
physical mechanism. As long as p* is not reached, the rock behaviour corresponds to elastic
reduction of the porosity. Above p*, the post-elastic behaviour phase starts, systematically
inducing a decrease in the permeability in case of hydrostatic compression. The notion of
low or high mean stresses is defined with respect to the position of the ratio p '/p* on the
scale [0,1].
With consolidated sandstones, the stress levels associated with C* and Pc are very close
and only the effective compaction pressure C* is used to characterise ductile failure.
the porosity independent of the deviatoric stress. Intergranular cracks develop between C '
and Ps (probably due to cementation failure), inducing a faster increase in the acoustic emis-
sion rate. These cracks, whose preferential orientation is approximately parallel to the axial
stress, allow relative movement of the grains thereby causing an overall dilatation of the
porous space. Performing several unloading/reloading cycles between C ' and Ps would
demonstrate a reduction in the rock stiffness due to the effect of the damage. The change in
permeability is then related to the initial rock permeability.
– Rocks of low initial permeability, i.e. poorly connected rocks, will be highly sensitive
to the increased connectivity, due to the development of microcracks, and their per-
meability will therefore increase.
– For rocks of high initial permeability, i.e. well-connected rocks, the microcracks do not
modify the overall connectivity but allow grain movements, which increase the
tortuosity and/or reduce the size of the pore throats, thereby reducing the permeability.
Intragranular cracks develop on approaching the stress peak; subsequent coalescence of
these cracks finally results in failure through formation of a shear band crossing the sample.
The bulk permeability is then controlled by the properties of the localisation zone, which
may in particular obstruct the flow of fluids. The stress peak is marked by a peak in the
acoustic emission rate. The maximum acoustic emission rate associated with a brittle failure
(dilating strain) is less than the maximum rate associated with a ductile failure (compacting
strain).
In addition, since very few events generate acoustic activity in carbonates, the effective
critical pressures must be determined using the stress-strain curves only, creating a larger
dispersion than for the sandstones [Vajdova et al., 2004].
This paragraph is based on the experimental results obtained by Yale and Crawford
(1998) on various oil-saturated carbonates of porosities ranging from 14% to 42%. The tests
followed various stress paths to approximate the in situ conditions corresponding to extrac-
tion of oil from a reservoir.
q
Hardening cap
Shear
localisation
https://telegram.me/Geologybooks
Initial yielding
p’
Figure 1.2.10
Carbonate failure/plastification model.
The stress paths with low mean effective stress resulted in brittle failure of the samples.
The stress paths with higher mean effective stress were used to determine two yield surfaces
(Figure 1.2.10). Engstrøm (1991) also observed the existence of progressive strain-
hardening of chalks for this type of loading.
According to Yale and Crawford (1998), the first yield surface (initial yielding) corre-
sponds to the stress state from which the curve ΔΦ − p ' deviates from its initial linear slope
by more than 10%. This surface is to be compared with the effective compaction pressures
C* associated with sandstones.
The second yield surface (hardening cap) is reached at the point of inflection of the curve
ΔΦ − p ' and is therefore to be compared with the collapse pressures Pc. The fundamental
difference between sandstones and carbonates lies in the significant gap between the two
yield surfaces for carbonates: the ratio between the critical hydrostatic pressures of
hardening cap and initial yielding reaches 1.8 for 20% porosity carbonates.
Strain hardening of the samples is positive (irreversible strain can only increase by increasing
the load) and mainly controlled by the porosity of the carbonates tested: for low-porosity
samples, the initial hardening threshold stresses are higher and the gap between the initial
yielding and the hardening is larger. In addition, dispersion of experimental results is more
pronounced for higher-porosity samples due to the greater sensitivity to small-scale heterogeneities.
In case of brittle failure, the stress levels associated with damage onset and the failure
peak are very close for carbonates. Failure therefore occurs quite suddenly after the elastic
phase. As for sandstones and with the same reserves, the post-peak behaviour is sometimes
represented by elastoplastic strain-softening models.
30 Chapter 1 • Elements of Rock Mechanics. Fundamentals
1.2.6.4 Shales
The term shale designates rocks whose clay mineral content (clay volume divided by total
volume) is greater than 35% [Vernik et al., 1993] [Plumb, 1994]. It therefore includes rocks
of highly variable composition and structure. Clay minerals consist of stacks of elementary
sheets. Depending on the type of clay, the bonds between sheets are more or less strong.
Unlike kaolinite and illite, smectite has weak bonds, making it highly sensitive to water con-
tent; as a result it has the ability to fix water molecules between two neighbouring sheets and
exhibits a high swelling potential.
The study of swelling clays is highly complex and will not discussed here.
Very broadly speaking, we can identify two main categories of shales according to their
water content (weight of water divided by weight of solid particles) [Rousset, 1988]:
– “plastic” shales with high water content (> 15 %),
– “stiff” shales with low water content (< 10%).
https://telegram.me/Geologybooks
Since the elastic limit of plastic shales is virtually zero, they can withstand high strains
with no apparent macroscopic failure. After a strain hardening phase, loadings at low mean
effective stress lead to a “rounded” stress peak, followed by softening behaviour. We then
obtain a stress plateau demonstrating the residual strength retained by the shale.
This type of behaviour is generally represented by a Cam-Clay type yield surface elasto-
plastic model (Figure 1.2.11).
Strain softening
Strain hardening
p’
Figure 1.2.11
Cam-Clay model yield surface.
The behaviour of stiff shales is more like that of an elastobrittle rock. They may even
exhibit very high strengths [Lin, 1981]. At low mean stresses, a sharp peak is observed, fol-
lowed by a rapid drop in the stress. The brittle failure data of stiff shales can be represented by
a linear Mohr-Coulomb criterion or a nonlinear criterion [Green et al., 1973] [Lin, 1981] [Sar-
gand and Hazen, 1987] [Cook et al., 1991] [Steiger and Leung, 1992] [Gavito, 1996]. Few
experimental data are available, however, to close the yield surface for high mean stresses.
In addition, plastic shales exhibit pronounced delayed response. At a given stress state,
the rock may therefore undergo strains over a long period of time, which may stabilise or
Chapter 1 • Elements of Rock Mechanics. Fundamentals 31
not. Stiff shales may also develop more or less large delayed strains. Viscous behaviour
models may be required to take into account these phenomena [Boidy, 2002].
Depletion tests in the laboratory are generally conducted on cylindrical samples placed in a
triaxial cell, with test procedures leading to a stress state more or less close to the in situ con-
ditions. Depletion can be simulated by either reducing the pore pressure at constant axial
stress to simulate a constant weight of earth (“true depletion” tests), or by increasing the
effective stresses at constant pore pressure. Obviously, simulating depletion by reducing
pore pressure is more representative of reservoir production mode (see § 3.1.3).
One disadvantage with replacing the pressure reduction by an increase in effective
stresses is that it introduces Biot’s coefficient. Depletion tests are in fact mainly carried out
to characterise the strain behaviour of the rock and, in this case, the relevant effective stress
https://telegram.me/Geologybooks
is Biot’s.
The loading path followed in terms of the evolution in axial stress and confining pressure
may be:
– isotropic ( Δpc ' = Δσa ');
– oedometric (i.e. uniaxial, the confining pressure evolution is then controlled automat-
ically to maintain the condition Δεr = 0 );
– proportional ( Δpc ' = K Δσa ' ), in order to follow the reservoir stress path when known
(see § 3.1.1).
Numerous experimental studies demonstrate that the behaviour of reservoir rocks is
highly dependent on the stress path [Holt, 1990] [Warpinski and Teufel, 1990] [Rhett and
Teufel, 1992a] [Ruisten et al., 1996] [Schutjens and de Ruig, 1996] [Keaney et al., 1998]
[Schutjens et al., 1998]. Consequently, if the stress path is known, it should be followed in
the tests conducted to determine the reservoir rock characteristics.
Occluded Connected
porosity porosity
= +
Solid matrix
Figure 1.3.1
Porous medium seen as the superimposition of two continuous media.
dm
+ div w = 0 (1.3.1)
dt
Chapter 1 • Elements of Rock Mechanics. Fundamentals 33
The fluid mass content per initial volume unit m is simply related to the Lagrangian
porosity Φ by the following relation which expresses complete saturation of the medium:
m = ρ fl Φ (1.3.2)
where rfl designates the fluid density.
The relative flow vector of fluid mass w is directly proportional to the relative flow rate
of the fluid with respect to the skeleton:
⎛
w = ρ fl V V =φ ⎜ v fl − vs ⎞⎟ (1.3.3)
⎝ ⎠
where f is the Eulerian porosity, which is the ratio of the current porous volume to the cur-
rent total volume, and V is the filtration vector.
Assuming infinitesimal transformations, the Lagrangian and Eulerian porosities are sim-
ply related by:
( )
https://telegram.me/Geologybooks
Φ = 1+ ε φ
where ε = tr ε .
( )
where ρo = ρso 1 − Φo + ρofl Φo .
or (
g fl = g fl pp , T )
34 Chapter 1 • Elements of Rock Mechanics. Fundamentals
1 ∂g fl ∂g fl
= − s fl = (1.3.5b)
ρ fl ∂pp ∂T
Note that the fluid-specific enthalpy is related to its specific free energy yfl:
pp
g fl = ψ fl + (1.3.6)
ρ fl
By differentiating the constitutive equations (1.3.5b), we obtain:
d ρ fl dpp
= − 3α fl dT (1.3.7a)
ρ fl K fl
dpp dT
ds fl = −3α fl + cp (1.3.7b)
https://telegram.me/Geologybooks
ρ fl T
where Kfl designates the fluid bulk modulus, 3α fl the fluid volumetric thermal expansion
coefficient and cp the fluid volumetric specific heat capacity at constant pressure.
For a low compressibility liquid and small pressure and temperature variations, we can
obtain a linearised expression of the fluid constitutive equations by integrating equations
(1.3.7) with Kfl, 3α fl and cp constant:
⎡ pp − pop ⎤
1
=
1 ⎢
ρ fl ρo ⎢
1−
K fl
+ 3α fl T − To ( )⎥⎥
fl ⎣ ⎦
pp − pop T − To
s fl = sofl − 3α fl + cp
ρofl To
By applying the first law of thermodynamics to the porous medium seen as the superimposi-
tion of the solid skeleton and the interacting interstitial fluid, we obtain the energy equation,
which expresses the conservation of energy:
dE
dt
=σ :
dε
dt
(
− div hfl w + q + F .w ) (1.3.9)
where E designates the internal energy density of the porous medium per initial volume unit,
hfl = g fl − T s fl the specific enthalpy of the fluid and q the outgoing heat flow vector.
The second law of thermodynamics leads to the entropy balance:
( )
dS ⎛ q⎞
+ div s fl w ≥ − div ⎜ ⎟ (1.3.10)
dt ⎝T⎠
https://telegram.me/Geologybooks
where S designates the entropy density of the porous medium per initial volume unit. The ≥
sign shows that there is a spontaneous production of entropy: ϕ T , where ϕ is the overall
dissipation.
By introducing the porous medium density of free energy ψ = E − T S , using the above
two relations, we derive the Clausius-Duhem inequality which shows that the dissipation ϕ
is non-negative:
ϕ = ϕs + ϕ fl + ϕth ≥ 0 (1.3.11)
dε dm dT dψ
ϕs = σ : + g fl −S − (1.3.12)
dt dt dt dt
represents the intrinsic dissipation, which turns out to be related to the solid skeleton alone.
If we introduce the entropy and free energy of the skeleton per initial volume unit,
Ss = S − m s fl and ψ s = ψ − m ψ fl , the fluid constitutive equations (1.3.5a) and the relation
m = ρ fl Φ lead to the following expression of the intrinsic dissipation:
dε dΦ dT dψ s
ϕs = σ : + pp − Ss − (1.3.13)
dt dt dt dt
ϕ fl = ⎡ − grad pp + ρ fl F ⎤ ⋅V (1.3.14)
⎣ ⎦
represents the dissipation associated with the fluid and more particularly its movement with
respect to the skeleton.
q
ϕth = − ⋅ gradT (1.3.15)
T
represents the thermal dissipation.
36 Chapter 1 • Elements of Rock Mechanics. Fundamentals
Due to the highly different nature of these dissipations, we adopt the decoupling assump-
tion which consists in replacing inequality (1.3.11) by three separate inequalities:
dε dΦ dT dψ s
ϕs = σ : + pp − Ss − ≥0 (1.3.16)
dt dt dt dt
ϕ fl = ⎡ − grad pp + ρ fl F ⎤ ⋅V ≥ 0 (1.3.17)
⎣ ⎦
q
ϕth = − ⋅ grad T ≥ 0 (1.3.18)
T
After identifying the spontaneous production of entropy ϕ T , inequality (1.3.10) finally
leads to the thermal equilibrium equation:
⎡ dS
( )
⎤
T ⎢ + div s fl w ⎥ = − div q + ϕM
https://telegram.me/Geologybooks
⎣ dt ⎦
where ϕM = ϕs + ϕ fl represents the mechanical dissipation density.
w
ρ fl (
=V = Λ − grad pp + ρ fl F ) (1.3.19)
Note that the volume forces often prove to be negligible compared with the hydraulic
gradient.
For a low compressibility interstitial fluid, the hydraulic conductivity L is expressed as a
function of the intrinsic permeability k, which depends solely on the geometry of the porous
network, and of the dynamic viscosity of the interstitial fluid mfl:
k
Λ= (1.3.20)
μ fl
k kg
Λ= + (1.3.21)
μ fl pp
Chapter 1 • Elements of Rock Mechanics. Fundamentals 37
When these models are included in finite elements simulation software, we may intro-
duce variation laws of the hydraulic parameters k and kg as a function of the porosity, which
represents a simplified way of taking into account the evolution of the porous network and
thereby of including this aspect of the hydro-mechanical coupling in the numerical model.
Thermal dissipation expression (1.3.18) shows that the heat conduction law must relate the
heat flow vector to the force responsible for heat production − grad T . In this case, its sim-
plest form is the linear relation given by Fourier’s law. For an isotropic medium, it is writ-
ten:
q = − λT grad T (1.3.22)
where lT is the thermal conductivity.
https://telegram.me/Geologybooks
Intrinsic dissipation expression (1.3.13) shows that, due to its nature, the skeleton density of
free energy depends on the external state variables: ε , Φ and T (observable variables whose
variation can be controlled externally), their associated variables being σ , pp and Ss.
According to the postulate of local state [Coussy, 2004], the skeleton density of free energy
is independent of the evolution rates and can finally be expressed in the form:
(
ψ s = ψ s ε , Φ, T , χ1 , χ2 ,..., χN ) (1.3.23)
where the variables ci are “internal” state variables (their variation cannot be controlled
externally).
Since the intrinsic dissipation cannot be negative, we have:
⎛ ∂ψ s ⎞ ⎛ ∂ψ s ⎞ ⎛ ∂ψ s ⎞ ∂ψ s
⎜σ − ⎟ :d ε + ⎜ pp − ⎟ dΦ − ⎜ S + ⎟ dT − ⋅dχ ≥ 0 (1.3.24)
⎝ ∂ε ⎠ ⎝ ∂Φ ⎠ ⎝ ∂T ⎠ ∂χ
We will now consider evolutions where the internal variables are constant. Since ine-
quality (1.3.24) must be satisfied for any evolution in the external variables independently of
each other, we obtain the following constitutive equations:
Note that equations (1.3.25) and (1.3.26) remain valid for evolutions including a varia-
tion of internal variables. Complementary evolution laws governing the evolution of internal
state variables need to be introduced however.
Intrinsic dissipation expression (1.3.12) leads to the expression of the density of free
energy and the constitutive equations of the porous medium relating the external state vari-
ables ε , m and T to their associated variables σ , gfl and Ss:
(
ψ = ψ ε , m, T , χ1 , χ2 ,..., χN ) (1.3.27)
∂ψ ∂ψ ∂ψ
σ= g fl = S=− (1.3.28a)
∂ε ∂m ∂T
For porous media with no chemical activity, however, it is simpler to introduce the fluid
mass content directly into the constitutive equations (1.3.26):
https://telegram.me/Geologybooks
We will now consider a porous medium whose skeleton exhibits isotropic linear elastic
behaviour. We can therefore adopt for ys a quadratic expression ψ sel which depends solely
on the first and second invariants of the strain tensor.
By differentiating the constitutive equations (1.3.26) with Gs = Gsel = ψ sel − pp Φ , we
obtain the porous skeleton thermoporoelastic constitutive equations which involve only
coefficients which are independent of the interstitial fluid:
⎛ 2G ⎞
d σ = ⎜ Kd − d ε 1 + 2G d ε − b dpp 1 − 3αd Kd dT 1
⎝ 3 ⎟⎠
dpp
d Φ = b dε + − 3αΦ dT (1.3.29)
N
dT
dSs = 3αd Kd d ε − 3αΦ dpp + Cε
T
where Kd is the skeleton bulk modulus, G the skeleton shear modulus, b Biot’s coefficient,
1/N the pore compressibility, 3ad the skeleton volumetric thermal expansion coefficient,
3aΦ the pore volumetric thermal expansion coefficient and Cε the skeleton volumetric heat
capacity at constant strain.
Chapter 1 • Elements of Rock Mechanics. Fundamentals 39
Assuming that the solid matrix is homogeneous and isotropic, we can determine expres-
sions relating coefficients b, N and aΦ, to the characteristics of the solid matrix [Coussy,
2004]:
Kd 1 b − Φo
b = 1− = (1.3.30)
Ks N Ks
αd = αs (
αΦ = b − Φo αs) (1.3.31)
where Ks is the bulk modulus of the solid matrix and as the thermal expansion coefficient of
the solid constituent.
thermodynamic system involve the fluid mass content m. They are obtained from constitu-
tive equations (1.3.28b) and the fluid constitutive equations (1.3.7).
Differentiating relation (1.3.2), we obtain:
dm d ρ fl
= dΦ + Φ
ρ fl ρ fl
dm dpp
= dΦ + Φ − 3Φα fl dT
ρ fl K fl
( )
d m s fl = s fl dm − 3Φα fl dpp + m cp
dT
T
By introducing the above relations in equations (1.3.29), we can eliminate the
Lagrangian porosity variation, replacing it by the variation in fluid mass content. Finally, we
obtain the constitutive equations:
⎛ 2G ⎞
d σ = ⎜ Kd − d ε 1 + 2G d ε − b dpp 1 − 3αd Kd dT 1
⎝ 3 ⎟⎠
dm dpp
= b dε + − 3αm dT (1.3.32)
ρ fl M
dT
dS − s fl dm = 3αd Kd d ε − 3αm dpp + Cεd
T
1 1 Φ
with = + (1.3.33)
M N K fl
40 Chapter 1 • Elements of Rock Mechanics. Fundamentals
The skeleton characteristics Kd, G, b and ad, prove to be drained properties of the porous
medium, i.e. they can be measured from tests where the interstitial pressure is maintained
constant ( dpp = 0). The drained volumetric heat capacity at constant strain Cεd takes into
account the fluid mass content via the term m cp .
Biot’s modulus M and the differential thermal expansion coefficient am depend on the
interstitial fluid. For a porous medium saturated by a low compressibility fluid, Biot’s
modulus and the differential thermal expansion coefficient can be considered as constant to
the first order of approximation. When the interstitial fluid is highly compressible, however,
they depend on the interstitial pressure and the temperature.
The drained constitutive equations (1.3.32) can be combined to obtain an undrained
expression of these equations:
https://telegram.me/Geologybooks
⎛ 2G ⎞ dm
d σ = ⎜ Ku − ⎟ d ε1 + 2Gd ε − bM 1 − 3αm Ku dT 1
⎝ 3 ⎠ ρ fl
dpp dm
= − bd ε + + 3αm dT (1.3.35)
M ρ fl
dm dT
dS − s fl dm = 3α Ku d ε − 3α m M + Cεu
ρ fl T
(
Φ = 1+ ε φ )
A first order approximation yields:
d Φ = dφ + Φo d ε
The constitutive equation (1.3.29) then leads to the equation governing the variation in
elementary Eulerian porosity f:
dpp
(
dφ = b − Φo d ε + ) N
− 3αΦ dT (1.3.36)
Chapter 1 • Elements of Rock Mechanics. Fundamentals 41
pp = 0 + pp
Note that the effective stress σ defined in this way corresponds to Terzaghi’s effective stress
https://telegram.me/Geologybooks
(Figure 1.3.2).
s s — PP1
pp 0 pp
pp 0 pp
= +
pp pp
0
Figure 1.3.2
Load decomposition.
Since the matrix behaviour is assumed to be linear and isotropic with a bulk modulus Ks,
decomposition of the stresses (1.3.37) is associated with the following decomposition of the strains:
pp
ε =ε − 1 (1.3.38)
3 Ks
where ε represents the strains produced by σ .
Figure 1.3.3 illustrates the validity of Biot’s semilinearity assumption for a 19.8%
porosity reservoir sandstone. This experimental curve is obtained from a hydrostatic test
where the confining pressure and the pore pressure are increased simultaneously, thereby
avoiding any variation in the effective stress and the associated nonlinear strain.
Since only the strains ε include nonlinear variations, potential Gs is taken to be the sum
of the quadratic potential of the linear poroelasticity Gsel leading to the constitutive equa-
tions (1.3.29) and a third-order nonlinear potential H which is purely a function of ε :
( ) ( )
Gs ε , pp = Gsel ε , pp + H ε ( ) (1.3.39)
42 Chapter 1 • Elements of Rock Mechanics. Fundamentals
120
100
80
PP (MPa)
60
40
Experimental curve
20
Model
0
–0.004 –0.003 –0.002 –0.001 0
Figure 1.3.3
https://telegram.me/Geologybooks
∂Gsel ∂H
σ= +
∂ε ∂ε
∂Gsel 1 ∂H
Φ=− − tr
∂pp 3 Ks ∂ε
Hence,
⎛ 2G ⎞ ∂H
σ = ⎜ Kd − ⎟ ε 1 + 2G ε − b pp 1 +
⎝ 3 ⎠ ∂ε
(1.3.40)
pp1 ∂H
Φ − Φo = b ε + − tr
N 3Ks ∂ε
Equations (1.3.40) can be decomposed into a hydrostatic part and a deviatoric part:
1 ∂H
σ = Kd ε − b pp + tr
3 ∂ε
∂H 1 ∂H
s = 2Ge + − tr (1.3.41)
∂ε 3 ∂ε
pp 1 ∂H
Φ − Φo = b ε + − tr
N 3 Ks ∂ε
Chapter 1 • Elements of Rock Mechanics. Fundamentals 43
tr σ ε
where σ = , s = σ − σ 1 and e = ε − 1. Note that s = s and e = e .
3 3
H=D
ε3
3
)(
+ ( F − D ε I2 − 3 I3 ) (1.3.42)
where I2 = ε11 ε22 + ε22 ε33 + ε33 ε11 − ε12 ε21 − ε23 ε32 − ε31 ε13 and I3 = det (e). F and D are
constants which have the same dimensions as stress.
Taking the principal stress coordinate system, we obtain the constitutive equations:
D � 2F 2
s = Kd e - b pp � e
3
https://telegram.me/Geologybooks
s11 � 2G e11 �
F-DÈ 2
� �
e � 2e11e22 � e332 � 2e11e33 - 2 e112 � 2e22 e33 ˘˙ (1.3.43a, b, c)
3 ÍÎ 22 ˚
pp D � 2 F
F - Fo � b e � - e2
N 3 Ks
The expressions associated with the other principal stress directions are obtained simply
by circular permutation.
After differentiating equations (1.3.43a) and (1.3.43c) we can explicitly define a tangent
drained bulk modulus, a tangent Biot’s coefficient and a tangent pore compressibility func-
tion of e . We obtain:
� � � �
ds = Kdt e d e - bt e dpp
t (1.3.44a, b)
Ê 1ˆ
� �
d F � bt e d e � Á ˜
ËN¯
� e � dpp
D � 2F
� �
Kdt e � Kd � 2
3
e
D � 2F
with � �
bt e � b - 2
3 Ks
e (1.3.45a, b, c)
t
Ê 1ˆ 1 D � 2F
� �
ÁË N ˜¯ e � N -
3 Ks2
e
t
Kdt Ê 1ˆ bt - Fo
Note that the following relations are satisfied: bt � 1 - and Á ˜ � .
Ks ËN¯ Ks
44 Chapter 1 • Elements of Rock Mechanics. Fundamentals
Equation (1.3.43a) can be used to express e as a function of the mean effective stress
s � s � pp and therefore relate the tangent drained bulk modulus, the tangent Biot’s coeffi-
cient and the tangent pore compressibility to the effective mean stress alone:
D � 2F
� �
Kdt s � Kd2 � 4
3
s
� �
bt s � 1 - �1 - b �2 - 4 D3�K2 F Ks (1.3.46a, b, c)
s s
t 2
Ê 1 ˆ 1 - Fo Ê 1- b ˆ D � 2F s
ÁË N ˜¯ � K - Á K ˜ - 4
s Ë s ¯ 3 Ks2 Ks2
Figure 1.3.4 shows that the tangent drained bulk modulus does in fact vary as a function
of the effective means stress s � s � pp .
https://telegram.me/Geologybooks
10000
8000
6000
Kot (MPa)
4000
2000 pp = 1 MPa
pp = 51 MPa
0
0 20 40 60
s + pp (MPa)
Figure 1.3.4
Variation law of the tangent drained bulk modulus on a sandstone with 19%
porosity [Bemer et al., 2001].
It is more difficult to define a tangent shear modulus using Biot’s semilinear model. From
the deviatoric part of the constitutive equations (1.3.43b), we must obtain a relation of type
��
d s � 2Gt e de (1.3.47)
i.e., according to the principle direction 1,
��
ds11 � 2Gt e de11 (1.3.48)
Chapter 1 • Elements of Rock Mechanics. Fundamentals 45
�2
F-DÈ
3 Î
� � � � � �
- 2e11 � e22 � e33 d e11 � e11 � e22 - 2e33 d e22 � e11 - 2e22 � e33 d e33 ˘
˚
To obtain an expression of type (1.3.48), relation e33 � e22 must necessarily be satisfied,
which amounts to having a condition of axisymmetry around principal axis 1. In this special
case, equation (1.3.43b) leads to relation:
q � 3G ed �
9
4
� �
F - D ed2 (1.3.49)
2
� �
where q � s 22 - s11 is the deviatoric stress and ed � e22 - e11 � ed the associated devia-
3
toric strain.
� �
dq � 3Gt ed d ed (1.3.50)
with � �
Gt ed � G � 3
F-D
2
ed (1.3.51)
Equation (1.3.49) can be used to express ed as a function of the deviatoric stress q and
thereby relate the tangent shear modulus to the deviatoric stress alone:
��
Gt q � G2 � F - D q� � (1.3.52)
To date, the validity of this variation law has not been checked experimentally.
The poroelastic model described here is the modified Cam-Clay model [Roscoe and Bur-
land, 1968]. This model was developed to represent the isothermal behaviour of normally
consolidated or slightly overconsolidated clays. It is an isotropic poroelastoplastic behaviour
model with isotropic strain hardening, defined for a water-saturated porous medium and
therefore initially associated with a model describing the viscous flow of a low compress-
ibility fluid. Extending poroelastoplasticity to unsaturated porous media results in relatively
complex behaviour models. Some examples can be found in [Alonso et al., 1990] and [Dan-
gla et al., 1997].
implement yield functions which depend only on the first two invariants. The mean stress p
and the deviatoric stress q are generally used in rock mechanics [Charlez, 1994]:
tr s 3 tr s
p� and q � s : s with s � s - 1
3 2 3
The volume strain ev and the deviatoric strain ed correspond respectively to the mean and
deviatoric stresses:
2 tr e
ev � tr e and ed � e : e with e � e - 1
3 3
Stresses p and q are the variables associated with strains ev and ed by the density of free
energy ys. We obtain:
s :d e � p d ev � q d ed
https://telegram.me/Geologybooks
Rather than a plastic variation in fluid mass content, it is generally preferred to introduce
a plastic Lagrangian porosity variation [Dangla et al., 1997]. We then write:
d F � d Fe � d F p (1.3.56)
A thermodynamic state is characterised by the external state variables: ev, ed and Φ, the
internal state variables: evp , edp , Φp and variables characterising the strain hardening c :
�
y s � y s ev , ed , F, evp , edp , F p , c �
Using the energy separation assumption, we can express the free energy of the skeleton
in the following form [Coussy, 2004]:
�
y s � y sel ev - evp , ed - edp , F - F p � U c� � � (1.3.57)
Chapter 1 • Elements of Rock Mechanics. Fundamentals 47
where y sel represents the proportion of the free energy of the skeleton recoverable as
mechanical energy and therefore depends only on reversible strains and porosities. U repre-
sents the quantity of energy trapped through the strain-hardening phenomena and is there-
fore related to the strain-hardening variables.
By introducing equations (1.3.54) and (1.3.57) in inequality (1.3.53), we obtain the fol-
lowing intrinsic dissipation expression:
d evp d edp d F p �U d c
js � p �q � pp - � �0 (1.3.58)
dt dt dt � c dt
which demonstrates that the thermodynamic forces associated with evp , edp , Φp and c , are
respectively p, q, pp, and the thermodynamic force V defined by the relation:
�U
V�- (1.3.59)
�c
https://telegram.me/Geologybooks
dq � 3G ( d ed - d evp ) (1.3.60a, b, c)
dpp
d F - d F p � b( d ev - d evp ) �
N
where p ' � p � b pp represents the elastic mean effective stress.
For a low compressibility interstitial fluid, equation (1.3.60c) can also be expressed as a
function of the fluid mass content m:
dm
rofl
�
- d F p � - b d ev - d ev p � � dp
M
(1.3.61)
The constitutive equations (1.3.60) define a model whose elastic part is linear. The
model defined by Roscoe, however, assumes that the constitutive equations are in fact non-
linear. This special formulation is based on Terzaghi’s hypothesis, which assumes elastic
and plastic incompressibility of the solid matrix. The elastic and plastic effective stresses are
then identical to Terzaghi’s effective stresses, defined by: s ' � s � p1 [Charlez, 1994].
The choice of a nonlinear constitutive equation is based on experimental observations
which show that, under hydrostatic loading, the void ratio decreases in direct proportion to
the logarithm of the absolute value of the mean effective stress. Hence,
d p'
dee � -k (1.3.62)
p'
48 Chapter 1 • Elements of Rock Mechanics. Fundamentals
where k is the swelling coefficient. For small strains and an incompressible matrix, the void
ratio is related to the volume strain by d ev � de 1 � eo . The nonlinear volumetric constitu-
tive equation of the modified Cam-Clay model is then written:
k d p'
d ev - d evp � - (1.3.63)
1 � eo p '
dq
d ed - d edp � (1.3.64)
3G
The model formulated in this way is thermodynamically admissible [Coussy, 2004].
In view of the various constitutive equations defined above, two formulations of the
Cam-Clay model can be considered:
https://telegram.me/Geologybooks
� � 12 ÈÍÍÎ� p � b p � � q ˘
2 2
f p, q, pp , pcr � p
� pcr - pcr2 ˙
2
˙˚
where b and M are constants and pcr is a positive strain-hardening variable called the (cur-
rent plastic) critical effective pressure which is associated with the scalar thermodynamic
force V .
pb ' � p � b pp is the plastic mean effective stress, not to be confused with the elastic
mean effective stress p ' � p � bpp . The yield surface f is then written:
� � 12 ÈÍÍÎ� p '� p � � q ˘
2 2
f pb ', q, pcr � b cr
- pcr2 ˙ (1.3.65)
2
˙˚
� � � �
In plane pb ', q , the plasticity criterion f pb ', q, pcr � 0 defines a family of ellipses
passing through the origin, having a maximum value at pb ' � - pcr and cutting the x-axis at
pb ' � - 2 pcr (Figure 1.3.5). The latter point, which corresponds to the current elastic limit of
the material under hydrostatic loading, is called the effective consolidation pressure. Double
Chapter 1 • Elements of Rock Mechanics. Fundamentals 49
Zero dilatancy
q
H=0
Plastic
Plastic dilatancy H < 0
contractancy
H>0
M
Compression Tension
Figure 1.3.5
Cam-Clay model yield surface.
https://telegram.me/Geologybooks
the value of the critical pressure represents in fact the absolute value of the maximum plastic
mean effective pressure previously undergone by the material. A material is therefore said to
be normally consolidated (or overconsolidated) if the current plastic mean effective stress is
equal to (or less than in absolute value) the consolidation effective pressure.
Hence, �
d evp � d l pb '� pcr �
q
d edp � d l (1.3.66)
2
�
d F p � d l b pb '� pcr �
with d l �
1 È
HÎÍ � �
pb '� pcr dpb '�
q dq ˘
2 ˚
˙ � 0, where H is the strain-hardening modulus.
For points on the yield surface belonging to the semi-ellipse located in the half-plane
defined by pb '� pcr � 0 , flow rule (1.3.66) shows that d evp � 0 and d F p � 0 . The material
is then plastically compacting and the plastic porosity decreases.
Inversely, for points on the yield surface belonging to the semi-ellipse located in the
half-plane defined by pb '� pcr � 0, we have d evp � 0 and d F p � 0. The material is then
plastically dilating and the plastic porosity increases.
For the point located at the intersection between the straight line of equation pb '� pcr � 0
and the ellipse, flow rule (1.3.66) shows that d evp � 0 and d F p � 0 . Plastic strain then takes
place with no variation in volume and the plastic porosity remains unchanged.
� �
The Cam-Clay model introduces a potential h p, pp , pcr not associated with the criterion to
specify the evolution equations governing the scalar hardening variable c:
https://telegram.me/Geologybooks
� � 1 b - Fo
� �
2
h p, pp , pcr � p � b pp � pcr (1.3.67)
� �
2 2
1 - Fo
d c � dl
�h
� dl
b - Fo
p � b pp � pcr � � (1.3.68)
� pcr
� �
2
1 - Fo
Relations (1.3.66) and (1.3.68) show that variable c can be identified with the plastic
void ratio ep, which represents the irreversible variation of void ratio with respect to a given
initial reference state:
F p - Fo evp
ep � (1.3.69)
� �
2
1 - Fo
d pcr
de p � - l - k � � pcr
with l � k (1.3.70)
d pcr 1 b - Fo
�- d evp (1.3.71)
l -k
� �
pcr 2
1 - Fo
Chapter 1 • Elements of Rock Mechanics. Fundamentals 51
0.0
Oedometric curve
–0.2
–0.4
Volumetric strain (%)
k
1 + eo l
–0.6 1 + eo
1
–0.8
–1.0
–1.2 1
–1.4
1 10 100
https://telegram.me/Geologybooks
Figure 1.3.6
Poroplastic behaviour of an argillite in an oedometric test
(under Terzaghi’s hypothesis).
Relations (1.3.59) and (1.3.70) lead to the following expression of trapped energy:
È p ˘
� � � �
o exp - e
U e p � l - k pcr Í ˙
ÍÎ l - k ˙˚
(1.3.72)
H�
1 b - Fo
�
p p � b pp � pcr �� p � b p � (1.3.73)
� �
l -k 2 cr p
1 - Fo
Consequently, for yield points such that the material exhibits plastically compacting
behaviour, the hardening modulus is positive and the critical pressure increases, which
corresponds to a strain-hardening phase. Inversely, for yield points such that the material has
a plastically dilating behaviour, the hardening modulus is negative and the critical pressure
decreases, which corresponds to a strain-softening phase. For yield points such that the
plastic strain takes place with no variation in volume or plastic porosity, the hardening
modulus is zero and the material behaves as a plastically perfect material. For these points,
the strains which are now only plastic occur at constant stresses and interstitial pressure. The
state is said to be critical.
The locus of critical states is defined by the following relations:
È p ˘ È p ˘
o exp - e
p � b pp � pcr o exp - e
Í ˙�0 q- pcr Í ˙�0
ÍÎ l - k ˙˚ ÍÎ l - k ˙˚
52 Chapter 1 • Elements of Rock Mechanics. Fundamentals
� �
In plane pb ', q , it is the straight line of equation q � - pb (Figure 1.3.5).
�1 - Fo � � l - k � ÈÍÊÁ 1
2
˘
1 ˆ q
d evp � � ˜ dpb '� dq ˙
b-F ÍÁ p pb ' ˜¯ 2 pcr pb ' ˙
o ÎË cr ˚
�1 - Fo � � È ˘
2
Í q2 ˙
d edp � l -k Í �q
dpb '�
� �
dq ˙ (1.3.74)
b-F 2
Í pcr pb '
4
pcr pb ' pcr � pb '
https://telegram.me/Geologybooks
o ˙
Î ˚
df p � b d evp
d evp �
b-F Í p ' 2 � h2 ˙
o Î b ˚
d edp � (1.3.75)
b-F 2 - h2 Í p ' 2 � h2 ˙
o Î b ˚
df p � b d evp
νd
σH
' = σv' , where s ' are effective stresses and nd the drained Poisson’s ratio.
1 − νd
Chapter 1 • Elements of Rock Mechanics. Fundamentals 53
In most cases, the in situ stress state does not correspond to this ideal scheme and the
stress distribution is more complex, depending on geological history. We identify:
– regional stresses which are related to major geological phenomena (plate tectonics);
indications of their orientation and sometimes their amplitudes are now available
[Reinecker et al., 2005],
– local stresses which are affected by local accidents such as folds, faults, salt domes;
stress rotations may then lead to orientations and amplitudes differing considerably
from the original scheme.
In foothill regions, for example, tectonic forces dominate and the maximum horizontal
stress may be the highest of the principal stresses [Last et al., 1995].
We assume that the lithostatic vertical stress is one of the principal stresses.
Maximum horizontal
stress (sHmax)
Drilling induced
fracture
Figure 1.4.1
Breakouts and fractures induced in the stress field.
54 Chapter 1 • Elements of Rock Mechanics. Fundamentals
Breakouts and drilling-induced fractures can be detected by systems recording the hole
geometry, multi-arm callipers and well imaging probes.
P1 azimuth
C1-3
C2-4
https://telegram.me/Geologybooks
Figure 1.4.2
Example of breakout detection.
Calliper Calliper
Depth
Depth
Calliper Calliper
Depth
Depth
Breakout
Washout
and washout
Calliper
Depth
KeySeat
Figure 1.4.3
Interpretation of CALLIPER data [Plumb & Hickman, 1985].
56 Chapter 1 • Elements of Rock Mechanics. Fundamentals
N E S W N N E S W N
270 1 2 3 4 5 6 90
180
https://telegram.me/Geologybooks
Figure 1.4.4
Interpretation of UBI (a) and FMI (b) images in terms of breakouts
[Zoback et al., 2003].
Figure 1.4.5
FMI images interpreted in terms of drilling-induced fractures. These fractures
are characterised by low vertical extension and irregular shape and can only
appear on one side of the hole (HEF petrophysical consulting site).
Chapter 1 • Elements of Rock Mechanics. Fundamentals 57
60
300 60 300 10%
60
40
5%
20 40° 45°
270 90 270 5% 10% 15% 20% 90
https://telegram.me/Geologybooks
Figure 1.4.6
Comparison between MSIP fast shear azimuth and OBMI induced fractures
[Franco et al., 2005].
In offshore, the density log can be calibrated on measurements taken in geotechnical sur-
veys performed before installation of platforms or subsea pipelines.
Breakdown pressure
https://telegram.me/Geologybooks
ISIP
Pressure
Leakoff pressure
Time
Figure 1.4.7
Hydraulic fracturing test.
The pressure is increased by injecting a fluid at constant flow rate. The pressure curve
can be subdivided into five phases: a linear increase phase, a nonlinear phase during which
the rate of increase slows down, the maximum pressure (breakdown pressure), a more or
less steady decrease indicating the opening of a fracture, lastly the relaxation curve after the
injection pump stops. The Leak-Off Pressure (LOP) corresponds to the end of the linear sec-
tion with change of slope. During the relaxation phase, the Instantaneous Shut In Pressure
(ISIP) or closure pressure is assimilated to the minimum stress in situ. The shut-in pressure
corresponds to the level of pressure exerted by the fluid inside the fracture, just sufficient to
keep the edges of the fracture open. Consequently, the minimum stress amplitude can be
determined by analysing the pressure during fracturing and closure of the fracture. Several
methods are available for determining closure pressure from the fall off curve. One of these
uses a change in linearity of the pressure decay on a graph of pressure plotted against the
square root of time [Desroches and Kurkjian, 1998, Zoback, 2008]. If the amplitude of the
closure pressure is less than the overburden at this depth, then the smallest principal stress
will correspond to the horizontal minimum stress (Shmin).
Chapter 1 • Elements of Rock Mechanics. Fundamentals 59
The Leak-Off Test (LOT) is conducted during the drilling operations after cementing a
casing. It consists in increasing the pressure at the casing shoe to a value greater than the
LOP but less than the rock breakdown pressure (Figure 1.4.8). This operation is mainly used
to evaluate the maximum permissible mud pressure. In a variant of the LOT, the Extended
Leak-Off Test (ELOT), the pressurisation phase continues beyond the formation breakdown
pressure and the fracture propagates. The test consists of two or more pressurisation cycles
(Figure 1.4.9). Extended leak-off tests use procedures from leak-off tests, therefore open
bottomhole or “barefoot” well configuration is pressurised without using packers.
In practice, the quality of 50% of the LOTs is poor. There are several reasons for this:
open section too long, mud compressibility artefact, pumping stopped too soon or relaxation
phase badly recorded [Chardac et al., 2005]. LOT and ISIP generally overestimate the mini-
mum stress.
A good estimate of the minimum stress amplitude can generally be obtained from a
series of fracturing tests. ELOTs are reliable tests compared with LOTs, but they are not as
https://telegram.me/Geologybooks
5.0
4.5
ISIP
4.0 LOP
3.5
Pressure (MPa)
Volume (litres)
3.0
2.5
2.0
1.5
900
1.0
0.5
0.0 0
11700 12000 12300 12600 12900 13200 13500 13800 14100
Time (sec)
Figure 1.4.8
Leak-off test (credit ITC).
60 Chapter 1 • Elements of Rock Mechanics. Fundamentals
2nd shut in
Leak-off
Fracture initiation stage 1st re-opening stage 2nd re-opening stage Time
Pump rate
https://telegram.me/Geologybooks
Time = 1 hour
Figure 1.4.9
Typical ELOT result [Enever et al., 1996].
pressure to stabilise at each step. All steps must last the same time, for example 30 min. The
fracture pressure can be identified from a graph of injection pressures against flow rates. The
SRTs obviously give an upper limit of the minimum horizontal stress.
In all cases an approach with a “reconciliation” diagram between all the tests performed
is necessary [Desroches and Kurkjian, 1999].
Pp:pore pressure,
σ ΔT :thermal stress,
ΔP:difference between the well pressure during drilling and the pore pressure.
The measurements of shear wave velocities made with crossed dipole sonic logging tools
such as DSI, MSIP, Sonic scanner, Wavesonic (fast and slow Dt shear) supply anisotropy
indicators. These indicators are expressed as percentages. For example, DT-based anisot-
ropy is:
Dtslow * − Dt fast
* 100
( Dtslow + Dt fast ) / 2
After extraction, a core expands more in the direction of the maximum stresses. This expan-
sion generates microcracks perpendicular to the three principal stresses. Two methods are
based on this relaxation principle:
– ASR (Anelastic Strain Recovery),
– DSCA (Differential Strain Curve Analysis).
With these methods, the core must be oriented, which is often costly, and highly accurate
strain measurements are required. With ASR, the measurements must be taken immediately
62 Chapter 1 • Elements of Rock Mechanics. Fundamentals
after recovery since anelastic strains develop in less than 50 hours. A detailed interpretation
of the tests is provided in the book by Fjaer et al. (1991). DSCA requires isotropic recon-
finement. Although inaccurate, these two methods provide an idea of the orientation, and
even the amplitude, of the stresses [Cui et al., 2006]. The measurements become very diffi-
cult to interpret in anisotropic and/or fractured rocks.
Geological strata contain faults, fractures and discontinuities at several scales. Friction in
these existing discontinuities limits the stresses.
At equilibrium, Jaeger and Cook demonstrated that the ratio of the maximum and mini-
mum effective stresses had to remain less than a certain value which depends on the coeffi-
cient of friction of the discontinuity m:
https://telegram.me/Geologybooks
2
σ '1 ⎡ 1 ⎤
σ '3 ⎢
(
≤ ⎢ μ2 + 1 ) 2 + μ⎥
⎥
⎣ ⎦
The values of s '1 and s '3 depend on the tectonic system:
– normal fault stress regime (sv > sH max > sh min): s '1 =�s 'v and s '3 =�s 'h min,
– strike-slip fault stress regime (sH max > sv > sh min): s '1 =�s 'H max and s '3 =�s 'h min,
– reverse stress regime (sH max > sh min >�sv): s '1 =�s 'H max, s '3 =�s 'v.
According to Byerlee [1978], the coefficient of friction of the discontinuities m lies
between 0.6 and 1.
Geomechanical modelling tools can be used to constrain the stress tensor, checking the
stability or instability of the boreholes where clearly identified breakouts and induced frac-
tures appear.
These tools require the following input data:
– the assumed directions and amplitudes of the minimum and maximum horizontal stresses,
– the pore pressure,
– the geomechanical properties of the rocks where the breakouts appear,
– a failure criterion (e.g. Mohr Coulomb).
Indications such as mud losses during drilling may also be used.
The discipline of oil reservoir geomechanical modelling is in full expansion. The first
studies were conducted to explain the significant subsidence observed in some mature fields
with highly porous (North Sea chalk) or poorly consolidated (Zuata field, Venezuela)
reservoirs [Charlez, 1997]. The scope of geomechanical modelling is much broader and
Chapter 1 • Elements of Rock Mechanics. Fundamentals 63
concerns all stress-dependent problems: modification of the mud window, instability of the
borehole walls, modification of flow directions, reactivation of fractures or faults,
interpretation of modifications in 4D seismic data, microseismicity. The stress modifications
are induced by reservoir depletion or, on the contrary, injection of fluids during hydrocarbon
production or gas storage (natural or CO2).
The zone considered for geomechanical modelling must include a volume larger than the
reservoir since the mechanical properties of surrounding formations have an impact on the
mechanical behaviour of the reservoir. The width of the sideburden will be ideally 3 times
the horizontal extension of the reservoir region. The height of the underburden will be of the
same order of value as the depth of the reservoir region.
Geomechanical modelling may be more or less complex, 2D or 3D depending on the
https://telegram.me/Geologybooks
problem to be solved. The geometry of the various layers is derived from seismic interpreta-
tion, constrained by the available borehole data. The number of stratigraphic units to be
taken into account depends on the borehole data available, especially the seismic impedance
contrast (product of the formation density by the compression wave velocity). The dynamic
elastic properties from the sonic logs and the static elastic properties used in the geomechan-
ical models are in fact closely related.
The mesh will be oriented according to the orientation of the principal stresses.
The geological strata exhibit discontinuities which may be more or less detectable and
more or less cemented or open. They share the following characteristics: low shear strength,
negligible tensile strength and hydraulic conductivity often much greater than the wall rock.
Several categories can be identified:
– diaclases: rock fractures, due to compressive, tensile or shear failures related to the
tectonic stresses; the two parts of the rock have not moved, however;
– faults: fractures identical to the diaclases but which caused a relative movement of
two parts of the wall rock. A slip has therefore occurred along this fault and the throw
varies from a few tens of centimetres to a few hundred metres;
– sedimentary joints: these are the joints separating two strata deposited at different
times under different conditions.
Discontinuities are generally areas of high deformability, where failure is easier and
where the fluids can flow.
All realistic models of a structure and its environment must take into account the possi-
ble discontinuities. Geometric discontinuities are generally correctly taken into account.
Resistivity, nuclear and gamma ray logs give a good estimation of the mineralogy and, in
particular, the location of the shale layers [Chardac et al., 2005].
The compression and shear wave velocities (Vp and Vs if available) and the density logs
provide information on the dynamic elastic properties, but the plasticity or viscosity proper-
ties can only be obtained by conducting laboratory tests.
A few calibration points can be obtained from the laboratory data, for the conversion of
sonic logs into static elastic properties.
If no S-wave records are available, the Vs velocity is estimated from Vp using empirical
formulae that depend on the lithology [Castagna et al., 1985].
The Vp and Vs velocities are used to calculate the undrained dynamic moduli:
⎛ 4 ⎞
Kudyn = ρ ⎜Vp2 − Vs2 ⎟ and Gdyn = ρVs2, with
⎝ 3 ⎠
Eudyn = ρVS2
( 3VP2 − 4VS2 ) and νdyn = (VP2 − 2VS2 )
(VP2 − VS2 ) u
2 (VP2 − VS2 )
The static properties can be deduced from the dynamic properties [Vidal-Gilbert et al.,
2005]. The procedure consists of two phases. Firstly, “fluid saturation” correction is carried
out to change from undrained dynamic elastic moduli to dynamic drained moduli. Secondly,
dynamic drained moduli are changed to static drained moduli.
One of the problems frequently studied in rock physics, especially for analysis of logs,
samples and seismic data, is the prediction of seismic velocities in rocks saturated with a
given fluid from another saturation state. This fluid substitution problem can be predicted
using the Biot-Gassmann theory (1951). Gassmann calculated, for a quasi static load, the
difference in elastic modulus between a drained porous medium and the same medium in
undrained state, in which local variations in fluid content are nil during mechanical loading.
The Gassmann equation can be expressed as:
Ku Kd K fl
= +
Ks − Ku Ks − Kd (
φ Ks − K fl )
with
Ku: undrained bulk modulus
Kd: drained bulk modulus
Ks: bulk modulus of the matrix (mineral bulk material forming the rock)
Kfl: fluid bulk modulus
Chapter 1 • Elements of Rock Mechanics. Fundamentals 65
⎛K ⎞
with: b = 1− ⎜ d ⎟
https://telegram.me/Geologybooks
⎝ Ks ⎠
1 b−φ φ
and = +
M Ks K fl
In this expression, we can see that the first term in the expression of 1/M is often much
smaller that the second since Ks is always much bigger than Kfl. Neglecting this first term,
we obtain a very simple expression:
b2
Ku ≈ Kd + K , [Rasolofosaon and Zinszner, 2003].
φ fl
The error thus committed has been estimated to be less than 5%.
In order to apply the Gassmann equation to wave propagation (dynamic case), the fre-
quency must be low enough for the variation in saturating fluid induced by passage of the
seismic wave to be uniform throughout the porous space (no gradient). This low frequency
limitation explains why the Biot-Gassmann equation is well suited to seismic frequencies
(< 100 Hz). While this relation can also be applied for the log analysis frequency band, it
may prove less relevant for ultrasound measurements on samples. The latter point is open to
discussion since Rasolofosaon and Zinszner (2004) demonstrated that the experimental
results on samples seemed to be in agreement with the Biot-Gassmann theory, for various
fluids saturating the porous medium.
66 Chapter 1 • Elements of Rock Mechanics. Fundamentals
The Biot-Gassman equation is then frequently used to estimate the dynamic drained
moduli from undrained dynamic moduli:
2
⎛ K dyn ⎞
K fl ⎜ 1 − u ⎟
⎜⎝ Ks ⎟⎠ Eudyn
Kddyn = Kudyn − with Kudyn = where:
⎛
⎜1 −
K fl ⎞
⎟φ−
K fl ⎛
⎜1 −
Kudyn ⎞
⎟
(
3 1 − 2νu )
⎜⎝ Ks ⎟⎠ Ks ⎜⎝ Ks ⎟⎠
Ks: bulk modulus of the solid matrix (of the material forming the grains),
K fl : bulk modulus of the fluid saturating the pores,
f: porosity (deduced from the density measurements, for example).
The bulk modulus of the rock matrix Ks depends on the mineral composition, but also to
https://telegram.me/Geologybooks
the second order on pressure and temperature. To a first approximation, the orders of magni-
tude indicated in Table 1.5.2 (dynamic values) will be used.
Table 1.5.2 Mean values of dynamic bulk and shear moduli in the main minerals forming sedimentary rocks
[Zinszner and Pellerin, 2007].
1
K fl =
Sw (1 − Sw )
+
Kw Koil
with:
Kw : water bulk modulus,
Koil : oil bulk modulus,
Sw : water saturation, determined using log analyses, for example.
Chapter 1 • Elements of Rock Mechanics. Fundamentals 67
The water bulk modulus varies with salinity, pressure and temperature (Figure 1.5.1).
For brine at a concentration of 15 g/L, Kfl = 2.2667 + 0.0063 p in GPa with p = pressure
expressed in MPa [Zinszner and Rasolofosaon, 2008].
3.5
3.0
2.5
2.0
1.5
https://telegram.me/Geologybooks
1.0
0.5
0 50 100 150 200 250 300 350
Temperature (°C)
Figure 1.5.1
Bulk modulus of a brine against pressure, temperature and salinity
[Batzle and Wang, 1992].
The bulk modulus of oil depends on its density, as well as the pressure and temperature
(Figure 1.5.2).
Pa 25
M
Pa 50
2000 M
Pa
1500
1000
500
0
0 50 100 150 200 250 300 350
Temperature (°C)
Figure 1.5.2
Bulk modulus of oil against pressure, temperature and composition
[Batzle and Wang, 1992].
68 Chapter 1 • Elements of Rock Mechanics. Fundamentals
The estimated dynamic drained moduli can then be changed to static drained moduli
using empirical correlations [Wang, 2000].
For reservoir rocks, we can use the correlations determined by Wang, who identifies two types:
– if Edstat < 15 GPa then Edstat = 0.4145 Eddyn − 1.0595 (GPa ) and
– if Edstat > 15 GPa, then Edstat = 1.1530 Eddyn − 15.1970 (GPa ) .
For shales, we can use correlations determined for all lithologies [Lacy, 1997, Cooper
and Hatherly, 2003] or for shales only [Horsrud, 2001], according to the following
sequences:
– Vp Vs: P and S wave velocities,
– Ed: dynamic modulus,
– Es: static modulus,
– UCS: unconfined compressive strength.
For example:
https://telegram.me/Geologybooks
– Ed = 0.265(Vp )2.04 [Lacy, 1997] with Vp in km/s and Ed in millions of psi (all lithologies),
– UCS = A.Vp3 [Cooper & Hatherly, 2003],
– ES = 0.0428( Ed )2 + 0.2334 Ed (for clays),
Es = 0.076(Vp )3,23 G = 0.03(Vp )3.3 and UCS = 0.77(Vp )2.93 (for shales) [Horsrud, 2001].
The static Young’s moduli are lower than the dynamic moduli [Wang, 2000; Yale et al.,
1995]. The static Poisson’s ratios are greater than or equal to the dynamic Poisson’s ratios.
For both static and dynamic properties, laboratory measurements remain preferable
when determining local or regional relations of the zone of interest.
Obviously no matter which approach is chosen, the measurements cannot be considered
as completely reliable. Current geomechanical models are deterministic. The values
assigned to the geomechanical properties must take into account the risk factor associated
with the problem studied.
The traditional displacement boundary conditions consist in blocking the horizontal dis-
placements of the vertical edges and the vertical displacements of the horizontal edges,
unless the upper edge coincides with that of the natural terrain.
The vertical displacements are fixed at the bottom and the horizontal displacements on
the lateral sides.
The contact surfaces between layers and the faults are generally fixed in a first simula-
tion. Friction can then be introduced, or not, depending on the resulting stress state.
2
Geomechanics, Drilling
and Production
https://telegram.me/Geologybooks
Drilling bits can be divided into two broad categories: roller-cone bits and PDC (Polycrys-
talline Diamond Compact) bits (Figure 2.1.1).
Roller-cone bits have several toothed cones which are free to turn as the bit is rotating.
Tri-cone is the dominant type. The teeth can either be ground in the matrix or consist of
tungsten carbide buttons inserted into holes in the cones. The teeth are designed to crush the
rock as the bit rotates (Figure 2.1.2).
PDC bits consist of several discs containing numerous cutting edges made of tungsten
and diamond. They cut the rock by shearing, layer after layer. These bits, more modern than
roller-cone bits, can reach high drilling rates in relatively soft formations. They are also
more efficient in hard rocks, provided that certain recommendations are respected (speed
and weight on bit), especially for new bits which are more fragile.
Drillers have two main concerns:
– reaching maximum efficiency under ordinary conditions,
– solving “problems” under exceptional conditions which could eventually damage the
bit, in particular.
The “problems” most frequently encountered are excessive vibrations, whose effects at
the surface arouse the driller’s attention, although lateral vibrations and precession some-
times remain unnoticed.
These excessive vibrations are generally observed when drilling through hard rocks at
low rate of penetration (regular bounce or precession), but the TRAFOR recordings taken by
IFP Energies nouvelles have also demonstrated bounce activity (more chaotic than the pre-
vious activity), while drilling rapidly through soft rocks; expulsion of the cuttings in “plugs”
seems to be the most likely cause for the excitation [Putot, 1995].
Until recent years in the drilling industry, “rock drillability” and “drillstring dynamics”
were quite independent disciplines [Pessier and Fear, 1992].
70 Chapter 2 • Geomechanics, Drilling and Production
a b
Pilot pin
Locking ring
Journal
Special Shirtail
metal inlays Gage
Bit leg
Nozzle
O-ring seal boss
Trademark
Figure 2.1.1
Examples of bits: roller-cone (a) and PDC (b) [Nguyen, J-P., 1993].
Steel
Steel
F
(a) (b)
Figure 2.1.2
Rock destruction modes: punching for the roller-cone bit, shearing for the new
PDC bit [Hélène Geoffroy, 1996].
“Rock drillability” concerns the ability of the drill bit to penetrate and remove the rock,
in the context of the bottomhole, controlled from the surface via the drillstring. This special-
ity generally deals with ordinary drilling situations and aims at optimising bit concepts with
respect to cut and hydraulics.
“Drillstring dynamics” studies examine the types of vibration which would damage the
drilling system equipment (drillstring and bit), using in particular surface indicators more or
less able to detect the various types of vibration at the bottom. This speciality also analyses
the conditions leading to a vibratory malfunction.
Chapter 2 • Geomechanics, Drilling and Production 71
With traditional rock drillability, it is impossible to correctly study the reduced drilling
rate regimes with high dissipation at the cutting face and therefore validly assess the drilling
efficiency [Warren and Smith, 1985]; the advance speed is reduced considerably by the
vibrations; in addition, the dynamics of the bottomhole drilling system, in particular, is
strongly related to the aggressivity of the cutters and the conditions under which the cuttings
are expelled. The distinction between the two disciplines therefore becomes quite arbitrary.
Drillability in dynamic regime is the end result of combining the two disciplines.
As regards drillability, one direction consists in analysing drillings in relatively soft forma-
tions. Tool performance, mainly that of monobloc “PDC” tools, is generally not a major problem,
provided that the cuttings are suitably evacuated, which nevertheless demands well-designed tool
hydraulics, good control and, in particular, sufficient hydraulic power [Wardlaw, 1971].
It seemed useful to describe the phenomenon associated with a tool cleaning fault,
through balance equations both as regards the divided solid (rock) and the drilling mud,
relying on simplified behaviour laws [Putot, 1995]. The evolution of the mixture of fluid and
rock particles is governed by a “source” term including the conditions under which the rock
https://telegram.me/Geologybooks
is broken by the cutters, together with terms of retention on the particles crushed but not
necessarily expelled, function of the entrainment capacity of the drilling mud.
Note that the hydraulics is a determining factor in soft rocks for evacuation of the cuttings
but also essential in “hard” rocks for good cooling of the cutters, guarantee of their integrity.
Under normal conditions, the d-exponent indicator can be used to “subtract” the effects
of modifications of the “loading” W and N. Its “logarithmic” definition offers the advantage
of slightly reducing the experimental dispersion, since the signal is generally very noisy due
to fluctuations related to the mineralogical characteristics of the rocks at grain scale and to
the pore pressure.
Plotted against the depth for a given lithology, the d-exponent provides an image of the
compaction. In a normal situation, the greater the depth, the greater the compaction. It can
be used to detect any compaction anomaly, for example areas of pore overpressure
[Mouchet and Mitchell, 1989].
Real-time drilling interpretation is commonly based on observing the normalised rate of
penetration RD alone, which provides information on the lithology changes and would also
give an indication on bit operation. The following expression was proposed by Warren (1984):
n
⎛ W ⎞
RD = k ⎜ (2.1.2)
⎝ σ D2 ⎟⎠
https://telegram.me/Geologybooks
where:
s is the unconfined compressive strength of the rock, frequently compared with a hardness.
k is an empirical proportionality factor related to the bit efficiency (ability to break the rock).
n is an exponent characterising, in principle, the local failure mechanism but which, in
practice, is more representative of a degree of uncertainty related to the position on the curve
d = f(W) (Figure 2.1.3).
n = 2 is the value generally used to characterise field data; it corresponds to a normal
drilling regime, for which increasing the weight on bit W significantly increases the perfor-
mance, i.e. the penetration per revolution, d = R/N. We simply take the logarithm of relation
(2.1.2) to obtain the d-exponent (2.1.1).
A decrease in n indicates a priori a deterioration of drilling conditions, generally due to
difficult expulsion of rock cuttings and sometimes more or less sudden damage to the drill
bit. The weight then has less effect on the normalised penetration d. These approximations
reflect the existence of an “S-shaped curve” whose upper plateau corresponds to saturation
of d with W (Figure 2.1.3).
For both roller-cone bit and PDC, Falconer et al. (1988) demonstrated that using the
assumptions adopted in his model, RD (ability of the bit to penetrate the rock) is inversely
proportional to the shear strength t of the rock; paradoxically, for the roller-cone bit,
although the kinematics consists mainly of punching, the efficiency of the shear mechanism
is the determining factor when assessing penetration. The paradox can be explained by the
fact that the implicit assumption t/s = constant is often made.
Falconer et al. (1988) proposed a quantity FORS proportional to the drilling resistance
defined later by Detournay and Defourny (1992) for the PDCs:
W σ
FORS = = (2.1.3)
RD D2 k
Chapter 2 • Geomechanics, Drilling and Production 73
d = R/N
n<1
n=1
n=2
W
https://telegram.me/Geologybooks
Figure 2.1.3
Sigmoid shape of the graph of penetration per revolution d against weight W:
the exponent n reflects the local behaviour.
FORS is the ratio of rock penetration resistance s divided by bit efficiency factor k.
Cunningham (1978) proposed a numerical frame very similar to expression (2.1.2),
where the exponents assigned to the hardness s and the weight W in the expression of their
dependency with respect to the normalised penetration differ – which is not the case in
(2.1.2) – and are fully defined by the rock hardness; he improves his description by explic-
itly introducing the drop in performance resulting from the pressure difference between mud
pressure and pore pressure, using the following empirical relation where he assumes that the
deviations from linearity R(N) are due to effects of pressure difference Δp:
NW a
R= (2.1.4)
k1σ 1.5 a
d + NW Δp ( )0.75
Numerous laboratory experiments have corroborated these expressions.
The previous literal expression, essentially empirical, showing the effect of interstitial pres-
sure, explains the role of this pressure in the shape of the curves R(N) at fixed W; it provides
information on the decrease in penetration d = R/N with speed N, useful to characterise the
dependency of torque on bit with angular speed, especially to assess the stick-slip mechanism.
As an illustration, Cunningham observes:
– linearity R(N) for Δp = 0,
– R(N) in N0.65 for Δp = 20 MPa (soft rock),
– R(N) in N0.85 for Δp = 20 MPa (hard rock).
For more recent studies concerning the effect of pore pressure on drilling performance,
interested readers can refer in particular to Peltier and Atkinson (1986), Lesso and Burgess
(1986), Detournay and Atkinson (1991), Hanson (1991).
74 Chapter 2 • Geomechanics, Drilling and Production
0.5
0.4
0.3
TD
0.2
0.1
0
0 0.005 0.010 0.015 0.020 0.025 0.030 0.035
RD
https://telegram.me/Geologybooks
Figure 2.1.4
Example of TD – RD values in a shale layer at a depth of approximately 1 800 m.
Detournay and Defourny (1992) completely revised the formulation for PDC bits by
introducing assumptions on the local mechanics of the cutters and by including the forces on
the entire tool. To characterise the results, variables S and E quite similar to the previous one
are used, called respectively drilling resistance and specific energy. Detournay and
Defourny also characterise the operation of roller-cone bits in the same system of variables,
if only to clearly identify their specificity.
Note that, for PDC bits, the analysis strategy used by Kuru and Wojtanowicz (1988) is
very similar although the experimental verification does not seem to have been as thorough
as that conducted by Detournay with Glowka’s data (1989).
Falconer and Detournay use a similar approach to characterise normal system operation,
by aligning points representative of bit performance in the plane of variables. This straight
line is the analogue of a rock failure criterion under global load. In the remainder of the
document, we will refer only to Falconer’s “roller-cone bit” terminology. Warren’s
mechanical model, expressed in response variables RD and TD is then written:
T D = c1 + c 2 R D (2.1.7)
We see immediately the similarity of (2.1.7) with Coulomb’s law, provided that in relation
(2.1.2) n is given a value of 2, which corresponds to the “normal” regime. However, it is proba-
bly more correct to position (2.1.7) with respect to a simple Amonton type interface friction law.
These normal conditions are represented on Figure 2.1.5 by a straight line.
We assume that neither the rock parameters determining the characteristics of the
straight line (ratio t/s of the shear strength to the compressive strength) nor the bit condition
76 Chapter 2 • Geomechanics, Drilling and Production
parameters are affected during the observation interval; only the two response components
RD and TD (or S and E) vary according to the disturbances on the input: weight on bit W and
speed of rotation N, which are nevertheless already mean values. Variations in rock “hard-
ness” s may also affect the response.
TD
0.1
https://telegram.me/Geologybooks
0
0 0.1 RD
Figure 2.1.5
Normal conditions in the TD – RD plane.
The slope representative of normal operating conditions lies between 0.07 and 0.15 for “ref-
erence shales”. These rocks are considered as favourable to analyse the wear factor, for exam-
ple, since the ratio t/s changes very little within this class, unlike the other rock categories.
The deviations on TD and on the slope values are attributable to a “non-shale” facies or to
transfers within the weight/torque torsor which are generally symptomatic of dysfunctions.
Refer to the publication by Falconer and Defourny for the quantitative effects of the wear
on the operating points: the drilling resistance is increased, at constant weight on bit W,
according to relation (2.1.3), which decreases RD. The complementary effect is a substantial
drop in TD indicating a decrease in shearability (Figure 2.1.6).
The criterion indicates the increasing energy contribution with axial thrust, which can be
interpreted as increasing “friction” at the borehole wall with confinement.
Another type of improvement was obtained by taking into account bit wear, since this
has an important effect on the torque. This correction was therefore introduced in the estima-
tions as an indicator TD for the torque, sensitive to cutter wear, like the more traditional indi-
cator RD which corresponds to the axial dimension. Combined use of these two indicators
yields a more reliable diagnosis than with a single indicator, especially when studying tran-
sitions to degraded or dangerous regimes.
Chapter 2 • Geomechanics, Drilling and Production 77
New bit
TD (sharp punch)
Worn bit
(blunt punch)
https://telegram.me/Geologybooks
RD1/2
Figure 2.1.6
Effects of wear on the relation of bit behaviour in the plane of reduced
variables, especially on the slope and the Y-intercept [Falconer and Defourny,
1992].
Figure 2.1.7
Example of PDC type bit, showing a wear flat (DBS document).
https://telegram.me/Geologybooks
This action is in line with the studies also conducted by SANDIA [Glowka, 1987, 1989]
and the University of Minnesota [Detournay, 1991, 1996]. The distribution of elementary
forces on the diamond impregnated cutters and therefore the ability of the bit to cut down the
rock have been predicted more accurately. It has led to a strategy for positioning the cutters
which is better than that associated with the new bit.
Crushed rock
Flat
Grain and
Damaged area powder
Crushed rock
Shear band
Intact rock
Damaged area
Figure 2.1.8
On the left: Cutting mechanisms observed in sandstone with new PDC blades.
On the right: Representation of the damaged region below the flat [Geoffroy,
1996].
d r
d
d KIc
2
l
https://telegram.me/Geologybooks
Figure 2.1.9
Cutting mechanisms depending on grain size r and penetration per revolution d
[Detournay, 1996].
Bit signature:
wear and blunting decreasing Rock signature:
performance = abrasive regime specific energy with threshold l
d
III Insufficient hydraulics
II Shear cutting poor cleaning conditions
efficient regime inefficient extrusion of
crushed material
New bit
III
sharp cutters
II Blunt PDC
cutters
Improving control by
l
choosing specific design
W
I Abrasive action,
inefficient rubbing
Figure 2.1.10
Bit penetration d against weight W exerted [Detournay, 1996].
Chapter 2 • Geomechanics, Drilling and Production 81
Regime I: d r
The first “cutting” regime is that of surface spalling and inefficient abrasion, exhibiting high
specific energies. Grain plastification occurs. In the penetration per revolution – weight on
bit axis system, this regime marks out a weight efficiency threshold which increases with the
size of the wear flat [see also Kuru and Wojtanowicz, 1988]. The order of magnitude of the
penetration per revolution is less than that of the grain size.
Regime III: d l
The third regime is considered as brittle. It appears from a depth l proportional to Irwin’s
https://telegram.me/Geologybooks
P P
R
a a b a a
3 E
Sphere D
(FE)
C F Cone Cavity
Sphere B G Sphere model
Mean pressure Pm/Y
Cone A
https://telegram.me/Geologybooks
0
1.0 10 100
E* tan E*a
Non-dimensional strain or
Y YR
Figure 2.1.11
Johnson model (1985).
2a
Elastic Plastic
Crack l
Figure 2.1.12
Normal indentation of a wedge into rock
[Damjanac and Detournay model, 1995].
Chapter 2 • Geomechanics, Drilling and Production 83
In practice, the characteristic of the global rock drilling bit system in the required opera-
tion is also a straight line, in the plane with the weight on the x-axis and the penetration per
revolution on the y-axis. The slope of the line is representative of the intrinsic specific
energy of the rock. It is very high for a sandstone, the specific energy being very low. If the
weight on bit is not precisely controlled, we rapidly enter abrasion mode (regime I) or mode
III which, due to accumulation of chips, soon causes cleaning problems.
In contrast, the specific energy is high in shales, increasing with confining pressure. The
bit is controlled better with the weight.
The weight efficiency threshold is related to the presence of a more or less pronounced
wear flat. It is worthwhile converting this information in a plane containing the reduced
variables weight and torque and interpreting using Detournay’s formalism.
The link between the specific energy and the cutter angle of attack may also be
considered; if very large, energy is lost while cutting, but there is better tool control.
The cutting test indicated previously can be used to obtain the rock properties, provided
https://telegram.me/Geologybooks
E* a
ε=
YR
where E* is the equivalent modulus of elasticity, R the indentation radius, a the contact cir-
cle radius, Y the yield stress.
For the conical indentation, this strain, independent of the load, is:
E * tg β
Y
The punch is conical, with half-angle at the top p/2-b. The radius of the contact circle is
a. Sharp cones (b close to p/2) favour plasticity, as do small radii R.
84 Chapter 2 • Geomechanics, Drilling and Production
E*/Y is the ratio of the modulus of elasticity to a yield criterion, generally the shear
failure modulus. For a highly cohesive rock this ratio will be moderate (several hundred), for
a poorly cemented rock it will be higher (several thousand). Note that E*/Y is the reciprocal
of the strain at failure.
The formalism on the transition of mechanisms provides a mechanical explanation of the
Johnson or Detournay indices. Including the specific fracture work, it introduces competi-
tion between failure and plasticity. This is in fact the meaning of the cutting mechanism
boundary l indicated by Detournay.
Drillability is traditionally studied in terms of steady state regimes, with limited fluctuation
of the characteristic quantities of the system.
By definition, the study of dysfunctions concerns transitions, whether due to a poorly
https://telegram.me/Geologybooks
managed control parameter – generally the weight, but possibly the speed of rotation or the
hydraulics – whose untimely variation triggers the phenomenon, or through the sudden
appearance of a rock formation transition for which the current control is unsuitable, thereby
inducing an unfavourable effect.
This description is also valid for dysfunctions such as excessive vibration in hard rocks
or for an incident related to clogging in soft rocks.
As we have seen in 2.1.1.2, normal behaviour is characterised by an operation line in a
suitable plane of variables (Figure 2.1.5) relating mean quantities (RD – TD) and which sug-
gests a Mohr-Coulomb type failure criterion with, in particular, a steady increase in the nor-
malised torque TD when the confinement effect RD increases (Falconer representations for
the roller-cone bit, Detournay for the PDC).
A significant departure of the operation points away from the “normal” line may come
from:
(i) A radical modification of the rock behaviour at failure (sudden lithological transition).
(ii) A vibratory dysfunction during the interaction of the bit with a very hard rock or on
the contrary degenerating into clogging/jamming with a soft rock for which the con-
trol was not adapted on time. Clogging designates conditions under which the
balance of divided solid material (cuttings) is in excess, jamming refers to conditions
which are “even more irreversible” for which physicochemistry makes the bonds
between aggregates and bit structure even stronger.
(ii) is in fact an exaggerated form of (i).
(i) In this case, after facies transition, the bit is no longer under optimised conditions
allowing efficient drilling but we are nevertheless outside the field of normal opera-
tion.
(ii) In this case, operation is disturbed on leaving the transition and drilling cannot
continue in the long term; an inappropriate modification of the control or lack of
reactivity by the driller during a facies transition may cause these untimely and often
irreversible excursions.
Chapter 2 • Geomechanics, Drilling and Production 85
description of the well-known effect of mechanical efficiency threshold [Putot, 1995, 1997].
Putot’s model represents a significant improvement over Warren’s semi-empirical
representations in case of defective hydraulics since it considers, in a simplified way, the
phenomena actually involved. Obviously, the disadvantage is that a large number of parameters
must be identified. However they all have a physical significance. The advantage of the analysis
concerns a realistic description of the conditions leading to a bit clogging mechanism.
Figure 2.1.13
Examples of damage to roller-cone bits. Top left: broken cone. Top right: worn
tooth. Bottom left: broken tooth. Bottom right: jamming.
86 Chapter 2 • Geomechanics, Drilling and Production
A dynamic version of the bit coupled with the cuttings concentration has been developed
[Putot and Constantinescu, 1997]: the state of the working face has two dynamic variables (in the
longitudinal and circumferencial directions) and a concentration variable; the solution of the
differential system combining these state variables is as much one of a problem of dynamics as
one of a problem of concentration evolution. The influence of the initial conditions and the
severity of the ramps forming the loading paths (in particular weight and torque control) have been
studied. Compared with the quasi-static version, this model offers highly significant advantages:
influence of bit dynamics on concentration evolution – and improved understanding of the drilling
system dynamics in case of clogging “problems”; but also, introduction of the rotation degree of
freedom and consequently torque taken into account in the response observation criteria.
Validation has been carried out semiquantitatively on a GdF Suez site [Putot et al., 2000].
rotates so irregularly that the drill team is clearly aware that something is wrong. In other
cases, for example in case of transverse vibrations, the anomaly is undetected at the surface
and can lead to the failure of the equipment, observed too late!
The problems due to vibrations are varied and often extremely serious. Frequently, the
damaged equipment must be brought to the surface, the borehole is irregular and, during
coring operations, the samples are damaged (e.g. core discing). Vibrations are correlated
with a reduced rate of penetration. Limiting the vibrations generally involves higher costs,
due to the need for more sophisticated equipment [Jardine et al., 1994].
It is important to dispel the notion that vibrations may benefit the rock drilling process
since, while the rate of penetration may increase in the short term, this comes at the expense
of shorter lifetime of the bit and drillstring. Although vibrations such as bit noise are used as
a seismic signal while drilling, these are usually of a different frequency and amplitude from
the destructive vibrations.
Drillstring vibrations can be divided into three types: axial, torsional and transverse:
– Axial vibrations cause bit bounce and rough drilling, behaviour that destroys bits,
damages Bottom Hole Assemblies (BHA), increases total drilling time and may be
detected at surface.
– Torsional vibrations cause irregular downhole rotation that fatigues drill collar con-
nections, damages the bit and slows drilling. The vibrations are recognised at the drill-
floor by fluctuations in the power needed to maintain a relatively constant rate of sur-
face rotation.
– The more destructive transverse vibrations may be unleashed with no sign at surface.
Deep in the hole, the BHA interacts with the borehole wall, generating very violent
transverse shocks; this leads to an out-of-gauge and rough hole and the shocks can
damage components of the BHA. The bottomhole equipment (optional) used to assess
the severity of this type of vibration significantly increases the drilling costs.
Vibrations of all three types may occur simultaneously or one at a time, depending on the
case, during rotary drilling and are more violent in vertical or low-angle wells, where the
drillstring may move more freely than in high-angle wells with a long offset.
Chapter 2 • Geomechanics, Drilling and Production 87
lations exert high cyclic stresses on the drillpipe and slow drilling [Pavone and Desplans, 1994].
The high speeds achieved during the slip phase are also one of the causes of transverse
vibrations due to centrifugal force [Théron et al., 2001].
Axial vibrations are also excited, since the axial drilling rate depends on the angular
speed. Lateral accelerations are frequently 10 times greater than axial accelerations.
However, despite the relation with rotation, analysis of transverse vibrations cannot rely
on a harmonic model. This is important since, in the past, it was believed that by “fine-
tuning” the drillstring – varying its rate of rotation and weight on bit – to avoid a resonant
frequency, transverse vibrations that have been initiated downhole may be stopped. Studies
by ANADRILL reveal, on the contrary, that this does not always work.
Field observations have led to the conclusion that transverse vibration problems appear
in hard or abrasive formations. Shocks are often seen as the MWD (Measurement While
Drilling) or adjacent stabiliser passes from a soft to a hard formation, without changing the
speed of rotation.
At the heart of the generation of transverse vibrations is a self-sustaining interaction
between the rotating drillstring and the wellbore [Théron et al., 2001]. Once initiated, this
interaction is hard to stop. Once shocks are observed downhole, varying the speed of rota-
tion will often affect only the number and severity of the shocks. Shocks are not eliminated
until the speed of rotation is reduced to a much lower value than the rate at which they com-
menced. Thus, significant shocks may be sustained at all but the lowest speeds of rotation.
Variation in transverse vibration with lithology is due to changes in the rock’s coeffi-
cients of friction and restitution. As friction between the drillstring and rock increases, more
energy per impact is transformed from rotary to transverse motion. In some cases, friction
may be exaggerated through the digging action of stabilisers clashing with the borehole.
The coefficient of restitution determines what proportion of the kinetic energy at impact
is absorbed by the formation. Low restitution results in significant energy absorption.
Because limestones and sandstones have high coefficients of friction and restitution, they
are more likely to generate high shocks than soft shales – an intuitive result.
88 Chapter 2 • Geomechanics, Drilling and Production
supported by the rock that has been removed. This results in the development of a stress con-
centration at the borehole wall. If the rock is not strong enough, the wall will fail.
In OBD, the mud density is determined to prevent a fluid influx from the formation
(Pmud > Ppore). If the mud density is too low compared with the mechanical stability crite-
rion, compressive failures could occur, possibly causing the well to collapse. The mud den-
sity must therefore be adjusted according to the most restricting criterion (pore pressure or
compressive stability).
The mud density cannot be increased infinitely, however, since if it is too high, tensile
failures caused by hydraulic fracturing will develop in the rock (Figure 2.2.1).
Friable
sandstone
Hole
collapse
Brittle
Compressive shale
failure
Hole
closure Salt
Figure 2.2.1
Failure by stress concentration around the borehole [Last & McLean, 1995].
Chapter 2 • Geomechanics, Drilling and Production 89
The mud pressure must therefore lie within a window (Figure 2.2.2) whose:
– lower limit is the larger of two values, which are the pore pressure Ppore and the mini-
mum mud pressure necessary to ensure stability of the walls Pcollapse,
– the upper limit is given by the fracture propagation pressure Pfrac in the terrain, i.e.:
max (Ppore, Pcollapse) < Pmud < Pfrac
Fracture
Stable
https://telegram.me/Geologybooks
region
Depth
Collapse
Figure 2.2.2
Mud weight window.
The mud weight window may be very narrow when the – abnormal – pore pressure is close
to the hydraulic fracturing pressure, a situation observed in numerous parts of the world, for
example the North Sea [Ward et al., 1994] and the Potwar Basin in Pakistan [Kadri, 1991].
The true mud pressure in the borehole depends on the static weight of the mud column
increased by the dynamic effect of the flow (known as ECD – Effective Circulating
Density), together with occasional fluctuations as the drillstring moves (pistoning or
suction). In view of these fluctuations, the borehole stability conditions are often borderline.
In OBD, the aim is to work at a pressure slightly above the lower limit for several reasons:
– to detect impregnated reservoirs,
– to limit the risks of fracturing and mud losses,
– to increase drilling performance.
Compressive failures are highly dependent on the in situ stress field. Failures occur pro-
gressively. Spalling appears towards the minor stresses in the first phases followed by stabi-
lisation with borehole breakout or continued enlargement of the borehole if the spalls come
loose from the wall and fall off (Figure 2.2.3).
90 Chapter 2 • Geomechanics, Drilling and Production
Diameter (mm)
200 240 280 320 360
s 2385
mm NS
3°W
Lon
ga 2395
xis
mm
=3
40
24
mm
Depth (m)
=2 2405
xis
rt a
o
Sh
2425
https://telegram.me/Geologybooks
Figure 2.2.3
Schematic illustration of spalling in the damaged area around a borehole.
Detection of borehole breakout using a Calliper [McLellan, 1994].
Maury and Sauzay (1987) identify more precisely three destabilisation mechanisms cor-
responding to three different stress models at the wall (sa = axial stress, sq = tangential
stress, radial stress = mud pressure):
– mode A, with “intermediate” axial stress (Pmud < sa < sq), indicated by the presence
of narrow, elongated spalls,
– mode B, with “intermediate” tangential stress sq (Pmud < sq < sa), producing curved
spalls concentric with drilling,
– mode C, shear failure when the mud pressure is high enough to open the natural frac-
tures (sq < Pmud < sa), generating small wedge-shaped spalls (Figure 2.2.4),
– mode D, extension failure.
The temperature variations induced by the circulation of mud may have considerable
influence on borehole stability [Maury, 1994]. Local cooling (at well bottom) releases the
compressive stresses around the borehole, possibly leading to temporary stability which will
disappear if the temperature rises (e.g. if the mud circulation stops). Inversely, local heating
(at the top of the well) may cause instabilities. Maury mentions that, in some cases, mud
cooling devices have proved an efficient way of reducing instability. It is important to know
the temperature distribution, or at least have an estimation (Figure 2.2.5). A number of soft-
ware programs capable of predicting the temperature variations in the borehole are now
available for operators.
Chapter 2 • Geomechanics, Drilling and Production 91
Internal pressure
Intermediate stress (or mud weight)
Mode A
"A" shear
Borehole elongation
https://telegram.me/Geologybooks
Intermediate
Medium Medium
"B" shear
High
"C" shear
Very high D
"D" extension
Hydraulic fracturing
Figure 2.2.4
Failure modes and spalling process [Maury & Sauzay, 1987].
92 Chapter 2 • Geomechanics, Drilling and Production
Temperature (°C)
20 40 60 80 100 120 140 160 180
100
17"1/2 Drilling phase (Q —~ 3300 l/min)
12"1/4 Drilling phase (Q —~ 2400 l/min)
2000
Static
17"1/2
Depth (m)
temperature
13"3/8 3000
70
12"1/4
4000
9"5/8
85
https://telegram.me/Geologybooks
8"1/2 5000
173°C/
7" liner 135 5000m
Figure 2.2.5
Temperature profiles during mud circulation [Maury, 1994].
The choice of drilling fluid is important for the swelling shales; a high electrolyte con-
centration may reduce crystalline swelling but it is recommended to use polymers and seal
narrow fractures with gilsonite.
1.5
Normalised enlarged hole size
1.4
1.3
1.2
1.1
Model-predicted extent of yielded zone
https://telegram.me/Geologybooks
1.0
0 10 20 30 40 50
Time (day)
Figure 2.2.6
Hole enlargement against time in a shale formation
[Hawkes and McLellan, 1997].
Bedding
plane
Spalling Bedding
breakout plane
1m
Figure 2.2.7
https://telegram.me/Geologybooks
In this type of formation, changing from a water-base mud to an oil-base mud has no
effect and increasing the mud density favours mud invasion of the fracture network
[Santarelli et al., 1992]. This problem is solved by adapting the mud rheology.
In some extreme cases, fault invasion will reduce the local effective stresses and the
shear stresses may cause lateral displacement of the borehole axis [Maury, 1994] with the
consequences we can well imagine: drillstring sticking, or even breaking (Figure 2.2.8).
u
sn
a) t b)
Consequences: Consequences:
Tight hole Stuck in holes
Difficulties to run-in-hole Difficulties to pull out of hole
Abnormal torque Drillstring failures (?)
Need for reaming Need for back reaming
Figure 2.2.8
Lateral displacement of the borehole axis induced by shear relaxation along
pre-existing faults or fractures [Maury, 1994].
Chapter 2 • Geomechanics, Drilling and Production 95
Salt
washout
Salt
Dolomite
Salt Kicks
movement
Brine/oil/
gas/H2S
Fracture Filtration
losses losses
Mud /
brine
Mud /
brine
Figure 2.2.9
Instabilities in salt [Williamson et al., 1997].
96 Chapter 2 • Geomechanics, Drilling and Production
et al., 1992; Charlez, 1997]. The main equations given by Aadnoy and Chenevert (1987)
will be mentioned below.
The following assumptions are made: linear elasticity and isotropy of the in situ rock mechani-
cal properties. We assume that the principal stresses in the formation are: sv vertical stress, sH
maximum horizontal stress, and sh minimum horizontal stress. Stresses near the wellbore are
described in a coordinate system where the z-axis is parallel to the wellbore and the y-axis is hori-
zontal (Figure 2.2.10). Transformation of coordinates (sv, sH, sh) to (x, y, z) is carried out by two
rotations, a rotation “a” around the sv axis and a rotation “i” around the y-axis. Angle i represents
the wellbore deviation and angle a the azimuth with respect to the maximum horizontal stress.
Pw represents the mud pressure in the wellbore of radius R.
sv z
sh i
y
sH a
sz
x
Pw
sr
sq
Figure 2.2.10
Transformation system and notation.
Chapter 2 • Geomechanics, Drilling and Production 97
The stresses in cylindrical coordinates are given by the following expressions as a func-
tion of the distance r to the centre of the wellbore.
⎛ σx + σ y ⎞ ⎛ R2 ⎞ ⎛ σ x − σ y ⎞ ⎛ R4 R2 ⎞
σr = ⎜ ⎟ ⎜1 − 2 ⎟ + ⎜ ⎟ ⎜ 1 + 3 4 − 4 2 ⎟ cos 2θ
⎜⎝ 2 ⎟⎠ ⎝ r ⎠ ⎜⎝ 2 ⎠⎟ ⎝ r r ⎠
⎛ R4 R2 ⎞ R2
+ τ xy ⎜ 1 + 3 − 4 ⎟ sin 2θ + p
⎝ r 4 2
r ⎠ r2 w
⎛ σx + σ y ⎞ ⎛ R2 ⎞ ⎛ σ x − σ y ⎞ ⎛ R4 ⎞ ⎛ R4 ⎞ R2
σθ = ⎜ ⎟ ⎜1 + 2 ⎟ − ⎜ ⎟ ⎜ 1 + 3 4 ⎟ cos 2θ − τ xy ⎜ 1 + 3 4 ⎟ sin 2θ − 2 pw
⎜⎝ 2 ⎟⎠ ⎝ r ⎠ ⎜⎝ 2 ⎠⎟ ⎝ r ⎠ ⎝ r ⎠ r
( ) Rr cos 2θ − 4ντ
2 R2
σ z = σ zz − 2ν σ x − σ y sin 2θ
https://telegram.me/Geologybooks
2 xy
r2
⎡⎛ σ − σ ⎞ ⎤⎛ R4 R2 ⎞
x y
τrθ = ⎢⎜ ⎟ sin 2θ + τ xy cos 2θ ⎥ ⎜ 1 − 3 4 + 2 2 ⎟
⎢⎜⎝ 2 ⎟⎠ ⎥⎝ r r ⎠
⎣ ⎦
⎛ R2 ⎞
τrz = ⎡τ xz cos θ + τ yz sin θ ⎤ ⎜ 1 − ⎟
⎣ ⎦⎝ r2 ⎠
⎛ R2 ⎞
τθ z = ⎡ −τ xz sin θ + τ yz cos θ ⎤ ⎜ 1 + ⎟
⎣ ⎦⎝ r2 ⎠
With
( )
σ x = σ H cos2 a + σ h sin2 a cos2 i + σv sin2 i
(
σ y = σ H sin2 a + σ h cos2 a )
( )
σ zz = σ H cos2 a + σ h sin2 a sin2 i + σv cos2 i
( ) )
τ yz = 0.5 σ H − σh sin ( 2 a sin i
( )
τ xz = 0.5 σ H cos2 a + σ h sin2 a − σv sin 2i
( ) )
τ xy = 0.5 σ H − σh sin ( 2 a cos i
98 Chapter 2 • Geomechanics, Drilling and Production
( )
σ z = σ zz − 2ν σ x − σ y cos 2θ − 4ντ xy sin 2θ
τrθ = 0
τrz = 0
τθ z = 2 ⎡ −τ xz sin θ + τ yz cos θ ⎤
⎣ ⎦
In an anisotropic horizontal stress field (sH max > sh min), concentrations of maximum
circumferential stress sq are observed on the wellbore walls in the direction of the minimum
https://telegram.me/Geologybooks
horizontal stress (Figure 2.2.11). We can therefore see that breakouts will occur in these
zones, depending on the hardness of the rock.
160
sqq
140
Stress at wellbore wall (MPa)
120
100
80
60
szz
40
20
srr
0
0 90 180 270 360
Angle around the hole (from south)
Figure 2.2.11
Stresses sq around a vertical wellbore under an east-west compressive stress
field [Zoback et al., 2003].
σ + σz 1
σ1 = θ + (σθ − σ z )2 + 4τθ2zz
2 2
σ + σz 1
σ2 = θ − (σθ − σ z )2 + 4τθ2zz
2 2
σ3 = pw
Chapter 2 • Geomechanics, Drilling and Production 99
Any compressive failure criterion can be applied to estimate the collapse failure: Mohr
Coulomb, which does not take into account the intermediate stress, Von Mises, Drucker
Prager, etc. [McLean and Addis, 1990].
These models have demonstrated in particular that horizontal wellbores are mechanically
more stable if they are drilled parallel to the maximum horizontal stress [Zoback et al., 2003].
Sand production designates transport of grains by the production fluid, from the reservoir
rock into a wellbore.
In gas and light oil reservoirs, or in offshore production, above a certain proportion, sand
production causes a number of undesirable problems such as damage to wellbore pumps and
wellhead erosion (a leak may be created by sandblasting in less than an hour in a gas well),
plugging of perforations or even total invasion of the production column. Several methods
are available to address these problems, but they are expensive to implement and often
reduce the production rate.
In heavy oil reservoirs, if sand production prevention systems are used, the oil flow rates
are considerably reduced (from 7/15 m3 a day to 0.5 m3 a day, for example) and the well
becomes uneconomical. In this case, it is acceptable to produce oil and sand simultaneously,
as with the Cold Heavy Oil Production with Sand (CHOPS) method which may prove
profitable under certain conditions. Due to the high viscosity and low production rates,
pump and valve erosion problems are minor and the sand can be transported up to the
surface without difficulty using progressive cavity pumps. If thermal stimulation is used, for
example Steam Assisted Gravity Drainage (SAGD), the oil viscosity is lowered and systems
must be installed to prevent sand production.
In all cases, therefore, anticipating the onset of sand production is a major challenge.
The physical problem is extremely complex and involves numerous parameters: rock
and reservoir characteristics, wellbore completion type and production conditions. Some
triggering mechanisms are well-known, for example damage to soft formations by drilling
or depletion, and damage under the effect of hydrodynamic forces; other, such as water cut,
erosion at the surface of cavities, etc., are much less well-known.
Next Page
vary considerably, for example a sand volume of 1 to 200 L over a period of 1 to 500 hours.
Geomechanics
and Reservoir
https://telegram.me/Geologybooks
It has become standard practice in reservoir engineering to use a geocellular 3D model for
the reservoir and the surrounding rocks in order to study the possible combined effects
between fluid production and the mechanical strains of the structure. It is in fact very impor-
tant to predict possible subsidence of the surface as oil and gas production proceeds.
Even a slight variation in pore pressure in a reservoir may cause significant displace-
ments in the subsurface layers, especially if the reservoir is very poorly or weakly cemented,
which increases the risk of collapse and the development of shear zones. Combined prob-
lems such as these are encountered when studying the stability of CO2 sequestration sites,
where it is important to check that the injection of large quantities of gas into the structure
will not destroy its long-term mechanical stability.
Exploitation of an oil reservoir is very expensive and entails a high degree of uncer-
tainty. These costs and uncertainties can be reduced by reservoir studies conducted by the oil
companies before or during exploitation. These reservoir studies are generally based on
numerical simulations of flows in the reservoir in order to predict the quantities of hydrocar-
bons present and their production over time, optimise reservoir production and assess the
risks associated with production. Reservoir simulators give a highly accurate description of
the fluid and heat transfers in the reservoir but do not model (or do so in a very simplified
manner only) the mechanical behaviour of the reservoir during its production. Production in
certain types of reservoir (poorly consolidated reservoirs, fractured or faulted reser-
voirs, etc.) is accompanied by major mechanical effects likely to threaten the integrity of the
reservoir and its wells, generate subsidence of the surface layers and modify the flows of the
reservoir fluids, and therefore any production predictions made by the reservoir engineers.
Modelling these mechanical effects therefore requires a geomechanical model which must
be coupled with the reservoir simulation.
We will first describe the hydraulic diffusivity equation, a classical model of the single-
phase and isothermal flows in porous media. This diffusivity equation involves two
important parameters: the compressibility and the permeability reservoir rock. We will then
demonstrate that these two parameters may be highly dependent on the stress state in the
reservoir. In this case, it becomes essential to model the mechanical behaviour of the oil
136 Chapter 3 • Geomechanics and Reservoir
reservoir during production. We then describe how the coupling between the reservoir and
geomechanical models can be achieved. Two examples of coupling between reservoir and
geomechanical simulators are described for a high porosity reservoir and a faulted reservoir.
Lastly, we will study the evolution of stresses in the reservoir during depletion and the
measurements available to validate it.
fluids in porous media (see Bear and Buchlin, 1991, for example). To simplify the
description, we will restrict ourselves to single-phase isothermal flows in porous media. We
write f the porosity of the porous medium (Eulerian porosity, which is the ratio of the pores
to the current total volume) and r the density of the saturating fluid. The conservation of
fluid mass can then be expressed as:
∂
( ρφ ) + div ( ρφ v ) = 0 (3.1.1)
∂t
with v the seepage velocity of the fluid in the porous medium. The latter is generally related
to the fluid pressure p via Darcy’s law which, in the absence of gravity, can be written:
k
φv = − grad p (3.1.2)
μ
with k the rock intrinsic permeability and m the fluid dynamic viscosity.
The fluid is assumed to be compressible, of compressibility Cf defined by:
∂ρ = ρ0C f ∂p (3.1.3)
with r0 the density of the fluid under fluid pressure p0.
A reservoir rock “compressibility” Cr is introduced at the same time, to take into account
a variation of the porous medium porosity as a function of the saturating fluid pressure:
∂φ = φ0Cr ∂p (3.1.4)
with f0 the initial porosity of the porous medium given for the initial fluid pressure p0. Intro-
ducing equations (3.1.2) to (3.1.4) in the conservation of mass equation (3.1.1), we obtain
after linearisation:
∂p k
φ0 (C f + Cr ) − Δp = 0 (3.1.5)
∂t μ
Chapter 3 • Geomechanics and Reservoir 137
Associated with representative boundary and initial conditions, equation (3.1.5) gives the
evolution of saturating fluid pressure in the porous medium. It is important to note, however,
that the compressibility Cr is not an intrinsic property of the reservoir rock but depends on
the stress path followed by the reservoir during its depletion. In addition, the permeability of
the porous medium may vary significantly during depletion of a reservoir due to changes in
effective stress in the reservoir.
Equation (3.1.5) shows that the evaluation of recovery is based mainly on the estimation of
reservoir rock compressibility, which represents the strain of the porous volume induced by
the variation in reservoir pressure:
1 ΔVp
Cr =
https://telegram.me/Geologybooks
(3.1.6)
Vp Δpp
So defined, the reservoir rock compressibility depends on the stress path followed by the
reservoir during production. This stress path may be quite different from the isotropic or
oedometric loading paths traditionally adopted in the laboratory to evaluate Cr. In particular,
it may be heterogeneous and vary during depletion (see Chapter 3.1.3).
Concerning this point, note that the compressibilities defined by Zimmerman describe
exclusively the volume response of a rock subjected to hydrostatic loading σ = − pc 1 and to
a pore pressure pp [Zimmerman 1991]:
1 ∂Vb 1 ∂Vp
Cbc = − Cpc = −
Vb ∂p pp Vp ∂p pp
c c
1 ∂Vb 1 ∂Vb
Cbp = − Cpp = − (3.1.7)
Vb ∂p pc Vb ∂p pc
p p
where Vp designates the porous volume and Vb the bulk volume, the compressibility closest
to definition (3.1.6) being obviously Cpp.
Assuming linear poroelastic behaviour of the reservoir rock, the compressibilities so
defined can be expressed as a function of the drained bulk modulus Ko and Biot’s
coefficient b [Boutéca, 1992]:
1 b Cbp
Cbc = Cpc = =
K0 φ0 K0 φ0
b b − φ0 (1 − b )
Cbp = Cpp = (3.1.8)
K0 φ0 K0
138 Chapter 3 • Geomechanics and Reservoir
1 1
Cbc = Cbp = Cpc = Cpp = (3.1.9)
K0 φ0 K0
Still assuming linear poroelastic behaviour of the reservoir rock, we can determine
expressions for Cr as a function of the loading path considered. The three types of loading
path generally followed to simulate depletion in the laboratory are the hydrostatic,
oedometric and proportional paths (see Chapter 3.1.3):
1 ⎛ b2 b − φo ⎞
Crhydrostatic = ⎜ + ⎟
φo ⎝ Ko Ks ⎠
⎛ ⎞
1 ⎜ b2 b − φo ⎟
https://telegram.me/Geologybooks
Croedometric = ⎜ + ⎟ (3.1.10)
φo ⎜ 4G Ks ⎟
Ko +
⎝ 3 ⎠
1 ⎛ 1 + 2 K b2 b − φ ⎞
o
Crproportional = ⎜ + ⎟
φo ⎝ 3 Ko
Ks ⎠
For a hydrostatic loading path, we obtain the expression for Cpp expressed in this case as
a function of the bulk modulus of the solid matrix.
Table 3.1.1 compares the compressibilities estimated from expressions (3.1.8) and
(3.1.10). The poroelastic parameters used are extracted from the thesis of D. Bévillon (2000)
and are representative of a highly compactable reservoir chalk. Note that the values obtained
largely depend on the assumptions made concerning the loading path.
Table 3.1.1 Estimation of chalk compressibility as a function of the loading path followed.
In addition, for some rocks, the compressibility may depend, like the drained bulk modu-
lus, on the pore pressure and stress state [Bemer et al., 2001].
Chapter 3 • Geomechanics and Reservoir 139
also that of the flow. Consequently, when the authors observe an increase in permeability
with the dilatancy of the sample, the permeability is measured parallel to the main axis of the
cracks. The petroleum industry, however, is more concerned by the change in horizontal
permeability.
Schutjens and de Ruig (1996) used a special cell to measure the change in radial perme-
ability (flow perpendicular to the preferential orientation of the cracks). Sandstone samples
with high initial porosity and permeability levels (porosity between 21% and 32%, radial
permeability between 0.1 mD and 500 mD) have been tested under hydrostatic and oedo-
metric loading paths. The true depletion tests showed a decrease in radial permeability.
However, the test conditions (loading paths and initial permeability levels of the samples)
correspond a priori to cases where the axial permeability also decreases. By comparing tests
with measurement of the axial permeability and tests with measurement of the radial perme-
ability, conducted on the same rock of low initial porosity and along a deviatoric loading
path with low mean effective stress, it would be possible to check whether the radial perme-
ability increases in the same way as the axial permeability.
Only very few data are available in the literature on the behaviour of a rock subjected to
a real triaxial loading path and on the anisotropy of the permeability tensor. Al-Harthy et al.
(1998) conducted depletion tests on various sandstones for three different stress states: true
triaxial (po; 0.8*po; 0.6*po), triaxial (po; 0.8*po; 0.8*po) and hydrostatic (po; po; po). The
results have shown a decrease in the vertical permeability (direction of the major principal
stress) with the drop in pressure, this loss of permeability being lower for true triaxial load-
ing than for traditional triaxial loading or hydrostatic loading.
Al-Harthy et al. (1999) completed these results by studying, still for sandstones, the
anisotropy of the permeability tensor and the influence of the loading path on the change in
permeabilities in the three main directions of the true triaxial test. For samples exhibiting
horizontal stratification, they measured vertical anisotropy factors (ratios between the verti-
cal permeability and a horizontal permeability) of up to 40; this factor is obviously greater
for true triaxial loading (1.25*po; po; 0.75*po) than for hydrostatic loading (po; po; po). The
horizontal anisotropy factors (ratios between the two horizontal permeabilities) measured
140 Chapter 3 • Geomechanics and Reservoir
are more limited and lie between 1.1 and 1.7, but this difference in horizontal permeabilities
may prove extremely important when installing a horizontal well. The results obtained do
not indicate a clear trend as to the influence of depletion on the anisotropy of permeabilities;
a slight decrease in anisotropy with increase in effective stresses has nevertheless been
observed on a reservoir sample exhibiting horizontal stratification.
Some authors have tried to define variation laws (mainly empirical) expressing perme-
ability as a function of the stress path followed. The small permeability variations in the
elastic range can be estimated using the experimental law proposed by David et al. (1994):
( )
⎡− a p '− p ' ⎤
k = ko e⎣ o ⎦
defined by the derivative of the function p ' = f ( q) (effective mean stress – deviatoric stress):
1 ∂k ∂p '
− =a +b
k0 ∂q ∂q
where ko is the permeability before damage.
For low porosity rocks, permeability is likely to increase as the rock is subjected to
shear-induced dilation. This increase may be related to the strains in the rock, the dilating
strain in a given direction being associated with the width of the crack created.
In case of ductile failure of the reservoir rock (very unlikely in situ), no permeability
reduction law is available.
Currently therefore, there is no generic law describing the change in intrinsic permeabil-
ity of a reservoir rock as a function of the stress path.
strain) for each mesh of the reservoir model [Boutéca, 1992]. Some simulators nevertheless
allow these parameters to vary with the reservoir fluid pressure.
The main geomechanical effects affecting production are the compaction and associated
increase in the recoverable quantities of fluid. Obviously, these effects are more noticeable
for highly compactable reservoir rocks (poorly cemented sandstones generally qualified as
poorly consolidated, highly porous chalks, etc.). Compaction is due to the decrease in reser-
voir pressure and the associated increase in effective compressive stresses in the reservoir.
The mechanical behaviour of poorly consolidated reservoir rocks is complex, being highly
nonlinear and depending on the stress path and the temperature. With highly porous chalks,
the mechanical behaviour of the reservoir rock may also be considerably affected by varia-
tions in water saturation (“water weakening”) [Hermansen et al., 2000; Homand, 2000;
Matà, 2001]. Compaction phenomena play a significant role in the production of reservoir
fluids, as demonstrated by the Bachaquero fields [Merle et al., 1976] and Zuata [Charlez,
1997] in Venezuela. For highly compactable reservoirs, therefore, it is essential to predict
the change in reservoir porosity during production using strain or stress data provided by a
geomechanical model associated with the reservoir model.
The stress variations resulting from reservoir production modify the porous structure of
the rock and therefore its permeability. The permeability of highly consolidated and natu-
rally fractured or faulted reservoirs is extremely sensitive to stress changes during produc-
tion. The thermoporoelastic effects in these reservoirs may modify the conductivities of the
fractures and make waterflooding much less effective [Guttierez & Makurat, 1997]. For
these reservoirs, the thermoelastic effects are generally concentrated around injector wells,
where the temperature variations are higher, whereas hydromechanical effects affect the
entire reservoir. The thermohydromechanical behaviour of the fracture system is complex.
Macroscopic behaviour depends on the density and orientation of the fracture system, the
initial stress state and the stress path in the reservoir during production. Due to the high elas-
tic rigidity of the matrix (rock intact), the strains are mainly located on the fracture planes,
thereby generating a change in their hydraulic conductivity [Koutsabeloulis et al., 1994;
Heffer et al., 1994]. Geomechanical modelling of the reservoir is therefore required to esti-
mate the permeability changes in the reservoir, whether they are related to changes in the
porous structure or to activation of fractures or faults.
142 Chapter 3 • Geomechanics and Reservoir
with M Biot’s modulus, b Biot’s coefficient, G the rock shear modulus and K0 the drained
bulk modulus of the rock. σ and ε designate respectively the stress and strain tensors, 1
the Kronecker symbol, p the pore pressure and m the fluid mass intake with respect to the
initial total volume. The first poroelastic constitutive law is introduced in the mechanical
equilibrium equation which, in the absence of body force, is written:
div σ = 0 (3.1.13)
The strain tensor is therefore expressed as a function of the displacement vector u . The
previous equation therefore becomes:
G
G Δu + ( K0 + )grad u = b grad p (3.1.14)
3
In addition, the fluid mass intake m with respect to the fluid initial volume is expressed
from the mass balance and Darcy’s law as:
∂m ⎛ k ⎞
∂t
( )
= − div ρφ v = − div ⎜ ρ grad
⎝ μ
p⎟
⎠
(3.1.15)
The latter equation is introduced in the second poroelastic constitutive law, to give after
linearisation:
1 ∂p k ∂ε
− Δp = − b v (3.1.16)
M ∂t μ ∂t
with εv = tr ε = div u , the volume strain.
Chapter 3 • Geomechanics and Reservoir 143
The geomechanical problem therefore consists in solving the following system of equa-
tions system which describes the flows and elastic strain of the porous medium [see also
Boutéca, 1992; Coussy, 1995; Lewis and Schrefler, 1998]:
G
G Δu + ( K0 + )grad u = b grad p (3.1.17)
3
1 ∂p k ∂ε
− Δp = − b v (3.1.18)
M ∂t μ ∂t
The pore pressure p is the main unknown of equation (3.1.18) and the displacement u is
the main unknown of equation (3.1.17). The displacement u appears in the right member of
the “hydraulic” equation (3.1.18) via the volume strain. Similarly, the fluid pressure modi-
fies the mechanical equilibrium via the right member of equation (3.1.17). Equations
(3.1.17) and (3.1.18) are therefore coupled.
https://telegram.me/Geologybooks
⎡K L ⎤ ⎡ Δt u ⎤ ⎡ F ⎤
⎢ T ⎥⎢ ⎥=⎢ ⎥ (3.1.19)
⎢⎣ L E ⎥⎦ ⎢⎣ Δt p ⎥⎦ ⎣ R ⎦
with:
– u the discretised mechanical unknown,
– p the discretised pressure unknown,
– Δt the time difference operator,
– K the mechanical rigidity matrix or stiffness matrix,
– L the coupling matrix between mechanical and hydraulic unknowns,
– F the force vector,
– E the hydraulic flow matrix in porous medium,
– R the source term of the flow problem.
The hydraulic flow matrix E is generally decomposed in the form E = T – D, where T is a
transmissivity matrix and D the accumulation diagonal matrix. It is important to note that in
the case described, matrices K and E are considered as linear operators whereas for a more
realistic problem (elastoplastic behaviour, compaction, multiphase flow, etc.), these opera-
tors are nonlinear.
Various coupling levels can be considered when solving the matrix system (3.1.19):
– Conventional reservoir approach: in this approach, the coupling term in the second
equation of matrix equation (3.1.19) is reformulated as a function of the pressure
variation so that matrix D uses the rock and fluid compressibilities. The limitations of
144 Chapter 3 • Geomechanics and Reservoir
this approach have already been discussed. It can nevertheless be used to obtain an
estimation of the stress variations when the pressure variations calculated by the reser-
voir simulator are transmitted to the second member of the first equation of matrix
equation (3.1.19) via a geomechanical simulator. In this case, we speak of one-way
coupling or decoupling.
– Partially coupled: in this approach, the equations of matrix system (3.1.19) are
solved separately on different simulators. A reservoir simulator solves the hydraulic
problem and a mechanical simulator the geomechanical problem. Coupling is carried
out by exchanging results between the two simulators at different exchange periods.
In practice, coupling consists in calculating the pressure field p from the displacement
field u calculated during the previous geomechanical simulation with the equation:
E Δt p = R − LT Δt u (3.1.20)
This first calculation is carried out with a conventional reservoir simulator using a
https://telegram.me/Geologybooks
porosity correction calculated from a mechanical variable [Settari and Mourits 1994,
Settari and Mourits 1998, Bévillon 2000, Mainguy and Longuemare 2002]. Then, using
the pressure field found using (3.1.15), the displacement vector is calculated from:
K Δt u = F − L Δt p (3.1.21)
reservoir and its finite volume mesh are shown on Figure 3.1.1. The finite element mesh
used in ABAQUS includes the reservoir and the side, over and underburdens (Figure 3.1.2).
The simplified reservoir geometry and the choice of finite volume and finite element
discretisations simplify data exchange between the two models.
1000 m
Borehole
https://telegram.me/Geologybooks
100 m
3000 m
Figure 3.1.1
Reservoir mesh composed of 55 finite volume cells.
7000 m
3000 m
100 m
900 m
9000 m
Figure 3.1.2
Geomechanical mesh composed of 1296 Q20 finite elements.
146 Chapter 3 • Geomechanics and Reservoir
The reservoir pressure is initially 48 MPa and the reservoir is produced from the central
producing well whose bottomhole pressure is fixed at 10 MPa. The reservoir fluid is
assumed to be poorly compressible. The reservoir porosity is initially high (40%) and the
mechanical behaviour is assumed to be nonlinear and represented by a Cam-Clay model.
The behaviour of the surrounding formations is represented by a linear poroelastic model
which provides the changes in pressure and stress field. Porosity and permeability are low in
the surrounding formations.
https://telegram.me/Geologybooks
Figure 3.1.3
Deforming finite element mesh after 33 years production.
U3 Value
-7.93E-01
-7.32E-01
-6.71E-01
-6.10E-01
-5.49E-01
-4.88E-01
-4.27E-01
-3.66E-01
-3.05E-01
-2.44E-01
-1.83E-01
-1.22E-01
-6.10E-02
+2.73E-23
Figure 3.1.4
Vertical displacement isovalues on the surface after 33 years production.
Chapter 3 • Geomechanics and Reservoir 147
The results of the coupled simulations indicate compaction of the reservoir rock resulting
as surface subsidence (Figures 3.1.3 and 3.1.4). After 33 years production, the vertical
displacement reaches a maximum value of 75 cm on the surface and in the centre of the
reservoir. The hydromechanical calculations provide the change in pore pressure in the
surrounding formations. These results indicate overpressure in the area of the surrounding
formations located at the periphery of the reservoir, reaching 7 MPa afters 12 years
production (Figure 3.1.5). This overpressure is due to stress transfer from the reservoir to the
area of the surrounding formations in contact with the reservoir, since their rigidity is higher
than that of the reservoir. This phenomenon is generally called “arch effect” or “stress
arching”, [Osorio et al., 1998]. This stress transfer increases the pore pressure near the
reservoir, which cannot be dissipated due to the low permeability of the surrounding
formations.
https://telegram.me/Geologybooks
Figure 3.1.5
Pore pressure isovalues in the reservoir and sideburdens after 12 years
production.
higher recovery than the reservoir calculations at start of simulation (which indicates that the
reservoir calculation underestimated the effects of pore compression), then lower recovery
after a long period of time (which indicates that the reservoir calculation overestimated the
effects of pore compression).
16
Cp oedometric local
14
Cp mean
Cp elastic
12
Cp (1.E-9Pa-1)
10
6
https://telegram.me/Geologybooks
0
0 5 10 15 20 25 30 35 40 45 50
Pore pressure (MPa)
Figure 3.1.6
Various pore compressibilities used in conventional reservoir calculations.
24
22 Partial coupling
Cp elastic
20
Cp mean
18 Cp elastoplastic
Recuperation (%)
16
14
12
10
8
6
4
2
0
0 2000 4000 6000 8000 10000 12000
Time (day)
Figure 3.1.7
Comparison of the coupled calculation with various conventional reservoir cal-
culations using various compressibilities [Bévillon, 2000].
Chapter 3 • Geomechanics and Reservoir 149
Figure 3.1.8
Reservoir geometry [Longuemare et al., 2002].
Figure 3.1.8 shows the reservoir geometry taken into account in the reservoir simulator. The
reservoir model is a simple porosity model including faults whose permeability is described by
transmissivity multipliers. The faults taken into account in the geomechanical model associated
with the reservoir model are shown on Figure 3.1.9. Figure 3.1.10 shows the distribution
between the two types of rock identified in the reservoir according to their dolomite content.
The first type, with low dolomite content, has smaller drained Young’s modulus and effective
cohesion (Eo = 12 GPa, C' = 4 MPa) than the second (Eo = 24 GPa, C' = 11.3 MPa).
The reservoir simulations model the thermohydraulic behaviour of the reservoir over a
period of 5 years. The geomechanical model calculates the associated stress variations and the
strains normal and tangential to the fault planes (Figure 3.1.11). The latter are related to the
permeability changes in the faults via a phenomenological law [Koutsabeloulis and Hope,
1998]. The permeability changes in the faults are then used to update the fault transmissivities
included in the reservoir model. Figures 3.1.12 and 3.1.13 show the fault transmissivity
multipliers of the reservoir model after 5 years production. The activated faults (whose
permeability is increasing) correspond to the faults on Figure 3.1.11 stressed due to the pressure
and temperature variations in the reservoir and their orientation with respect to the direction of
maximum stress.
150 Chapter 3 • Geomechanics and Reservoir
Figure 3.1.9
Faults of the geomechanical model [Longuemare et al., 2002].
https://telegram.me/Geologybooks
Figure 3.1.10
Distribution of the two types of rock in the reservoir (dark blue E = 12 GPa,
light blue E = 24 GPa) [Longuemare et al., 2002].
Model: TFEA00
L005: Time/Months 59,90
Nodal 1-STRN Normal
Max = .43E-4 Min = 0
.391E-4
.352E-4
.313E-4
.274E-4
.234E-4
.199E-4
.156E-4
.117E-4
.781E-5
.391E-5
Figure 3.1.11
Strain isovalues normal to the fault planes after 5 years production
[Longuemare et al., 2002].
Chapter 3 • Geomechanics and Reservoir 151
Figure 3.1.12
Transmissivity multiplier in the x direction [Longuemare et al., 2002].
https://telegram.me/Geologybooks
Figure 3.1.13
Transmissivity multiplier in the y direction [Longuemare et al., 2002].
Δσ h
γ = (3.1.20)
Δpp
with Δpp < 0 for depletion. Note that expression (3.1.20) is not based on any behaviour or
boundary condition assumption.
Some authors adopt a different definition of the stress path, expressed as a function of
the variations in minimum effective horizontal stress sh' and effective vertical stress sv'
[Rhett and Teufel 1992a] [Schutjens and de Ruig 1996] [Hettema et al., 1998] [Schutjens
et al., 1998] [Yale and Crawford, 1998]. Giving, assuming elastic behaviour,
Δσ 'h Δ(σ h − bpp )
K= = (3.1.21)
Δσ 'v Δ(σv − bpp )
where b is Biot’s coefficient.
If we assume that the axial stress does not vary during depletion (no stress arching), we
https://telegram.me/Geologybooks
vo
K= (3.1.23b)
1 − vo
where no is the drained Poisson’s ratio.
In situ measurements taken on different fields (sandstones and chalks) show that the
Δσ h
stress path γ = followed by a reservoir during depletion varies between 0.2 and 1
Δpp
[Addis et al., 1998] and may be quite different from the path predicted assuming uniaxial
compaction. Thus, for the Groningen field (consolidated sandstone with 15%-20%
porosity), the stress path estimated using in situ measurements is γ = 0.4 , to be compared
with the value of 0.8 obtained using equation (3.1.23a) [Hettema et al., 1998]. The compila-
tion of 20 years in situ data gives a stress path K = 0.2 for the Ekofisk field, whereas uniax-
ial compaction tests in laboratory lead to values ranging from K = 0.4 for chalks in the Tor
formation, to K = 0.5 − 0.6 for the Lower Ekofisk formation (Rhett and Teufel, 1992a).
Chapter 3 • Geomechanics and Reservoir 153
It is important to note that use of expressions (3.1.23) to predict the stress path during
depletion from laboratory measurements is based on strong assumptions:
– uniaxial reservoir compaction,
– elastic behaviour of the rock during depletion,
– homogeneous drained Poisson’s ratio at reservoir scale, identical to that measured at
laboratory scale.
Various factors influencing reservoir behaviour during depletion question the above
assumptions; they are described in the next paragraph. From the point of view of scale
effects between the laboratory samples and the reservoir, note for example that in order to
calibrate the predicted subsidence values with in situ data, the rock stiffness obtained in the
laboratory must be divided by 2 for Norwegian chalk reservoirs and, inversely, multiplied
by 2 for the Groningen field sandstone [Santarelli et al., 1998]. Upscaling from laboratory
data to reservoir scale is a central problem, and one of increasing interest to rock mechanics
engineers in the petroleum community.
https://telegram.me/Geologybooks
Rr
– for = 0.1, a stress path γ = 0.84 independent of the ratio between the drained
Rh
Young’s moduli of the reservoir rock and the host rock;
Rr
– for = 0.01, a stress path varying between γ = 0.77 for E0r = E0h and γ = 0.93
Rh
E0h
for E0r = .
20
Kenter et al. (1998) generated a 3D mesh of the Shearwater field using seismic data,
producing a better representation of the true geometry of the reservoir, the surrounding
154 Chapter 3 • Geomechanics and Reservoir
layers, and the fractures. They demonstrated that, due to the special geometry of the
reservoir studied (inclined and elongated), depletion induces stress arching which reduces
the stresses at the top of the reservoir and increases them at its boundaries.
voir rock elastic limit is reached. The stress path (calculated according to the stress and pres-
sure state before depletion) changes from g = 0.7 – 0.8 when the plasticity criterion is
reached to g = 0.3 – 0.4 at end of depletion.
Influence of fractures
The presence of fractures may result in a non-homogeneous stress path in the reservoir. San-
tarelli et al. (1998) report, in particular, measurements taken on a North Sea poorly consoli-
dated sand reservoir which reveal the existence of two distinct zones: the first zone follows a
stress path g ≈ 0.7, while the second follows a lower stress path g ≈ 0.4 due to the presence of
a greater number of fractures (faults observable on the seismic data).
3.2.1 Subsidence
underground fluids. The articles by Christensen et al. (1988) and Boutéca et al. (1996)
report the main cases of subsidence related to pumping in aquifers and production of hydro-
carbons. Some of the most famous cases of subsidence will be mentioned below.
stress associated with the pressure drop in the aquifer. Although the ground surface subsided
slowly, there are numerous negative effects: fractured well casings, damage to roads and canals
and irrigation networks due to differential subsidence associated with high horizontal strains.
https://telegram.me/Geologybooks
Figure 3.2.1
Joseph Poland beside a telephone pole in San Joaquin valley approximately
marking the position of maximum subsidence in 1981
(Photo: Galloway et al., 1999, p. 23).
The first subsidence problems appeared end 1984, maximum subsidence was about 3
metres at end 1985, with a subsidence rate of 40 to 45 cm/year. Numerous subsidence-
induced casing failures were observed in the reservoir and the overburden [Yudovich and
Chin, 1989]. In addition, due to the marine environment the subsidence consequences were
more expensive. During the summer of 1987, seven platforms had to be raised 6 metres due
to the significant decrease in the air gap between the platform and the sea [Sulak and
Danielsen, 1989]. Seabottom subsidence results from reservoir compaction associated with
elastic then plastic strains (pore collapse). This strain generates a high irreversible reduction
of the porosity when the pore pressure decreases [Johnson et al. 1989]. Reservoir subsidence
may be increased by consolidation of clays in the overburden. Several reinjection operations
were carried out to limit the subsidence. Some of the gas produced was reinjected from 1985
to 1987 [Sulak and Danielsen, 1989]. While this limited the subsidence rate to about 26 cm/
year, it had to be stopped for financial considerations.
Water injection was then carried out to stabilise the pressure in the reservoir and enhance
hydrocarbon recovery. Water injection was first started in the Tor formation from end 1987
https://telegram.me/Geologybooks
then extended to the Ekofisk formation from 1994. In 1996, the reservoir pressure was
perfectly stabilised (even very slightly increased) although reservoir subsidence continued,
reaching a maximum value of about 7 metres (Figure 3.2.2). This permanent strain of the
chalk at constant pressure was put down to water weakening (weakening caused by the
increase in water saturation of the chalk) [Maury et al., 1996; Sylte et al., 1999;
Matà, 2001].
300
1986
400
500
600
1996
700
800
Figure 3.2.2
Subsidence as a function of depletion for the Ekofisk field (Hetema et al.
2002).
Chapter 3 • Geomechanics and Reservoir 159
Uh
Uv
Figure 3.2.3
Idealised representation of the horizontal and vertical displacements on the
surface of a subsidence zone.
160 Chapter 3 • Geomechanics and Reservoir
A delay of a few months to a few years is frequently observed between the pressure drop
in the reservoir and its consequences on the surface in terms of subsidence. As shown on
Figure 3.2.4, the linear poroelastic model cannot account for this delay, which must there-
fore be attributed to nonlinearity. This delayed effect can be explained by the usual vis-
coelastic behaviour of the rocks or by compaction of possible clay levels in the reservoir.
Due to the delay between reservoir depletion and the start of subsidence, long-term
monitoring of the displacements will be required after production stoppage and a subsidence
measuring system will therefore have to be maintained for many years.
Linear extrapolation
Field measurements
Subsidence
https://telegram.me/Geologybooks
Subsidence-depletion
delay
Prognosis based on
linear elasticity
Pressure depletion
Figure 3.2.4
Diagram of the subsidence-depletion delay effect [Hetema et al., 2002].
The main risks associated with subsidence are illustrated in the above-mentioned
examples. Subsidence is a serious problem for offshore platforms (reduced air gap) and near
seas and lakes (risk of flooding). In addition, Boutéca et al. (1996) point out that the risks
are highly dependent on the environment: ten centimetres subsidence is critical at Gronin-
gen, for example, but quite acceptable at other fields. Vertical subsidence is associated with
horizontal displacements which may damage buildings and surface facilities. The UK
National Coal Board classifies damage according to the size of the structure and the ampli-
tude of the displacements on the ground [Geertsma and Van Opstal, 1973]. Horizontal and
vertical strains between the surface and the reservoir may also fracture casings, destroy
numerous boreholes and reactivate faults.
A. Subsidence measurement
Boutéca et al. (1996) propose a review of subsidence and compaction measurement
methods. On the ground, subsidence is traditionally measured by levelling survey or GPS
positioning. More recently, Brink et al. (2002) used an InSAR (Interferometric Synthetic
Chapter 3 • Geomechanics and Reservoir 161
Aperture Radar) method to determine subsidence rates at the Lost Hills field in California.
This technique, based on processing images obtained using SAR radar on board satellite,
represents a powerful ground movement estimation tool. At sea, measurement can be carried
out less precisely by measuring the height variations between the platform and sea level.
Measurements of underground displacements in the reservoir and the overburden can also be
used. These measurements can be taken using conventional log analyses (measurement of
natural gamma radiation) or specialised log analyses (variation in casing lengths,
measurement using radioactive marker techniques, etc). For shallow fields, reservoir
compaction can be measured using a taut wire at the bottom of the well.
B. Subsidence prediction
Geertsma (1973) proposed an analytical model to provide an estimation of the subsidence
resulting from compaction of a cylindrical reservoir, of thickness h and radius R located at
depth z. Reservoir compaction is given by:
https://telegram.me/Geologybooks
h
C = Δh = ∫ cm ( z ) Δp( z )dz (3.2.1)
0
with cm the reservoir rock compressibility in oedometric condition (uniaxial condition and con-
stant vertical stress) which can be measured in laboratory and Δp the pressure variation in the
reservoir. If cm and Δp are homogeneous in the reservoir, the reservoir compaction is given by:
C = cm Δph (3.2.2)
Geertsma (1973) then gives the analytical expressions of the horizontal and vertical
elastic displacements on the surface in case of infinitely rigid underburden, overburden with
the same mechanical characteristics as the reservoir and reservoir with uniaxial behaviour.
The vertical displacement at the centre of the reservoir can be expressed as follows:
⎛ η ⎞ z
S = −2(1 − ν ) ⎜ 1 − ⎟ C with η = (3.2.3)
⎜⎝ 1 + η2 ⎟⎠ R
According to the latter equation, the ratio between maximum subsidence and reservoir
compaction (S/C) is determined by the ratio between depth of burial and the lateral extent of
the reservoir. Expression (3.2.3) shows in particular that small deeply buried reservoirs (large
h) will present no significant subsidences, even if their reservoir compaction is high. On the
contrary, large reservoirs at low depths (small h) may be potential candidates for subsidence.
Morita et al. (1989) later proposed a semi-analytical solution based on finite element cal-
culations, which does not assume that the reservoir and its confining formations have the
same mechanical properties.
The previous solutions are based on simplifying assumptions which do not allow accu-
rate calculation of the subsidence for a given field. In particular, the complex reservoir
geometry, heterogeneities in the reservoir and the surrounding formations and elastoplastic
constitutive laws must be taken into account to predict subsidence accurately. More accurate
162 Chapter 3 • Geomechanics and Reservoir
For the last century or so, human activity, in particular combustion of fossil hydrocarbons, has
increased the levels of greenhouse gases in the atmosphere. The CO2 concentration has increased
and could reach 1,000 ppm by the end of the century if no steps are taken to reduce emissions.
Experts agree on the fact that this increase is responsible for noticeable climate changes and vari-
https://telegram.me/Geologybooks
ability, with potentially disastrous impacts on the ecosystems and human activities.
As pointed out in the 2005 IPCC (Intergovernmental Panel on Climate Change) report,
Carbon Capture and Storage (CCS) is an efficient way of reducing CO2 emissions. If imple-
mented, the rate of greenhouse gas emission into the atmosphere could start to stabilise sig-
nificantly in the near future.
Storage in end-of-life oil or natural gas reservoirs offers several advantages, the most
important being that it is a known technique and virtually leaktight. These natural reservoirs
have proven their ability to store hydrocarbons for several millions of years. CO2 has already
been reinjected for many years in depleted reservoirs to reduce the oil viscosity, improve its
mobility and enhance recovery. While this is an EOR (Enhanced Oil Recovery) technique,
for sequestration of large quantities of CO2, the pressures considered exceed the pressures
that existed in situ before any human intervention, which is never the case with EOR.
CO2 storage in deep aquifers represents another possibility. Numerous deep aquifers
exist, onshore and offshore, some with areas of up to several thousand km2. Formed from
porous, permeable rocks, filled with brine unsuitable for human consumption, they could
store large quantities of CO2. Some are already used for this purpose (for example the Utsira
formation in the North Sea).
Hydraulic fracturing may be increased by the fact that CO2 injection will generate a
priori thermal perturbation with a cold front around the injectors. In some cases, fracturing
may propagate in the cap rock (see Chapters 2.4 and 2.5).
Pressure injection reduces the effective stresses normal to the faults and fractures and
therefore reduces their shear strength. The injector wells must be located far away from faults.
Shear stresses may be created at the top of the reservoir at the interface with the cap rock
[Hawkes et al., 2004].
3.2.2.2 Modelling
To quantify the risks, the most reliable approach is to conduct precise coupled geomechani-
cal-reservoir simulations. As when calculating subsidence, it will be essential to take into
account the geometry and mechanical properties of the reservoir and the surrounding forma-
tions.
Knowledge of the initial in situ stresses is also fundamental [Hawkes et al., 2005].
https://telegram.me/Geologybooks
The few studies already conducted indicate possible surface uplift in case of shallow
reservoir or aquifer (Figure 3.2.5) [Rutqvist et al., 2007, Vidal-Gilbert et al., 2009].
SMB15 SMB15
SMB17 SMB17
a) b)
U3 (m) U3 (m)
+5 = 810e-03 +2 = 054e-03 +4 = 409e-03 +2 = 709e-03
+5 = 814e-03 +1 = 428e-03 +4 = 126e-03 +2 = 425e-03
+4 = 558e-03 +8 = 019e-04 +3 = 842e-03 +2 = 142e-03
+3 = 932e-03 +1 = 759e-04 +3 = 559e-03 +1 = 859e-03
+3 = 306e-03 -4 = 500e-04 +3 = 276e-03 +1 = 575e-03
+2 = 680e-03 -1 = 076e-03 +2 = 992e-03 +1 = 292e-03
+2 = 054e-03 -1 = 702e-03 +2 = 709e-03 +1 = 008e-03
Figure 3.2.5
Variations of vertical displacement (m) a) at the reservoir top and b) at surface
during CO2 injection [From Vidal-Gilbert et al., 2009].
164 Chapter 3 • Geomechanics and Reservoir
Wells are drilled for various reasons: exploration, oil and gas production, water injection,
storage, etc. At some stage, all these wells will be permanently abandoned. Producer wells
have a lifetime of about twenty years, corresponding to the production period of an oil field.
After this phase, the well must be abandoned and plugged with suitable material, typically
cement-based. Even though the well has been plugged, there is a serious risk of leakage in
the long term which represents a threat to the environment. For example, migration of
hydrocarbons along the borehole may pollute the groundwater, the marine ecosystem or the
atmosphere in case of toxic gases. These leaks can be sealed, but repair costs are extremely
high for offshore installations. The main objective of oil well abandonment is therefore to
prevent the migration of any hydrocarbon reserves remaining at end of production by
limiting the flow of fluids along the borehole between the various geological layers.
Well closure and abandonment is therefore a critical issue regarding protection of the
environment and in particular protection of water resources. Since the techniques involved
in closure require a detailed analysis of the geological context, the well, variations in bot-
tomhole conditions expected over the long term, reservoir fluids, etc., each abandonment
procedure is unique.
Chapter 3 • Geomechanics and Reservoir 165
Typical
Typical «Best practices»
production well plug and abandonment
Drive
Bridge Plug No. 5
pipe
plug Surface isolation
Freshwater Plug No. 4
Isolate freshwater
Bridge
Saltwater
plug
Cut and pull
Surface
intermediate casing
casing
below freshwater base
Cement
retainer Plug No. 3
Intermediate Squeeze/isolate
casing noncemented annulus
Cut and pull production
casing near intermediate
Top of cement
casing shoe
Production
tubing Bridge
Plug No. 2
plug
Safety plug
Production
packer
Figure 3.2.6
Oil well abandonment procedure using cement plugs [Kelm and Faul, 1999].
166 Chapter 3 • Geomechanics and Reservoir
Installation of the well plug is a determining factor in closure quality. Wells are gener-
ally plugged with Portland cement to which sand and bentonite (for expanding cements) can
be added. Cement set accelerators can also be added, as well as dispersing agents to improve
the durability of the materials. For very large diameter wells, operators may also use con-
crete. With water wells, use of bentonite-based clay materials is sometimes recommended
[Calvert and Smith, 1994]. Presence of fluid in the well makes installation of the cement
plug difficult since it favours dilution and the creation of preferential flow paths of the
cement slurry [Smith et al. 1984]. The density difference between the cement slurry and the
fluid in the well may cause the cement slurry to flow to the bottom of the well, the result
being that no plug is really formed [Calvert and Smith, 1994, Crawshaw and Frigaard,
1999]. This type of flow can be avoided by installing a mechanical plug under the cement
plug, to ensure the cement slurry remains in place while setting. The installation procedure
must also ensure perfect adherence between plug and wall to avoid risks of leakage at the
interfaces. To improve adhesion between the casing and the cement plug, Evans and Carter
(1963) and Kelm and Faul (1999) recommend cleaning the tubing before installing the
https://telegram.me/Geologybooks
cement plug.
influence of this cooling phase to avoid excessive thermal expansion of the plug,
which could damage the primary cementation of the annulus (Figure 3.2.7).
0
-.3E7
-.6E7
-.9E7
-.12E8
-.15E8
-.18E8
-.21E8
-.24E8
-.27E8
-.30E8
-.33E8
-.36E8
-.39E8
-.42E8
-.45E8
-.48E8
-.51E8
-.54E8
-.57E8
-.60E8
Figure 3.2.7
Evolution of the radial stress around the well a) before cooling, b) after cooling
and c) after installing the plug and heating [Bosma et al., 2000].
https://telegram.me/Geologybooks
constitutive law in compression. In particular, this is the model used by Bosma et al.
(1999) and Ravi et al. (2002) to study the behaviour of the annuli cemented during the
well completion and production phases. In case of deep wells exposed to high
temperatures, the temperatures will modify the cement microstructure and therefore
affect the mechanical behaviour of the plug. It is therefore important to characterise
the mechanical behaviour of well plugs at high temperatures so that the plug
behaviour can be modelled during abandonment using suitable models.
4
400 years
3 150 years
125 years
q (MPa)
2
https://telegram.me/Geologybooks
100 years
1
75 years
50 years
0
-6 -4 -2 0 2
p' (MPa)
Figure 3.2.8
Example of tensile load applied in the shale caprock above an injector well
[Mainguy et al., 2007]. Evolution of the deviatoric and effective mean stresses
from end of production to 400 years after abandonment.
Reservoir monitoring is a production technology designed to measure, check and predict the
reservoir performance. Monitoring generally includes measurements of temperature, pres-
sure, flow rate and constituents, as well as surface deformation, 4D seismics and microseis-
micity which no longer concern the well alone but consider the reservoir as a whole with the
surrounding formations. These records can be used to map the progression of a heat front in
3D, identify reactivation of a fracture or fault and check the integrity of the cap rock.
10
-10
-20
https://telegram.me/Geologybooks
-30
-40
-50
200 m
Figure 3.3.1
Surface deformation over a period of 3.5 months as computed from tiltmeter
measurements on a SAGD field [Dubucq et al., 2008].
“Precision” GPS instruments and interferometric aperture radar techniques can also be
used to measure surface displacements, but with less accuracy.
The InSAR (Interferometric Synthetic Aperture Radar) technique uses two synthetic
aperture radars simultaneously, or the same radar used at different times. The point to point
phase differences in the images generated are then studied to find the altitude of the land.
Centimetric displacements can be measured in areas where the signal remains coherent [Mei
and Frosses, 2007].
The seismic technique is the geophysical method most commonly used to define sub-surface
structures. The most common implementation of this technique consists in reflection seismic
with multiple coverage. This method provides imaging of the sub-surface in 2 or 3
dimensions. Lithology and fluid information can be extracted from seismic data (wave
propagation velocities, amplitudes, etc.).
Chapter 3 • Geomechanics and Reservoir 171
Time-lapse seismic reservoir monitoring is the process of acquiring and analysing multi-
ple seismic surveys, repeated at the same site over calendar time, in order to image fluid-
flow effects in a producing reservoir. If each survey is “3-D seismic,” then the resulting set
of time-lapse data is termed “4-D seismic,” where the extra fourth dimension is calendar
time [Lumley, 1995].
In addition to 4-D seismic, there are other viable methods of time-lapse seismic monitor-
ing including repeated 2-D surface seismic, surface-to-borehole VSP (vertical seismic pro-
filing), and crosswell tomography.
Figure 3.3.2 shows two seismic cross-sections acquired at different times during the pro-
duction of the Duri field, Indonesia: before steam injection and after 5 months of steam
injection. The last seismic cross-section is the difference between the two previous ones.
The well outline visible in the third cross-section is due to a shallow borehole steam
leak.The time lapse seismic images show some changes in amplitudes along baseline reflec-
tors, due to a variation in fluid content.
https://telegram.me/Geologybooks
B
Time (s)
Time (s)
Time (s)
35
40
45
50
55
60
30
35
40
45
50
55
60
30
5
60
3
5
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
0.
Figure 3.3.2
Seismic cross-sections: (left) before steam injection, (centre) during injection,
(right) difference between the two cross-sections. S: steam zone, B: borehole
heating, P: polarity reversal [Lumley, 1995].
172 Chapter 3 • Geomechanics and Reservoir
Similarly, Figure 3.3.3 shows the maps of amplitude differences between the seismic
surveys carried out in 2001, 2004 and 2005 on the Christina Lake field operated by SAGD.
The amplitude differences indicate the changes in the steam chamber around the wells.
4D:2001-04 4D:2001-05
a) 2004 b) 2005
Figure 3.3.3
https://telegram.me/Geologybooks
Maps of amplitude difference a) between 2001 and 2004 and b) between 2001
and 2005 [Zhang et al., 2005].
The images provided by reservoir seismic monitoring may help identify by-passed oil
volumes, which will then be drilled. The reserves extracted can therefore be increased,
thereby extending the economic life of an oil field.
4D seismic can be used to track the progress of the injected fluid front (water, gas, steam,
CO2, etc.) and may play a role in optimising the injection programs.
More recently, time-lapse seismic has been applied for objectives other than monitoring
saturation changes, such as pressure monitoring and compaction detection. Over the past
few years, seismic monitoring has been widely used to track fluid movements (saturation
effect). This approach is no longer sufficient and pressure effects must be taken into account
to improve the interpretation of 4D seismic data [Landro, 1999]. More precisely, the pres-
sure effect does not concern the pore pressure alone, but also the induced geomechanical
effects.
Decreasing the pore pressure in the reservoir will increase the mean effective stress; this
will lead to rock compaction and increase the P-wave velocity Vp in the reservoir. The
overburden and underburden will deform to fill the space created by reservoir compaction.
Vp may be reduced in these zones due to the arching effect, a phenomenon observed by
Hatchell et al. (2004) on a North Sea field. The velocity variations in the cap rock were even
greater than those observed in the reservoir.
Coupled geomechanical-reservoir simulations can be used to determine the porosity, in
situ stresses and saturations which will then be substituted in the Gassmann equation to cal-
culate the P- and S-wave velocity field not only in the reservoir but also in the entire block
considered. The studies conducted in this field are highly promising [Vidal et al., 2002,
Minkoff et al., 2004].
Chapter 3 • Geomechanics and Reservoir 173
Confining stress
https://telegram.me/Geologybooks
Energy release
Pore
pressure
Shear stress
Figure 3.3.4
Microseismic failure mechanism.
Initially observed in Colorado after injection of fluid waste for underground disposal in
the 1960’s, induced microseismicity soon became a subject of interest to oil producers in
order to map hydraulic fractures, the first publications dating back to the 1970’s.
In the following years, microearthquakes were observed in the highly depleted oil zones,
especially the Groningen (the Netherlands), Gazli (Uzbekistan) and Lacq (France) fields.
Since these were naturally seismic risk areas, implementation of monitoring was in some
respects expected to distinguish between natural seismicity and induced seismicity. Surface
monitoring was carried out, using seismology equipment.
In the 1980’s, Passive Seismic Monitoring (PSM) became widely used in the mining
industry to prevent the risk of cave-in, especially in South Africa where extraction of various
ores (coal as well as coal and diamonds) is carried out at greater and greater depths. In the
field of underground storage, this technique has been used to monitor storage in salt cavities.
In these fields, the measurement means are traditional and inexpensive, consisting of
geophones connected to recording systems. The frequencies used ranges from a few Hz (less
in seismology) to several hundred Hz. The most recent developments take into account
three-component sensors.
174 Chapter 3 • Geomechanics and Reservoir
In the oil industry, the depth of the reservoir may limit the use of this technique due to
signal attenuation particularly in the near surface.
Ideally, information derived from interpretation of microseismic data can be used to:
– locate possible sliding over a natural fault plane,
– locate the compacting zones,
– locate the hydraulic fracturing zones and determine the characteristics (width, length,
height, orientation, growth, complexity, etc.) of the fractures created,
– identify the natural fractures or faults contributing to fluid displacements,
– map the fluid displacements,
– locate the zones badly drained or not drained by the fluids injected,
– identify the mechanically active levels likely to play a role in well failure (casing),
– map the thermal fronts, etc.
“Short-term” applications mainly concern injection operations that are highly localised in
time and space. The operator expects interpretation of the measurements immediately after
the test, for hydraulic fracturing or injection control.
Monitoring of hydraulic fracture propagation (see 2.4.5) depends on the instruments
implemented (number of sensors) and the knowledge available on wave propagation veloci-
ties in the different layers.
Injection control consists in checking that injection of a given fluid (production water,
wastewater, industrial fluids, drilling mud, steam, CO2, etc.) in a reservoir or an
underground aquifer carried out in limited proportions (limited injection duration) proceeds
correctly. Under these conditions, the operator does not want to fracture the geological units
surrounding the reservoir layer or the aquifer, to avoid polluting the other layers.
Carried out near the injection zone, PSM can track possible fracturing of the rocks or
reactivation of natural fractures or faults. If this type of phenomenon is observed, injection
can be contained in the initial layer by simply adapting the injection parameters. On the con-
trary, if there is no microseismic activity and no brutal injection pressure change, there is a
strong likelihood that the rock has not been fractured.
To date, “long-term” or permanent applications are operational in mining and storage of
hydrocarbons in saline cavities. Their main role is to ensure the safety of people and
property. These applications are also implemented for safety considerations in two other
cases: oil production where permanent surface measures have already been imposed on
some operators (especially NAM in the Netherlands) and storage of nuclear waste (NIREX).
In traditional oil and gas applications, long-term PSM proves interesting for gas storage
in underground aquifers [Vidal et al., 2000; Deflandre et al., 2002], monitoring of oil pro-
duction, monitoring of steam chamber propagation in thermal production [Snell et al.,
1999], reinjection of production water and more recently CO2 sequestration.
Long-term applications can be expected to play an increasing role in meeting restrictions
on protection of the environment [Snell et al., 1999]. It is in fact likely that PSM will even-
tually be implemented in oil well abandonment procedures.
Chapter 3 • Geomechanics and Reservoir 175
Wellhead sensor
Microseismic
Permanent downhole source
geophones
Ds = (D P, Dq)
Figure 3.3.5
Principle of passive seismic monitoring with downhole instruments.
176 Chapter 3 • Geomechanics and Reservoir
clock and an electronic module equipped with a pressure sensor and a temperature probe.
This system allows bottomhole fracturing temperature and pressure logging during and
after injection and real-time logging of a large number of events due to the proximity of the
emission zone. In addition, proximity of the emission zone reduces the uncertainty on the
Mono-conductor
Logging cable
BOP
Tubing Motor
Electronic
module Two arms latch assembly
Downhole pressure
Digital transmission
Mono conductor
logging cable Downhole temperature
Packer Decoupling
Casing device Decoupling control
Microseismic sensor
source
Permanent magnet
Fractured zone
Figure 3.3.6
Diagram of the SIMFRAC logging probe during a fracturing experiment.
Chapter 3 • Geomechanics and Reservoir 177
location of microseismic events and produces a more accurate map of the structure. Lastly,
system installation costs are low since the tubing does not have to be pulled up to take the
measurement.
In the frac-treated well, however, the microseismicity activity emitted during injection
cannot be analysed due to the noise level in the well. The probe can only be used during the
minifracturing phase preceding injection of the proppant.
Acoustic probes equipped with three geophones can be lowered down on a cable in a
cased well not equipped with a production column. Coupling is provided by the opening of
two motor-driven arms.
The big advantage with well permanent seismic sensors is that measurements can be
taken without affecting production. This advantage will be increasingly important when
operating deep sea wells due to their specificities (oil type, low pressure and temperature).
Comprehensive systems cover the requirements of conventional well seismics and pas-
sive seismic monitoring.
Tubing
Analogic links
Decoupling
device Bow-spring
for coupling
Frame 3-component
geophone
module
Digital
transmission to
deeper levels
Figure 3.3.7
Well permanent seismic sensors [Laurent et al., 1999].
Introduction
At all times during the exploitation of an oil field, the engineers strive to optimise drilling
times, reduce uncertainties and production costs, and optimise the use of available data.
Geomechanics, which includes all thermohydromechanical phenomena, plays an impor-
tant role in every operation involved in the exploitation of hydrocarbons, from drilling to
production and right up to the time the wells are abandoned. Pressure changes in the reser-
voir modify the in situ stresses and cause strains not only to the reservoir but also to the
https://telegram.me/Geologybooks
entire sedimentary column. In turn, these stress modifications and strains change the stabil-
ity parameters of the walls of the wells to be drilled and the flow properties of the fluids.
The main purpose of this book is to provide engineers involved in oil field development
and petroleum engineering students, whether drillers or producers, the information they need
about the techniques and approaches currently used in geomechanics.
The book is divided into three broad sections.
The first section provides information on rock mechanics, describes the characterisation
of rocks in the laboratory and the modelling of their mechanical behaviour. This chapter also
describes how geomechanics determine the stresses applied to the material in place and why
they need to model the reservoirs and the surrounding material.
The second section deals with the role of geomechanics in the optimisation of drilling
(drillability and wellbore stability) and production. Sand production, which brings numerous
undesirable problems, can be predicted and in some cases controlled. Hydraulic fracturing
can be encouraged to stimulate production or on the contrary avoided so as not to reactivate
a fault, in case of CO2 sequestration.
The third section deals with the coupling between the production of fluids in the reser-
voir and the mechanical strains of the complete oil structure. The environmental aspects of
this interaction are discussed with respect to subsidence of the surface layers, CO2 seques-
tration and well abandonment. Geomechanical monitoring of reservoirs may prove essential
in some cases to optimise production or check cap rock integrity.
Warning! In rock mechanics the compressive stresses are positive, whereas in continuous
media mechanics, stresses are negative. The latter convention is adopted in the numerical models.
Each chapter will indicate whether the positive or negative compressive stresses are being used.
References
https://telegram.me/Geologybooks
Aadnoy, B.S. (1988) Inversion technique to determine the in situ stress field from fracturing data,
Paper SPE 18023 presented at the 63rd Annual Technical Conference and Exhibition, Houston,
Texas, October 2-5.
Aadnoy, B.S. and Chenevert, M.E. (1987) Stability of highly inclined boreholes, Paper SPE 16052
presented at the SPE/IADC Drilling Conference held in New Orleans, March 15-18.
Abaqus Analysis User’s Manual, Abaqus Version 6.4, Simulia, Dassault Systèmes, 2005.
Abass, H.H., Hedayati, S. and Meadows, D.L. (1996) Nonplanar fracture propagation from a horizontal
wellbore: experimental study, Paper SPE 24823, SPE Production and Facilities, pp. 133-137.
Addis, M. A. (1997) The stress-depletion response of reservoirs, Paper SPE 38720 presented at the
SPE Annual Technical Conference and Exhibition held in San Antonio, Texas, October 5-8.
Addis, M.A., Choi, X. and Cumming, J. (1998) The influence of the reservoir stress-depletion res-
ponse on the lifetime considerations of well completion design, Paper SPE 47289 presented at
the SPE/ISRM Eurock’98 held in Trondheim, Norway, July 8-10.
Adenot, F. (1992) Durabilité du béton: Caractérisation des processus physiques et chimiques de dégra-
dation du ciment, PhD. Thesis, Université d’Orléans.
Alfenore, J. (2001) How to control high productivity wells in deep water reservoirs, SPE Distinguished
Lecturer Program.
Al-Harthy, S.S., Jing X.D., Marsden, J.R. and Dennis, J.W. (1999) Petrophysical properties of
sandstones under true triaxial stresses I: Directional transport characteristics and pore volume
change. Paper SPE 57287 presented at the SPE Asia Pacific Improved Oil Recovery Conference
held in Kuala Lumpur, Malaysia, October 25-26.
Alonso, E. E., Gens, A. and Josa, A. (1990) A constitutive model for partially saturated soils, Géotech-
nique, Vol. 40, n°3, pp. 405-430.
Batzle, M. and Wang, Z. (1992) Seismic properties of pore fluids, Geophysics, Vol. 57, N° 11,
pp. 1396-1408.
Baumgarte, C., Thiercelin, M. and Klaus, D. (1999) Case studies of expanding cement to prevent
microannular formation, Paper SPE 56535 presented at the Annual Technical Conference and
Exhibition, Houston, Texas, October 3-6.
Bear, J. and Buchlin, J.M. (1991) Modelling and applications of transport phenomena in porous media,
Kluwer, Academic Publishers, Dordrecht.
Beaudoin, J.J. (2000) Calcium hydroxide in cement matrices: physico-mechanical and physico-
chemical contributions, In Materials Science of Concrete: Calcium Hydroxide in Concrete,
Special Volume, pp. 131-142.
180 References
Bemer, E., Boutéca, M., Vincké, O., Hoteit, N. and Ozanam, O. (2001) Poromechanics: from linear to
nonlinear poroelasticity and poroviscoelasticity, Oil & Gas Science and Technology – Rev. IFP,
Vol. 56, N° 6, pp. 531-544.
Bemer, E. and Lombard, J.-M. (2008) An experimental investigation of the evolution of rock porome-
chanical properties associated to chemical alteration processes, Proceedings of Third Internatio-
nal Conference on Coupled T-H-M-C Processes in Geo-systems (GEOProc 08), Lille, France.
Berryman, J. G. (1995) Mixture theories for rock properties, in Rock physics and phase relations – A
handbook of physical constants, edited by T. J. Ahrens, American Geophysical Union,
Washington, pp. 205-228.
Bévillon, D. (2000) Couplage d’un modèle de gisement et d’un modèle mécanique: Application à
l’étude de la compaction des réservoirs pétroliers et de la subsidence associée. PhD Thesis, Uni-
versité des Sciences et Technologies – Lille.
Biot, M. A. (1941) General theory of three-dimensional consolidation, Journal of Applied Physics,
Vol. 12, pp. 155-164.
Birch, F. (1961) The velocity of compressional waves in rocks to 10 kilobars, Part 2. J. Geophys. Res.,
66, pp. 2199-2224.
Boidy, E. (2002) Modélisation numérique du comportement différé des cavités souterraines, PhD The-
https://telegram.me/Geologybooks
Burger J., Gadelle, C.P. and Marrast, J.R. (1986) Development of a chemical process for sand control,
SPE 15410 presented at the SPE Annual Technical Conference and Exhibition, New Orleans,
Louisiana, October 5-8.
Byerlee, J.D. (1978) Friction of rocks, Pure and Applied Geophysics, Vol. 116, pp. 615-626.
Calvert, D.G. and Smith, D.K. (1994) Issues and techniques of plugging and abandonment of oil and
gas wells, Paper SPE 28349 presented at the Annual Technical Conference and Exhibition, New
Orleans, September 25-28.
Carde, C. (1996) Caractérisation et modélisation de l’altération des propriétés mécaniques due à la
lixiviation des matériaux cimentaires, PhD INSA Toulouse.
Carter, B.J., Desroches, J., Ingraffea, A.R. and Wawrzynek, P.A. (2000) Simulating fully 3D hydraulic
fracturing, In Modeling in Geomechanics, Ed Zaman, Booker and Gioda, Wiley Publishers.
Castagna, J.P., Batzle, M.L. and Eastwood, R.L. (1985) Relationships between compressional-wave
and shear-wave velocities in clastic silicate rocks, Geophysics, Vol. 50, N° 4, pp. 571-581.
Cerasi, P., Papamichos, E. and Stenebraten, J. F. (2005) Quantitative sand-production prediction: Fric-
tion-dominated flow model, Paper 94791 presented at the SPE Latin American and Caribbean
Petroleum Engineering Conference held in Rio de Janeiro, Brazil, June 20-23.
Chardac, O., Murray, D., Carnegie, A. and Marsden, J.R. (2005) A proposed data acquisition program
https://telegram.me/Geologybooks
for successful geomechanics projects, Paper SPE 93182 presented at the 14th SPE Middle East
Oil and Gas Show and Conference, Bahrain, March 12-15.
Charlez, P.A. (1991) Rock mechanics, Volume 1, Theoretical fundamentals, Editions Technip, Paris.
Charlez, P.A. (1994) Exemple de modèle poroplastique: le modèle de Cam-Clay, Ecole de Mécanique
des Milieux Poreux, Aussois, France, 30 Mai-3 Juin 1994.
Charlez, P.A., Lemonnier, P., Ruffet, C., Boutéca, M. and Tan, C. (1996) Thermally induced fractu-
ring: analysis of a field case in North Sea, Paper SPE 36916 presented the SPE European Petro-
leum Conference held in MiIan, Italy, October 22-24.
Charlez, P.A. (1997) Rock mechanics, Petroleum Applications. Editions Technip, Paris.
Chavez, J-C., Carruthers, J. and McMurdy, P. (2005) Water flooding efficiency in a scenario of
multiple induced fractures, an applied geomechanical study, Paper SPE 97526 presented at the
SPE International Improved Oil Recovery Conference in Asia Pacific held in Kuala Lumpur,
Malaysia, December 5-6.
Chen, G., Chenevert, M., Sharma, M. and Yu, M. (2003) A study of wellbore stability in shales inclu-
ding poroelastic, chemical, and thermal effects, Journal of Petroleum Science and Engineering,
Vol. 38, N° 3-4, June 2003, pp. 167-176.
Chen, X., Tan, C.P. and Detournay, C. (2003) A study on wellbore stability in fractured rock masses
with impact of mud infiltration, Journal of Petroleum Science and Engineering, Vol. 38, N° 3,
pp. 145-154.
Chi, J.M., Huang, R. and Yang, C.C. (2002) Effects of carbonation on the mechanical properties and
durability of concrete using accelerated testing method, Journal of Marine Science and Techno-
logy, Vol. 10, No. 1, pp. 14-20.
Chin, L.Y. and Thomas, L.K. (1999) Fully coupled analysis of improved oil recovery by reservoir
compaction, Paper SPE 56753 presented at the SPE Annual Technical Conference and Exhibi-
tion, Houston, Texas, October 3-6.
Christensen, O., Janbu, N. and Jones, M.E. (1988) Subsidence due to oil-gas production, Proceedings
of the International Conference on Behaviour of Offshore Structures (BOSS’ 88), Trondheim,
Norway, V1, pp. 143-157.
Cipolla, C.L. and Wrigth, C.A. (2002) Diagnostic techniques to understand hydraulic fracturing: what?
Why? and How?, Paper SPE 75359, SPE Production & Facilities, Vol. 17, N° 1, pp. 23-35.
Cook, J. M., Sheppard, M. C. and Houwen, O. H. (1991) Effects of strain rate and confining pressure
on the deformation and failure of shale. Paper SPE 19944 presented at the 1990 IADC/SPE Dril-
ling Conference held in Houston, Texas, February 27-March 2.
182 References
Cooper, G. A. and Hatherly, P. (2003) Prediction of rock mechanical properties from wireline data and
their use in drilling simulation, Paper SPE 83509 presented at the SPE Western Regional/AAPG
Pacific Section Joint Meeting held in Long Beach, California, May19-24.
Cossé, R. (1993) Basic of reservoir engineering, Editions Technip, Paris.
Coussy, O. (1991) Mécanique des milieux poreux, Editions Technip, Paris.
Coussy, O. (1995) Mechanics of Porous Continua, John Wiley & Sons Ltd., Chichester.
Coussy, O. (2004) Poromechanics, John Wiley & Sons Ltd., Chichester.
Crawshaw, J. P. and Frigaard, I. (1999) Cement plugs: stability and failure by buoyancy-driven mecha-
nism, Paper SPE 56959 presented at the Offshore Europe Conference held in Aberdeen, Scot-
land, September 7-9.
Cunningham, R.A. (1978) An empirical approach for relating drilling parameters, Paper SPE 6715,
Journal of Petroleum Technology, Vol. 30, N° 7, pp. 987-991, July 1978
Damjanac, B. and Detournay, E. (1995) Numerical modeling of normal wedge indentation in rocks, in
Rock Mechanics, Proceedings of the 35th U.S. Symposium, University of Nevada, Reno,
pp. 349-354. J.J.K. Daemen and R. A. Schultz, Balkema Editions, Rotterdam.
https://telegram.me/Geologybooks
Dangla, P., Malinski, L. and Coussy, O. (1997) Plasticity and imbibition-drainage curves for unsatura-
ted soils: A unified approach, Proc. of the 6th International Symposium on Numerical Models in
Geomechanics (NUMOGVI), Montréal, Quebec, July 2-4, Edited by S. Pietruszczak and G.
Pande. A.A. Balkema, Rotterdam, the Netherlands. pp. 141-147.
De Borst, R. and Nauta, P. (1985) Non-orthogonal cracks in a smeared finite element model. Enginee-
ring Computations, Vol. 2, N° 1, pp. 35-46.
De Sa, A. N., Tavares, A. F. C. and do Carmo Marques (1989) Gravel Pack in offshore wells, Proc. of
the 21th Annual Offshore Technology Conference, Houston, paper 6041.
De Waal, J.A. and Smits, R.M.M. (1988) Prediction of reservoir compaction and surface, Paper SPE
14214, SPE Formation Evaluation, Vol. 3, N° 2, pp. 340-346.
Deflandre, J-P. and Huguet, F. (2002) Microseismic monitoring on gas storage reservoirs: a ten-year
experience, Proc. 17th World Petroleum Congress, Rio de Janeiro, Brazil.
Deliac, E (1986) Optimisation des machines d’abattage à pics, PhD Thesis, Université Pierre et Marie
Curie.
Desroches, J. and Kurkjian, A.L. (1998) Applications of wireline stress measurements, SPE 48960
presented at the SPE Annual Technical Conference and Exhibition, New Orleans, September
27-30.
Desroches, J., Marsden, J.R. and Colley, N., M. (1998) Wireline open-hole stress tests in a tight gas
sandstone, Proceedings of the International Gas Research Conference, Cannes, France, Novem-
ber 6-9.
Detienne, J-L., Danquigny, J., Lacourie, Y. and Espy, M. (2002) Produced water re-injection on a low
permeability carbonaceous reservoir, Paper SPE 78482 presented at the 10th Abu Dhabi Interna-
tional Petroleum Exhibition and Conference, October 13-16.
Detournay, E. and Atkinson, C. (1991) Influence of pore pressure on the drilling response of PDC bits,
Rock Mechanics as a Multidisciplinary Science, Ed. by J.C. Roegiers, Balkema, Rotterdam,
pp. 539-547.
Detournay, E. and Defourny, P. (1992) A phenomenological model for the drilling action of drag bits,
International Journal of Rock Mechanics and Mining Sciences & Geomechanics, Vol. 29, N° 1,
pp. 13-23.
Detournay, E. (1996) Models for rock cutting, Proceedings of the Tool-Rock Interaction Workshop
held during the 2nd North American Rock Mechanics Symposium (NARMS-2), Montreal, June
22, Editions Balkema.
References 183
Dubucq, D., Chalier, G. and Xavier, J-P. (2008) Steam chamber monitoring using surface deformation
from radar (PS-InSAR) and tiltmeter measurements, Proceedings of the Canadian CSPG CSEG
CWLS Convention.
Dusseault, M. B. and Santarelli, F. J. (1989) A conceptual model for massive solids production in
poorly consolidated sandstones, Proceedings ISRM/SPE International Symposium on Rock at
Great Depth, Vol. 2, Balkema, pp. 789-798.
Dusseault, M, Geilickman, M. and Spanos, T. (1998) Mechanisms of sand production in heavy oils,
Proceedings 7th UNITAR Conference on Heavy Oil and Tar Sands, Beijing, N° 1998.123.
Economides, M.J., Watters, L.T. and Dunn-Norman, S. (1998) Petroleum well construction, John
Wiley & Sons.
Egermann, P., Bemer, E. and Zinszner, B. (2006) An experimental investigation of the rock properties
evolution associated to different levels of CO2 injection like alteration processes, Paper SCA
2006-34, Proceedings of the International Symposium of the Society of Core Analysts, Trond-
heim, Norway, September 12-16.
Eilers, L.H. and Root, R.L. (1974) Long-term effects of high temperature on strength retrogression of
cements, SPE, Paper 5028 presented at the 49th Annual Fall Meeting of the SPE of AIME,
Houston, Texas, USA, October 6-9.
https://telegram.me/Geologybooks
Elkins, L., Morton, D. and Blackwell, W. (1972) Experimental fireflood in a very viscous oil-consoli-
dated sand reservoir, Paper SPE 4086 presented at the Fall Meeting of the Society of Petroleum
Engineers of AIME, 8-11 October 1972, San Antonio, Texas.
Ellis, R.C. (1998) An overview of frac packs: a technical revolution (evolution) process, Journal of
Petroleum Technology, Vol. 50, N° 1, pp 66-68.
Enever, J.R., Yassir, N., Willoughby, D.R. and Addis, M.A. (1996) Recent experience with extended
leak-off tests for in situ stress measurements in Australia, Australian Petroleum Production and
Exploration Association Journal, pp. 528-535.
Engstrøm, F. (1991) Rock mechanical properties of danish North Sea Chalk, Proceedings of the Fourth
North Sea Chalk Symposium. Deauville, France, September 21-23.
Eronini, I. E., Somerton, W. H. and Auslander, D. M. (1982) A Dynamic model for rotary rock dril-
ling, Transactions of ASME, Journal of Energy Resources Technology, Vol. 104, N° 2, pp. 108-
120.
Evans, G.W. and Carter, L.G. (1963) Bonding studies of cementing compositions to pipe and forma-
tions, drilling and production practice 1962, American Petroleum Institute, New York.
Falconer, I.G., Burgess, T.M. and Sheppard, M.C. (1988) Separating bit and lithology effects from
drilling mechanics data, Paper SPE 17191 presented at the SPE/IADC Drilling Conference, Dal-
las, Texas, February 28 -March 2.
Fam, W.A. and Dusseault, M.B. (1998) Borehole stability in shales: a physico-chemical perspective,
Paper SPE 47301 presented at the SPE/ISRM Conference on Rock Mechanics in Petroleum
Engineering, Eurock’98, Trondheim, Norway, Balkema, Rotterdam, pp. 285-290.
Fear, M.J., Knowlton, R.H., Meany, N. and Warren, T.M. (1995) Experts discuss drill-bit design, field
performance, Journal of Petroleum Technology, 100-104, pp. 142-143.
Finol, A.S. and Sancevic, Z. A., (1995) Subsidence in Venezuela, in subsidence due to fluid
withdrawal, developments in petroleum science, edited by G. V. Chilingarian, E. C. Donaldson
and T.F. Yen, Elsevier Science, Amsterdam.
Fjær, E., Holt, R. M., Horsrud, P., Raaen, A. M. and Risnes, R. (1992) Petroleum related rock mecha-
nics, Elsevier, Amsterdam.
Franco, J.L.A., De La Torre, H.G., Ortiz, M.A.M., Wielemaker, E., Plona, T.J., Saldungaray, P. and
Mikhaltseva, I. (2005) Using shear-wave anisotropy to optimize reservoir drainage and improve
production in low-permeability formations in the North of Mexico, Paper SPE 96808 presented
at the SPE Annual Technical Conference and Exhibition held in Dallas, Texas, October 9-12.
184 References
Fung L. S. K., Buchanan, L., and Wan, R. G. (1992) Coupled geomechanical-thermal simulation for
deforming heavy-oil reservoirs. Proceedings of the CIM Annual Technical Conference, Calgary,
June 7-10.
Galloway, D.L., Hudnut, K.W., Ingebritsen, S.E., Phillips, S.P., Peltzer, G., Rogez, F. and Rosen, P.A.
(1998) InSAR detection of aquifer-system compaction and land subsidence, Antelope Valley,
Mojave Desert, California, Water Resources Research., Vol. 34, 2573-2585.
Galloway, D.L., Jones, D.R. and Ingebritsen, S.E. (1999) Land subsidence in the United States: U.S.
Geological Survey Circular 1182.
Garnier, A.J. and van Lingen, N.H. (1958) Phenomena affecting drilling rates at depth, Paper SPE
1097 presented at 33rd SPE Annual Meeting, Houston, in Petroleum Transactions, AIME,
Volume 216, 1959, pages 232-239.
Gassmann, F. (1951) Über die Elastizität Poröser Medien, Veirteljahrsschrift der Naturforschenden
Gesellschaft, in Zurich, 96, 1-23.
Gatto, P. and Carbognin, L. (1981) The Lagoon of Venice: natural environmental trend and man-
induced modification, Hydrological Sciences Bulletin, 26/4/12, pp. 379-391.
Gavito, D. G. (1996) A new rock strength model and its practical applications, Paper SPE 35322 pre-
sented at the International Petroleum Conference and Exhibition of Mexico City, March 5-7.
https://telegram.me/Geologybooks
Geertsma, J. (1973) A basic theory of subsidence due to reservoir compaction: the homogeneous case,
Verhandelingen van het Koninklijk Nederlands Geologisch Mijnbouwkundig Genootschap,
Vol. 28, pp. 43-62.
Geertsma, J., and Van Opstal, G. (1973) A numerical technique for predicting subsidence above com-
pacting reservoirs, based on the nucleus of strain concept, Verhandlingen van het Koniklijk
Nederlands Geologisch Mijnbouwkundig Genootschap, Vol. 28, pp. 63-78.
Geertsma, J. (1989) On the alert for subsidence, Agip Review, Vol. 6, pp. 39-43.
Geilickman, M. and Dusseault, M. (1997) Dynamics of wormholes and enhancement of fluid produc-
tion, Paper 97-09 presented at the 48th Annual Technical Meeting of the Petroleum Society in
Calgary, Canada.
Geilickman, M. Dusseault, M. and Dullien, F.A. (1994) Fluid saturated solid flow with propagation of
a yielding front, Proceedings of the SPE/ISRM Conference on Rock Mechanics in Petroleum
Engineering, Eurock’94, Delft, Balkema, Rotterdam.
Geoffroy, H. (1996) Etude de l’interaction roche-outil de forage: influence de l’usure sur les
paramètres de coupe, PhD thesis, Ecole Polytechnique, Palaiseau, France.
Geoffroy, H., Nguyen Minh, D. and Putot, C. (1997) Study on interaction between rocks and worn
PDC’s cutter, Proceedings of the 36th Symposium US on Rock Mechanics ISRM, New York.
Geoffroy, H., Nguyen Minh, D., Bergues, J. and Putot, C. (1998a) Frictional contact on cutters wear flat
and evaluation of drilling parameters of a PDC bit, Paper SPE 47323 presented at the SPE/ISRM
Conference on Rock Mechanics in Petroleum Engineering, Trondheim, Norway, July 8-10.
Geoffroy, H., Nguyen Minh, D., Maitournam, H., Bergues, J. and Putot, C. (1998b) Evaluation of dril-
ling parameters of a PDC bit, in Advances in Rock Mechanics, World Scientific Publishing Co
Ed., pp. 301-314.
Gérard, B. (1996) Contribution des couplages mécanique-chimie-transfert dans la tenue à long terme
des ouvrages de stockage des déchets radioactifs, PhD Thesis, Ecole Normale Supérieure,
Cachan, France.
Gidley, J.L., Holditch, S.A., Nierode, D.E. and Veatch, R.W. (1989) Recent advances in hydraulic
fracturing, Society of Petroleum Engineers, Richardson, TX, USA., Monograph SPE Vol. 12.
Glowka, D. A. (1989) Use of single-cutter data in the analysis of PDC bit designs: Part 1 – Develop-
ment of a PDC cutting force model. Part 2 – Development and use of the PDCWEAR computer
code, Journal of Petroleum Technology, August. Vol. 41, N° 8.
Glowka, D.A. (1987) Development of a method for predicting the performance and wear of PDC drill
bits, Sandia Report 86-1745, Sandia National Laboratories.
References 185
Green, S. J., Griffin, R. M. and Pratt, H. R. (1973) Stress-strain and failure properties of a porous
shale. Paper SPE 4242 presented at the SPE Drilling and Rock Mechanics Conference, Austin,
Texas, January 22-23.
Guéguen, Y. et Palciauskas, V. (1992) Introduction à la physique des roches, Editions Hermann, Paris.
Guéguen, Y. and Boutéca, M. (1999), Mechanical properties of rocks: Pore pressure and scale effects.
Oil & Gas Science and Technology – Rev. IFP, Vol. 54, N° 6, pp. 703-714.
Guillemot, J. (1986) Elements de géologie, Editions Technip, Paris.
Gutierrez, M. and Lewis, M. (1998) The role of geomechanics in reservoir simulation, Proceedings of
the SPE/ISRM Conference on Rock Mechanics in Petroleum Engineering, Trondheim, Norway,
Eurock’98, Trondheim, Norway, July 8-10, pp. 439-448.
Hanson, J.M. (1991) Pore pressure ahead of the bit, Paper SPE 21916 presented at SPE/IADC Drilling
Conference, Amsterdam, Netherlands, March 11-14.
Hatchell, P.J., van den Beukel, A., Molenaar, M.M., Maron, K.P., Kenter, C., Stammeijer, J.G.F., van
der Velde, J.J. and Syers, C.M. (2003) Whole earth 4D: reservoir monitoring geomechanics,
Expanded Abstracts of the 73rd Annual International Meeting of the Society of Exploration
Geophysicists, Dallas, pp. 1330-1333.
https://telegram.me/Geologybooks
Hawkes, C.D. and McLellan, P.J. (1997) A new model for predicting time-dependent failure of shales:
theory and application, Paper presented at the 48th Technical Meeting of the Petroleum Society
of CIM, Calgary, June, also published in the Journal of Canadian Petroleum Technology,
Vol. 30, N° 13, December 1999.
Hawkes, C.D., McLellan, P.J. and Bachu, S. (2005) Geomechanical factors affecting geological sto-
rage of CO2 in depleted oil and gas reservoirs, Journal of Canadian Petroleum Technology
vol. 44, N° 10, pp. 52-61.
Hebert, W.E. and Young, F.S. (1972) Estimation of formation pressure with regression models of dril-
ling data, Journal of Petroleum Technology, Volume 24, Number 1, pp. 9-15.
Heffer, K.J., Last, N.C., Koutsabeloulis, N.C., Chan, H.C.M., Gutierrez, M. and Makurat, A. (1994)
The influence of natural fractures, faults and earth stresses on reservoir performance – Geome-
chanical analysis by numerical modelling. North Sea oil and gas reservoirs – Norwegian Geo-
technical Institute (Revue), Vol. 192, pp. 201-211.
Hermansen, H., Landa, G.H., Sylte, J.E. and Thomas L.K. (2000) Experiences after 10 years of water-
flooding the Ekofisk Field, Norway. Journal of Petroleum Science and Engineering, Vol. 26,
N° 1-4, pp. 11-18.
Hettema, M.H.H., Schutjens P.M.T.M., Verboom B.J.M. and Gussinklo H.J. (1998) Production-indu-
ced compaction of sandstone reservoirs: The strong influence of field stress, Paper SPE 50630
presented at the European Petroleum Conference, 20-22 October 1998, The Hague, Netherlands.
Hettema, M., Papamichos, E. and Schutjens, P. (2002) Subsidence delay: field observations and analy-
sis, Oil & Gas Science and Technology – Rev. IFP, Vol. 57, N° 5, pp. 443-458.
Homand, S. (2000) Comportement mécanique d’une craie très poreuse avec prise en compte de l’effet
de l’eau: de l’expérience à la simulation, PhD Thesis, University of Lille.
Horsrud, P. (2001) Estimating mechanical properties of shale from empirical correlations, Paper SPE
56017, Journal SPE Drilling & Completion, Vol. 16, N° 2, pp. 68-73.
Horsrud, P., Holt, R., Sonstebo, E.F., Svano, G. and Bostrom, B. (1994) Time dependent borehole sta-
bility: laboratory studies and numerical simulation of different mechanisms in shale, Paper SPE
28060 presented at the Conference on Rock Mechanics in Petroleum Engineering, Eurock 94,
29-31 August, Delft, Netherlands.
Hu, T., Fournier, F., Royer, J.J. and Joseph, C. (2006) Impact of uncertainties in rock mechanical pro-
perties on subsidence evaluation, Proc. of the XIth International Congress of the International
Association for Mathematical Geology, University of Liège – Belgium.
186 References
Huang, W.S., Marcum, B.E., Chase, M.R. and Yu, C.L. (1998) Cold production of heavy oil from
horizontal wells in the Frog Lake field, Paper SPE 52636, Journal SPE Reservoir Evaluation &
Engineering, Vol. 1, N° 6, pp. 551-555.
Jaeger, J. C. and Cook, N. G. W. (1979) Fundamental of rock mechanics – Third Edition, Chapman
and Hall, London.
Jardine, S., Malone, D. and Sheppard, M. (1994) Putting a damper on drilling’s bad vibrations, Oilfield
Review, January 1994.
Johnson, K.L (1985) Contact mechanics, Cambridge University Press.
Johnson, J.P., Rhett, D.W. and Siemers, W.T. (1989) Rock mechanics of the Ekofisk reservoir in the
evaluation of subsidence, Paper SPE 17854, Journal of Petroleum Technology, Vol. 41, N° 7,
pp. 717-722.
Kadri, I. B. (1991) Abnormal formation pressures in post-eocene formation, Potwar Basin, Pakistan,
Paper SPE 21920 presented at the SPE/IADC Drilling Conference, Amsterdam, Netherlands,
11-14 March.
Keaney, G.M.J., Meredith, P.G. and Murell, S.A.F. (1998) Laboratory study of permeability evolution
in a ‘tight’ sandstone under non-hydrostatic stress conditions. Paper SPE 47265 presented at the
https://telegram.me/Geologybooks
Last, N., Plumb, R., Harkness, R., Charlez, P., Alsen, J. and McLean, M. (1995) An integrated
approach to evaluating and managing wellbore instability in the Cuisiana field, Colombia, South
America, Paper SPE 30464 presented at SPE Annual Technical Conference and Exhibition held
in Dallas, Texas, October 22-25.
Laurent, J., Deflandre, J.P, Dubois, J.C. and Huguet, F. (1999) Permanent Downhole Geophones: an
efficient approach for reservoir monitoring, Proceedings of the European Applied Research
Conference on Natural Gas 1999 (“Eurogas 99”) Bochum, Germany.
Lavergne, M. (1986) Méthodes sismiques, Editions Technip, Paris.
Lawn, B.R. and Wilshaw, T.R. (1975) Fracture of brittle solids, Cambridge University Press.
Lesso, J.R. and Burgess, T.M. (1986) Pore pressure and porosity from MWD measurements, Paper
SPE 14801 presented at the SPE/IADC Drilling Conference, Dallas, Texas, February 9-12.
Le Tirant, P. and Gay, L. (1972) Manuel de fracturation hydraulique, Editions Technip, Paris.
Lewis, R. W. and Schrefler, B.A. (1998) The Finite element method in the static and dynamic defor-
mation and consolidation of porous media, Second Edition, Wiley & Sons, New York.
Lin, W. (1981) Mechanical behavior of Mesaverde shale and sandstone at high pressure, Paper SPE
9835 presented at the SPE/DOE Low Permeability Gas Reservoirs Symposium, Denver, Colo-
https://telegram.me/Geologybooks
McLellan, P. (1994) Assessing the risk of wellbore instability in horizontal and inclined wells, Paper
HWC94-14 presented at the Canadian SPE/CIM/CANMET International Conference on Recent
Advances in Horizontal Well Applications, Calgary, Canada, March 20-23.
McLellan, P.J. and Cormier, K. (1996) Borehole instability in fissile, dipping shales, northeastern
British Columbia, Paper SPE 35634 presented at the SPE Gas Technology Symposium, Calgary,
Alberta, Canada, 28 April-1 May.
Mendelsohn, D.A. (1984) A review of hydraulic fracture modeling, part 1: general concepts, 2D
models, motivation for 3D modeling, Journal of Energy Resources Technology, Vol. 106, N° 3,
pp. 369-376.
Mendelsohn, D.A. (1984) A review of hydraulic fracture modeling, part II: 3D modeling and vertical
growth in layered rock, Journal of Energy Resources Technology, Dec 1984, Vol. 106, N° 4,
pp. 543-553.
Merle, H.A., Kentie, C.J.P., van Opstal, G.H.C. and Schneider, G.M.G. (1976) The Bachaquero study
– A composite analysis of the behavior of a compaction drive/solution gas drive reservoir,
Journal of Petroleum Technology, September, Vol. 28, N° 9, pp. 1107-1115.
Mindlin, R. D. (1949) Compliance of elastic bodies in contact, Journal of Applied Mechanics, Vol. 16,
https://telegram.me/Geologybooks
pp. 259-262.
Minkoff, S. E., Stone, C. M., Bryant, S. and Peszynska M. (2004) Coupled geomechanics and flow
simulation for time-lapse seismic modelling, Geophysics, Vol. 69, N° 1, pp. 200-211.
Mobach, E. and Gussinklo, H. J. (1994) In situ reservoir compaction monitoring in the Groningen
field, Proceedings of EUROCK’94, Rock Mechanics in Petroleum Engineering, A. A. Balkema
Publ., Rotterdam, The Netherlands, pp. 535-547.
Moricca, G. Ripa, G., Sanfilippo, F. and Santarelli, F.J. (1994) Basin scale rock mechanics: field
observations of sand production, Paper SPE 28066 presented at the SPE/ISRM Rock Mechanics
in Petroleum Engineering Conference Delft, The Netherlands, August 29-31.
Morita, N., Whitfill, D.L., Nygaard, O. and Bale, A. (1989) A quick method to determine subsidence,
compaction and in-situ stresses induced by reservoir depletion, Journal of Petroleum Techno-
logy, January 1989, SPE 17150, Vol. 41, N° 1, pp. 71-83.
Mouchet, J. P. and Mitchell, A. (1989). Abnormal pressure while drilling. Manuals techniques 2,
Boussens, France, Elf Aquitaine Editions.
Mouton, D.E., Wilton, B.S. and Ramos, G.G. (1998) Optimizing drilling strategies in a tectonic belt,
Pagerungan field, North of Bali, Paper SPE 39357 presented at the IADC/SPE Drilling Confer-
ence, Dallas, Texas, March 3-6.
Nguyen, J.P. (1993) Le forage, Editions Technip, Paris.
Nguyen Minh, D. (1974) Contribution à l’étude de la taille des roches, PhD thesis, University of Paris VI.
Noik, C. and Rivereau, A. (1999) Oilwell cement durability, Paper SPE 56538 presented at the SPE
Annual Technical Conference and Exhibition, Houston, Texas, October 3-6.
Nolte, K.G. (1979) Determination of fracture parameters from fracturing pressure decline, Paper SPE
8341 presented at the SPE Annual Technical Conference and Exhibition, Las Vegas, Nevada,
September 23-26.
Nolte, K.G. and Smith, M.B. (1981) Interpretation of fracturing pressures, Journal of Petroleum Tech-
nology, Vol. 33, N° 9, Sept 1981, pp. 1767-1775.
Nouri, A., Vaziri, H., Belhaj H. and Islam, R. (2004) Sand production control: a new set of criteria for
modeling based on large scale transient experiments and numerical investigation, SPE Jounal,
Vol. 11, N° 2, pp. 227-237.
Nouri, A., Vaziri, H., Kuru, E. and Islam, R. (2006) A comparison of two sanding criteria in physical
and numerical modeling of sand production, Journal of Petroleum Science and Engineering,
Vol. 50, N° 1, pp. 55-70.
References 189
Nur, A. and Murphy, W. (1981) Wave velocities and attenuation in porous media with fluids, Procee-
dings of the 4th International Conference on Continuum Models of Discrete Systems, Stoc-
kholm, pp. 311-327.
Onaisi, A., Durand, C. and Audibert, A. (1994) Role of hydration state of shales in borehole stability
studies, Paper SPE 28062 presented at the SPE/ISRM Conference on Rock Mechanics in Petro-
leum Engineering, Eurock’94, Delft, Netherlands.
Osorio, J. G., Chen, H.Y., Teufel, L. and Schaffer, S. (1998) A two-domain, fully coupled fluid-flow/
Geomechanical simulation model for reservoirs with stress-sensitive mechanical and fluid-flow
properties. Paper SPE 28062 presented at the SPE/ISRM Conference on Rock Mechanics in
Petroleum Engineering, Eurock’98, Trondheim, Norway, July 8-10.
Papamichos, E., Stenebraten, J., Cerasi, P., Lavrov, A., Vardoulakis, I., Fuh, G.H., Brignoli, M.,
Goncalves de Castro, C.J. and Havmoller, O. (2008) Rock type and hole failure pattern effects
on sand production, Proceedings of the 42nd US Rock Mechanics Symposium, San Francisco,
June 29-July 2, Paper ARMA 08-217.
Patron, E. (1999) Poromécanique des joints en graphite expansé matricé, PhD Thesis, Ecole Nationale
des Ponts et Chaussées, Paris.
https://telegram.me/Geologybooks
Parcevaux, P. A. and Sault, P. H. (1984) Cement shrinkage and elasticity: a new approach for a good
zonal isolation, Paper SPE 13176 presented at the 59th SPE Annual Technical Conference and
Exhibition, Houston, September 16-19.
Pavone, D. R. and Desplans, J. P. (1994) Application of sampling rate downhole measurements for
analysis and cure of stick-slip in drilling, Paper SPE presented at the SPE Annual Technical
Conference and Exhibition, New Orleans, Louisiana, September 25-28.
Peltier, B. and Atkinson, C. (1986) Dynamic pore pressure ahead of the bit, Paper SPE 14787, SPE
Drilling Engineering, Vol. 2, N° 4, pp. 351-358.
Perkins, T. K. and Weingarten (1988) Stability and failure of spherical cavities in unconsolitated sand
and weakly consolidated rock, Paper SPE 18244 presented at SPE Annual Technical Conference
and Exhibition, Houston, Texas, October 2-5.
Perkins, T.K. and Gonzalez, J.A. (1985) The effect of thermoelastic stressses on injection well fractu-
ring, Paper SPE 11332, SPE Journal, Vol. 25, N° 1, pp. 78-88.
Perrodon, A. (1989) Dynamique et méthode d’étude des bassins sédimentaires, Editions Technip,
Paris.
Pessier, R.C (1994) Different shales dictate fundamentally different strategies in hydraulics, bit selec-
tion and operating practices, Paper SPE 28322 presented at the 69th Annual Technical Confer-
ence and Exhibition, New Orleans, Louisiana, September 25-28.
Pessier, R. C. and Fear, M. J. (1992) Quantifying common drilling problems with mechanical specific
energy and a bit-specific coefficient of sliding friction, Paper SPE 24584 presented at the SPE
Annual Technical Conference and Exhibition, Washington, D.C., October 4-7.
Pistre, V., Kinoshita, T., Endo, T., Schilling, K., Pabon, J., Sinha, B., Plona, T., Ikegami, T. and John-
son, D. (2005) A modular wireline sonic tool for measurements of 3D (azimuthal, radial and
axial) formation acoustic properties, Proceedings of the 46th SPWLA Annual Logging Sympo-
sium held in New Orleans, Louisiana, United States, June 26-29.
Plona, T.J., Kane, M.R., Sinha, B.K., Walsh, J. and Viloria, O. (2000) Using acoustic anisotropy,
Proceedings 41th Annual SPWLA Logging Symposium.
Plumb, R.A. (1994) Influence of composition and texture on the failure properties of clastic rocks.
Proceedings of the Eurock 1994. Rock Mechanics in Petroleum Engineering, Balkema,
Rotterdam, pp. 13-20.
Plumb, R., Papanastasiou, P. and Last, N. (1998) Constraining the state of stress in tectonically active
settings, Proceedings of the SPE/ISRM Conference on Rock Mechanics in Petroleum Enginee-
ring: Eurok 98, Trondheim, July 8-10.
190 References
Plumb, R.A. and Hickman, S.J. (1985) Stress induced borehole elongation: a comparaison between the
four arm dipmeter and yhe borehole televiewer in the Auburn geothermal well, Journal of Geo-
physical Research, Vol. 90, N° B7, pp. 5513-5522.
Poland, J. F. and Lofgren, B. E. (1984) Case History No 9.13 San Joaquin Valley, California, USA, in
Poland, J.F., ed., Guidebook to studies of land subsidence due to ground-water withdrawal,
UNESCO Studies and Reports in Hydrology, N°. 40, pp. 263-277.
Putot, C.J.M., Guesnon J., Perreau P.J. and Constantinescu A. (2000) Quantifying drilling efficiency
and disruption: field data vs. theoretical model., Journal SPE Drilling & Completion, Vol. 15,
N° 2, pp. 118-125.
Putot, C.J.M. and Constantinescu, A. (1997) Modèle de foration fondé sur le couplage des effets
d’évacuation des déblais et de coupe de la roche, Symposium Saint-Venant, Editions ENPC,
Paris.
Putot, C. (1995) Modèle de foration fondé sur le couplage de la destruction de la roche et d’évacuation
des débris, Oil & Gas Science and Technology – Rev. IFP, Vol. 50, N° 4, pp. 491-516.
Putot, C.J.M., Perreau P.J. and Constantinescu A. (1998) Field data vs. Theoretical model to quantify
drilling efficiency and disruption, Paper SPE 50579 presented at the SPE European Petroleum
Conference, The Hague, Netherlands, October 20-22.
https://telegram.me/Geologybooks
Ramos, G.G., Katahara, K.W., Gray, J.D. and Knox, D.J.W. (1994) Sand production in vertical and
horizontal wells in a friable sanstone formation, North Sea, Paper SPE 28065 presented at the
SPE/ISRM Conference on Rock Mechanics in Petroleum Engineering, Eurock’94, Delft,
Netherlands.
Rasolofosaon, P. and Zinszner, B. (2002) Vérification expérimentale de la formule de Gassmann dans
les calcaires poreux, Oil & Gas Science and Technology, Rev. IFP., Vol. 57, N° 2, pp. 129-138
Rasolofosaon, P., and Zinszner, B. (2004) Laboratory petroacoustics for seismic monitoring feasibility
study, The Leading Edge, Vol. 23, N° 3, pp. 252-258.
Ravi, K., Bosma, M. and Gastebled, O. (2002) Safe and economic gas wells through cement design for
life of the well, Paper 75700 presented at the SPE Gas Technology Symposium, Calgary,
Alberta, Canada, April 30-May 2.
Reinecker, J., Heidbach, O., Tingay, M., Sperner, B. and Müller, B. (2005) The release 2005 of the
world stress map (available online at www.world-stress-map.org)
Renard, G., Nauroy, J.-F., Deruyter, Ch., Moulu, J.-C., Sarda, J.-P. and Le Romancer, J.-F (2000)
Production froide des huiles visqueuses, Oil & Gas Science and Technology – Rev. IFP, Vol. 55,
No. 1, pp. 35-66.
Rhett, D.W. and Teufel, L.W. (1992a) Effect of reservoir stress path on compressibility and permeabi-
lity of sandstones, Paper SPE 24756 presented at the SPE Annual Technical Conference and
Exhibition, Washington, D.C., October 4-7.
Rhett, D.W. and Teufel, L.W. (1992b) Failure criteria for high porosity North Sea Chalk. Proceedings
of the Fourth North Sea Chalk Symposium, Deauville, France, September 21-23.
Roscoe, K.H. and Burland, J.B. (1968) On the generalised stress-strain behaviour of ‘wet clay’, Engi-
neering Plasticity, Eds Heyman J. and Leckie, F.A., Cambridge University Press, pp. 535-609.
Rousset, G. (1988) Comportement mécanique des argiles profondes – Application au stockage de
déchets radioactifs, PhD Thesis, Ecole Nationale des Ponts et Chaussées.
Ruisten, H., Teufel, L.W. and Rhett, D. (1996) Influence of reservoir stress path on deformation and
permeability of weakly cemented sandstone reservoirs, Paper SPE 36535 presented at the SPE
Annual Technical Conference and Exhibition, Denver, Colorado, October 6-9.
Rutqvist, J., Birkholzer, J.T. and Chin-Fu Tsang, (2007) Coupled reservoir – geomechanical analysis
of the potential for tensile and shear failure associated with CO2 injection in multilayered
reservoir – caprock systems, International Journal of Rock Mechanics & Mining Sciences,
Vol. 45, pp. 132-143.
References 191
Santarelli, F.J., Dardeau, C. and Zurdo, C. (1992) Drilling through highly fractured formations: a pro-
blem, a model, and a cure, Paper SPE 24592 presented at the SPE Annual Technical Conference
and Exhibition, Washington, D.C., October 4-7.
Santarelli, F.J., Tronvoll J.T., M. Svennekjaer, Skeie H., Henriksen H. and Bratli R.K. (1998) Reser-
voir stress-path: The depletion and the rebound, Paper SPE 47350 presented at the SPE/ISRM
Conference on Rock Mechanics in Petroleum Engineering, Trondheim, Norway, July 8-10.
Sarda, J-P. (1984) Fracturation hydraulique, Rapport IFP 31911.
Sargand, S. M. and Hazen, G. A. (1987) Deformation behaviour of shales, International Journal Rock
Mechanics Mining Sciences, Vol. 24, No.6, pp. 365-370.
Schofield and Wroth (1968) Critical state soil mechanics, McGraw Hill, London.
Scholz, C. H. (1990) The mechanics of earthquakes and faulting, Cambridge University Press.
Schutjens, P.M.T.M. and de Ruig, H. (1996) The influence of stress path and permeability of an over-
pressurised reservoir sandstone: some experimental data, Physics and Chemistry of the Earth,
Vol. 22, N° 1-2, pp. 97-103.
Schutjens, P.M.T.M., Blanton T.L., Martin J.W., Lehr, B.C. and Baaijens, M.N. (1998) Depletion-
https://telegram.me/Geologybooks
Sylte, J.E., Thomas, L.K., Rhett, D.W., Bruning, D.D. and Nagel, N.B. (1999) Water induced compac-
tion in the Ekofisk Field, Paper SPE 56426 presented at the SPE Annual Technical Conference
and Exhibition held in Houston, Texas, October 3-6.
Theron, A., De Langre, E. and Putot, C. (2001) The effect of dynamical parameters on precession in
rotary drilling, Journal of Energy Resources Technology, Vol. 123, N° 3, pp. 181-186.
Tixier, M.P., Loveless, G.W. and Anderson, R.A. (1975) Estimation of formation strength from the
mechanical properties log, Paper SPE 4532-PA., Journal of Petroleum Technology, Vol. 27,
N° 3, pp. 283-293.
Tortike, W.S. and Farouq, A. (1993) Reservoir simulation integrated with geomechanics. Journal of
Canadian Petroleum Technology, Vol. 32, N° 5, pp. 28-37.
Tremblay, B., Sedgwick, G. and Forshner, K. (1996) Imaging of sand production in a horizontal pack
by X-ray computed tomography, Paper SPE 30248-PA, SPE Formation Evaluation, Vol. 11,
N° 2, pp. 94-98.
Tremblay, B., Sedgwick, G. and Forshner, K. (1996) Simulation of cold production in heavy oil reser-
voirs: wormhole dynamics, Paper SPE 35387 presented at the SPE/DOE 10th Symposium on
Improved Oil Recovery held in Tulsa, Oklahoma, April 21-24.
Tremblay, B., Oldakowski, K. and Settari, A. (1997) Geomechanical properties of oil sands at low
https://telegram.me/Geologybooks
effective stress, Paper 97-07 presented at the 48th Annual Technical Meeting of the Petroleum
Society in Calgary, Alberta, Canada.
Tremblay, B., Sedgwick, G. and Forshner, K. (1998) Modelling of sand production from wells on pri-
mary recovery, Journal of Canadian Petroleum Technology, Vol. 37, N° 3.
Tremblay, B., Sedgwick, G., Vu, D., Lillico, D., Jossy, E., Yuan, J-Y, Babchin, A. and Sawatzky, R.
(1998) Cold production in heavy oil reservoirs, Poster presented at the 15th World Petroleum
Congress, Beijing, China.
Vajdova, V., Baud, P. and Wong, T-F. (2004) Compaction, dilatancy and failure in porous carbonate
rocks, Journal of Geophysical Research, Vol. 109, N° B5, B05204.
Van Hasselt, J.P. (1992) Reservoir compaction and surface subsidence resulting from oil and gas-
production: a review of theoretical and experimental research approaches, Geologie En
Mijnbouw, Vol. 71, N° 2, pp. 107-118.
Van Oort, E., Nicholson, J. and D’Agostino, J. (2001) Integrated borehole stability studies: key to
drilling at the technical limit and trouble cost reduction, Paper SPE 67763 presented at the SPE/
IADC Drilling Conference, Amsterdam, Netherlands, February 27 -March 1.
Vardoulakis, I., Stavropoulou, M. and Papanastasiou, P. (1996) Hydro-mechanical aspects of the sand
production problem, Transport in porous media, Vol. 22, N° 2, pp 225-244.
Vaziri, H., Barree, B., Xiao, Y., Palmer, I. and Kutas, M. (2002) What is the magic of water in produ-
cing sand?, Paper SPE 77683 presented at the SPE Annual Technical Conference and Exhibi-
tion, San Antonio, Texas, September 29 -October 2.
Vaziri, H. (2004) Integrated soil and rock mechanics: a major step towards consistent and improved
sand production prediction and management, Paper 04-461 presented at the 6th North America
Rock Mechanics Symposium (NARMS), Houston, Texas, June 5-9.
Veeken, C.A. M., Davies, D. R., Kenter, C. J. and Kooijman, A. P. (1991) Sand production prediction
review: developing an integrated approach, Proceedings 66th Annual Technical Conference of
the SPE, Dallas, Paper SPE 22792 presented at the SPE Annual Technical Conference and Exhi-
bition, Dallas, Texas, October 6-9.
Vernik L., Bruno M. and Bovberg, C. (1993) Empiricial relations between compressive strength and
porosity of siliciclastic rocks. International. Journal of Rock Mechanics and Mining Science,
Vol. 30, No.7, pp. 677-780.
Vidal, S., Longuemare, P. and Huguet, F. (2000) Integrating geomechanics and geophysics for reser-
voir seismic monitoring, feasibility studies, Paper SPE 65157 presented at the SPE European
Petroleum Conference, Paris, October 24-25.
References 193
Vidal, S., Huguet, F. and Mechler, P. (2002) Characterising reservoir parameters by integrating seis-
mic monitoring and geomechanics, The Leading Edge, Vol. 21, pp. 295-301.
Vidal-Gilbert, S., Huguet, F., Assouline, L. and Longuemare, P. (2005) Hydromechanical modelling
of reservoir behaviour during underground gas storage exploitation, Proceeding of the 68th
European Association of Geoscientists and Engineers Conference & Exhibition, Madrid, Spain,
June 13-16.
Vidal-Gilbert, S., Nauroy, J-F. and Brosse, E., (2009) 3D geomechanical modelling for CO2 geologic
storage in the Dogger carbonates of the Paris Basin, International Journal of Greenhouse Gas
Control, Vol. 3, N° 3, pp. 288-299.
Vincké, O. (1994) An estimation of bulk moduli of sandstones as a function of confining pressure using
their petrographic and petrophysic description, Paper SPE/ISRM 28038 presented at the SPE
Conference on Rock Mechanics in Petroleum Engineering, Delft, Netherlands, August 29-31.
Vincké, O., Boutéca, M. J., Piau, J. M. and Fourmaintraux, D. (1998) Study of the effective stress at
failure, Proceedings of the Biot Conference on Poromechanics, Louvain-La-Neuve, Belgique,
September 14-16, Thimus et al Editions, Balkema, Rotterdam, pp. 635-639.
Wang, Z. (2000) Dynamic versus static elastic properties of reservoir rocks, in seismic and acoustic
velocities in reservoir rocks, SEG Geophysics Reprint Series, No 19.
https://telegram.me/Geologybooks
Ward, C.D., Coghil, K. and Broussard, M.D. (1994) The application of petrophysical data to improve
pore and fracture pressure determination in North Sea Central Graben HPHT wells, Paper SPE
28297 presented at the SPE Annual Technical Conference and Exhibition, New Orleans, Loui-
siana, September 25-28.
Wardlaw, H. W. R (1971) Optimization of rotary drilling parameters, PhD thesis, Texas University, Austin.
Warpinski, N.R., Moschovidis, Z.A., Parker, C.D. and Abou-Sayed, I.S. (1994) Comparison study of
hydraulic fracturing models test case: GRI staged field experiment N° 3, Paper SPE 275890,
SPE Production & Facilities, Vol. 9, N° 1, pp. 7-16.
Warren, T.M. (1984) Factors affecting torque for a roller cone bit, Paper SPE 11991-P.A., Journal of
Petroleum Technology, Vol. 136, N° 9, pp. 1500-1508.
Warren, T.M. and Smith, M.B. (1985) Bottomhole stress factors affecting drilling rate at depth, Paper
SPE 13381, Journal of Petroleum Technology, August, pp. 1523-1535.
Williamson, M.D., Murray, S. J., Atkins, W.S., Hamilton, T.A.P. and Copland, M.A. (1997) A review
of Zechstein drilling issues, Paper SPE 51182, SPE Drilling & Completion, Vol. 13, N° 3,
pp. 174-181.
Wong T-F., David C. and Zhu, W. (1997) The transition from brittle to cataclastic flow: mechanical
deformation, Journal of Geophysical Research, Vol. 102, N° B2, pp. 3009-3025.
Wright, C.A., Weijers, L., Davis, E.J. and Mayerhofer, M. (1999) Understanding hydraulic fracture
growth: tricky but not hopeless, Paper SPE 56724 presented at the SPE Annual Technical Con-
ference and Exhibition, Houston, Texas, October 3-6.
Wriput, D. and Zoback, M.D. (2000) Fault reactivation and fluid flow along a previously dormant
normal fault in the northern North Sea, Geology, Vol. 28, pp. 595-598.
Xu, W., Le Calvez, J. and Thiercelin, M. (2009) Characterisation of hydraulically-induced fracture
network using treatment and microseismic data in a tight-gas formation: a geomechanical
approach, Paper SPE 125237 presented at the SPE Tight Gas Completions Conference held in
San Antonio, Texas, June 15-17.
Yalamas, T., Nauroy, J.-F., Bemer, E., Dormieux, L. and Garnier, D. (2004) Sand erosion in cold
heavy oil production, Paper SPE 86949 presented at SPE International Thermal Operations and
Heavy Oil Symposium and Western Regional Meeting, Bakersfield, California, March 16-18.
Yale, D. P., Nieto, J.A. and Austin, S. P. (1995) The effect of cementation on the static and dynamic
mechanical properties of the Rotliegendes sandstone, In: Daemen, J.J.K. & Schulz, R. A. (ed.)
Proceedings of the 35th U.S. Symposium on Rock Mechanics, Balkema, Rotterdam,
pp. 169-175.
194 References
Yale, D.P. and Crawford, B. (1998) Plasticity and Permeability in Carbonates: Dependance on Stress
Path and Porosity, Paper SPE 47582 presented at the SPE/ISRM Conference on Rock Mecha-
nics in Petroleum Engineering, Trondheim, Norway, July 8-10.
Yamamoto, K., Gutierrez, M., T. Shimamoto and Maezumi (2000) Verification of a 3D hydraulic frac-
turing model against a field case, Paper SPE 59373 presented at the SPE/DOE Improved Oil
Recovery Symposium, Tulsa, Oklahoma, April 3-5.
Yudovich, A., Chin, L.Y. and Morgan, D.R. (1989) Casing deformation in Ekofisk, Paper SPE 17856,
Journal of Petroleum Technology, Vol. 41, N° 7, pp. 729-734.
Zhang, W., Youn, S. and Doan, Q. (2005) Understanding reservoir architectures and steam chamber
growth at Christina lake, Alberta, by using 4D seismic and crosswell seismic imaging, Paper
SPE 97808 presented at the SPE/PS-CIM/CHOA International Thermal Operations and Heavy
Oil Symposium, Calgary, Alberta, Canada, November 1-3.
Zhu, W. and Wong, T.F. (1997). The transition from brittle to cataclastic flow: permeability evolution,
Journal of Geophysical Research, Vol. 102, B2, Feb. 10, pp. 3027-3041.
Zimmerman, R. W. (1991) Compressibility of sandstones, Elsevier Science Publishers B.V.,
Amsterdam.
Zinszner, B. and Pellerin, F-M. (2007) A Geoscientist’s guide to petrophysics, Editions Technip, Paris
https://telegram.me/Geologybooks
Zoback, M.D., Barto, C.A., Brudy, M.O., Castillo, D.A., Finkbeiner, T., Grollimund, B.R., Moos,
D.B., Peska, P., Ward, C.D. and Wiprut, D.J. (2003) Determination of stress orientation and
magnitude in deep wells, International Journal of Rock Mechanics & Mining Sciences, Vol. 40,
pp. 1049-1076.
Main Symbols
https://telegram.me/Geologybooks
Abandonment 164
Abnormal pressure 165
Acid fracturing 117
Acoustic emission 27
Anelastic Strain Recovery 61
https://telegram.me/Geologybooks
Anisotropy 14
Aquifer 167
Arching effect 57 172
Axial permeability 139
Axial splitting 81
Axial vibrations 87
Biot’s coefficient 23 25 31
Biot-Gassmann 64
Borehole breakouts 53
Borehole imaging 128
Boundary conditions 68
Brazilian test 18
Breakdown pressure 58 119 120
Breakout 89 96
Brittle failure 26 28
Byerlee 62
Calliper 54
Cam-Clay 30 47
Carbonates 28
Carbonation 169
Carter’s model 123
Caving 93
Cement 165
Cement plugs 169
Cement slurry 166 168
CHOPS 99
https://telegram.me/Geologybooks
Cleaning problems 83
Clogging 84
Closure pressure 58
CO2 sequestration 135 162
Coal gasification 118
Compact growth 104
Compaction 141
Compression and shear wave velocities 64
Compressive failures 88
Conduction laws 36
Consolidation 155
Consolidation treatment 117
Core discing 86
Coupled geomechanical-reservoir simulations 147 172
Coupling 136
Coussy 32
Creeping formations 95
Cunningham 73
Cutter wear 76 77
Cuttings 71
Damage 140
Darcy 3 142
Decompacted 102
Depletion 136 151
Detournay and Defourny 75
Deviatoric stress 6
d-exponent 71
https://telegram.me/Geologybooks
Falconer 74
Fault transmissivity 149
Faults 63
Filter cake 132
Fluid loss coefficient 123
FORS 72
Frac-pack 115
https://telegram.me/Geologybooks
Hardening cap 29
Heavy oil reservoirs 99
Hollow cylinder test 111
Hydraulic conductivity 141
Hydraulic diffusivity equation 135
Hydraulic fracture propagation 174
Hydraulic fracturing 88 115 117
163 176
Hydrostatic loading 27
https://telegram.me/Geologybooks
Imaging data 54
In situ stress 52 53
Inclinometry 121
Initiation pressure 119
Instantaneous Shut In Pressure 58 120
Interferometric Synthetic Aperture Radar 160
Internal plugging 132
Iterative coupling 144
Jamming 84
Lagrangian porosity 3 26
Leakoff 119 120 125
Leak-Off Pressure 58
Leak-Off Test 59
Limestone 17
Loading path 21 138
Loss of integrity 166
Oedometric test 18
One-way coupling 144
Over Balanced Drilling 88
Overburden 52 142
Permeability 3
Plastic buckling 81
Plasticity 28
Plasticity criterion 48
Plug ageing 168
Poisson’s ratio 11
Pore overpressure 72
Poroelastic behaviour 38
Poroelasticity 41
Poroelastoplastic behaviour 45
Porosity 3
https://telegram.me/Geologybooks
SAGD 99 172
Softening behaviour 28
Solid matrix 41
Spalling 81 89
Specific heat capacity 34
Static moduli 11
Steam chamber 172
Steam injection 167 171
Step Rate Tests 59
Stick-slip 87
Strain-hardening 29 50
Strains 8
Stress path 8 137 151
Stress tensor 5
Subsidence 135 153 154
Tensile failures 88
Tensile strength 120
Thermal effects 130
U
https://telegram.me/Geologybooks
Washouts 95
Water injection 130
Watercut 101 116
Waterflooding 141
Weight On Bit 71
Wire-wrapped sand screens 114
Wormholes 101 104 164
Yield surface 30
Young’s modulus 11
https://telegram.me/Geologybooks